You are on page 1of 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/241457336

Simplified dragonfly airfoil aerodynamics at Reynolds numbers below 8000

Article  in  Physics of Fluids · July 2009


DOI: 10.1063/1.3166867

CITATIONS READS
84 10,777

2 authors:

David-Elie Levy Avraham Seifert

4 PUBLICATIONS   137 CITATIONS   
Tel Aviv University
145 PUBLICATIONS   5,016 CITATIONS   
SEE PROFILE
SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Improved performance and reduced noise of wind turbines View project

Feedback Flow Control View project

All content following this page was uploaded by Avraham Seifert on 05 March 2016.

The user has requested enhancement of the downloaded file.


PHYSICS OF FLUIDS 21, 071901 共2009兲

Simplified dragonfly airfoil aerodynamics at Reynolds numbers below 8000


David-Elie Levy and Avraham Seifert
School of Mechanical Engineering, Faculty of Engineering, Tel-Aviv University, Tel Aviv 69978, Israel
共Received 3 November 2008; accepted 9 June 2009; published online 7 July 2009兲
Effective aerodynamics at Reynolds numbers lower than 10 000 is of great technological interest
and a fundamental scientific challenge. The current study covers a Reynolds number range of
2000–8000. At these Reynolds numbers, natural insect flight could provide inspiration for
technology development. Insect wings are commonly characterized by corrugated airfoils. In
particular, the airfoil of the dragonfly, which is able to glide, can be used for two-dimensional
aerodynamic study of fixed rigid wings. In this study, a simplified dragonfly airfoil is numerically
analyzed in a steady free-stream flow. The aerodynamic performance 共such as mean and fluctuating
lift and drag兲, are first compared to a “traditional” low Reynolds number airfoil: the Eppler-E61. The
numerical results demonstrate superior performances of the corrugated airfoil. A series of low-speed
wind and water tunnel experiments were performed on the corrugated airfoil, to validate the
numerical results. The findings indicate quantitative agreement with the mean wake velocity profiles
and shedding frequencies while validating the two dimensionality of the flow. A flow physics
numerical study was performed in order to understand the underlying mechanism of corrugated
airfoils at these Reynolds numbers. Airfoil shapes based on the flow field characteristics of the
corrugated airfoil were built and analyzed. Their performances were compared to those of the
corrugated airfoil, stressing the advantages of the latter. It was found that the flow which separates
from the corrugations and forms spanwise vortices intermittently reattaches to the aft-upper arc
region of the airfoil. This mechanism is responsible for the relatively low intensity of the vortices
in the airfoil wake, reducing the drag and increasing the flight performances of this kind of
corrugated airfoil as compared to traditional low Reynolds number airfoils such as the Eppler
E-61. © 2009 American Institute of Physics. 关DOI: 10.1063/1.3166867兴

I. INTRODUCTION these mechanisms, limiting the aerodynamic performance.


Furthermore, unsteady flow features are initiated at low Rey-
Effective flight at Reynolds numbers lower than 10 000 nolds numbers and become more pronounced as the Rey-
共based on chord length c兲, is of current technological interest nolds number decreases. For example, harmonic unsteady
and a fundamental scientific challenge. The ability to operate laminar flow was found in the wake of the NACA 0012 at
very small and cheap flying systems in hostile environments Reynolds numbers lower than 60 000.6
is the main motivation for the development of miniature fly- The focus of the current study is on flight Reynolds
ing vehicles1 共MAVs兲. numbers, based on a representative wing chord length, in the
From the aerodynamic point of view, the small dimen- range 2000⬍ Re⬍ 8000. Contrary to flight at significantly
sions of such vehicles and the low velocities in which they higher Reynolds numbers, nature can provide guidance, with
are required to fly significantly reduce the lifting ability on proper understanding that will allow mimicking effective
one hand and amplify the viscous flow phenomena on the flight at these low Reynolds numbers. Many insects fly in the
other. These viscous phenomena generate high drag,2 signifi- Reynolds numbers range Re⬍ 20 000. Natural flight is char-
cantly reducing these small airplanes’ performance regarding acterized by unsteady aerodynamics 共flapping wings兲 that is
gliding and range factor 共CL/CD兲 as well as endurance factor able to enhance the lift and provide thrust. The flapping flight
共CL1.5 / CD兲 共Ref. 3兲 共where CL and CD are the aircraft lift is a primordial flight mode of insects and therefore the focal
and drag coefficients, normalized by the wing area兲. These point of many studies 共e.g., Ref. 7兲. Nevertheless, to simplify
factors express the distance and the loitering time, respec- the technological challenge of building miniature air ve-
tively, that a given configuration can achieve in glide. Fur- hicles, it is advantageous to avoid wing flapping. Therefore,
thermore, the rapid boundary layer momentum loss can our work focuses on analysis of the flow around fixed wings.
cause flow detachments from the wing even at very low in- Several insects are able also to glide, one of which is the
cidence. The characteristics of the separated regions depend dragonfly. Therefore the analysis of the flow features around
on the airfoil geometry, and could vary significantly between representative natural insect airfoil at glide can be a candi-
different airfoils. In many cases, these detachments may an- date flow model for inspiring fixed-wing engineering solu-
ticipate the airfoil stall; in other situations, they may reattach tions.
as a turbulent boundary layer through a local laminar sepa- The dragonfly flies at Reynolds numbers smaller than
ration, transition and reattachment, called “laminar 15 000 共Ref. 8兲 and can fly without significant wing motion,
bubble.” 4,5 Generally, the maximum lift is very sensitive to at a representative Reynolds number between 700 and 2400.9

1070-6631/2009/21共7兲/071901/17/$25.00 21, 071901-1 © 2009 American Institute of Physics

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-2 D.-E. Levy and A. Seifert Phys. Fluids 21, 071901 共2009兲

At a higher Reynolds number, it is hypothesized that the corrugation “valleys.” Furthermore, no spontaneous collapse
dragonfly actually flies without wing flapping.8 This assump- of lift or massive flow separation, characteristic to the burst
tion is acceptable due to the sufficiently low amplitude and of the laminar bubble, was found in either the corrugated or
frequency of the wing oscillations perpendicular to the flight the filled-smoothed airfoils. Sunada et al.16,17 found that the
direction, such as one wavelength of the wing motion ex- performances of finite wings investigated in a water tank at
tends forward a distance significantly larger than the repre- Re= 4000 are strongly affected by the leading edge vortices.
sentative wing chord. The dragonfly gliding is interesting In the above-mentioned studies, the importance of the vorti-
also from a biological point of view since this flight mode is ces created by the corrugations on the enhanced performance
a part of its natural flight repertoire.9 Furthermore, since our of the different corrugated airfoils was clearly indicated, but
aim is also to arrive at technologically viable insight that is the underlying physical mechanism was not revealed.
inspired by the dragonfly’s gliding flight and not only to More recently, a computational study of the flow around
investigate the natural dragonfly aerodynamics, the flapping a corrugated dragonfly airfoil at Reynolds numbers of 55 000
motion can essentially be neglected. and 68 000 was published.18 The flow around this corrugated
airfoil was simulated by an unsteady Navier–Stoke solver
and compared to experimental results. The flow was found
II. THE CORRUGATED DRAGONFLY AIRFOIL
inherently unsteady and dominated by 3D vortices particu-
The complex morphology of the dragonfly wing cross- larly at high angles of attack. However, the behavior of the
section 共airfoil兲 is very different from the traditional closed, flow at Reynolds numbers close to 60 000 is significantly
thick, and smooth airfoils of common low Re number air- different than those studied here, i.e., at Reynolds numbers
planes. It is a very thin and low cambered airfoil, character- below 8000. At these higher Reynolds numbers, a transition
ized by a complex morphology near the leading edge 共X / c from laminar to turbulent flow is expected, and fine-scale 3D
⬍ 30%, where X is the distance from the leading edge and c structures will appear in the flow.
is the chord兲. The leading edge is composed of strong veins An additional numerical analysis of a dragonfly airfoil
at top and bottom, joined by a thin membrane, with three was recently performed at a more relevant Reynolds numbers
rows of serrations.8 Downstream from the leading edge range, lower than 10 000.19 The flow was assumed to be
共termed Costa兲, the airfoil is corrugated. Note that this kind laminar and steady. The calculated lift to drag ratio of the
of airfoil is representative also of other flying insects, such as corrugated airfoil and the streamlined airfoil were found to
the locust.10 In addition to the clear structural advantages,11 be similar and sometimes slightly higher for the corrugated
the aerodynamic features of such corrugated airfoils at these airfoil. It was proposed that this slight advantage of the cor-
Reynolds numbers are still poorly understood. rugated airfoil is due to lower drag coefficients. It was at-
Simplified dragonfly airfoils and wing configurations tempted to explain this benefit by the appearance of negative
were tested and analyzed as fixed wings by many research- skin friction due to the reverse flow between the corruga-
ers. Rees12 compared the lift-drag polars of a wing with a tions. While this is qualitatively viable argument, it is orders
corrugated airfoil with these of a wing with a smoothed air- of magnitude smaller than the drag difference measured and
foil at Reynolds numbers lower than 900. These wings had calculated. Moreover, the authors did not consider the in-
an aspect ratio 关AR⫽共wing span length兲2/wing surface兴 of creased drag created by the momentum invested in the vor-
less than 1.12 therefore, three-dimensional 共3D兲 flow effects tices’ rotation and the alternation of the form drag. Therefore,
as wing-tip vortices should have a dominant role in the delay this argument does not explain the observations, considering
of the wing stall at higher angles of attack. Newman et al.8 only steady flows.
suggested that the corrugations provoke an early transition to
turbulent flow over the two-dimensional 共2D兲 airfoil, permit- III. MOTIVATION
ting the reattachment of the flow to the wing. The reattach-
ment of the flow allows the reduction of the drag coefficient As discussed above, corrugated airfoils provide superior
and also enhances the range factor 共CL/CD兲. However, the aerodynamic efficiency compared to “classical” smooth air-
Reynolds numbers of the above study were relatively high foils at the current range of Reynolds numbers. However, the
共Re⬎ 10 000兲, and this explanation cannot be extended to more detailed studies known to the authors at Reynolds num-
lower Reynolds numbers, where the flow is essentially lami- bers lower than 20 000 were essentially experimental and
nar. Buckholtz13 found, by flow visualization experiments, performed on finite wings. These measurements are affected
that laminar reattachment takes place downstream of the sta- by 3D effects, such as wing-tip vortices. Therefore, 2D air-
tionary separation bubbles caused by the airfoils corruga- foil properties cannot be isolated. Furthermore, only the in-
tions. Okamoto et al.14 isolated the influence of different tegral forces were measured and no detailed flow features
geometrical parameters and showed that the surface texture were presented. This is apparently due to the difficulty in
or the roughness of wings result in an increased maximum measuring properties such as pressures and velocities over
lift coefficient, CLmax and increased maximum lift-to-drag these very thin airfoils at very low speeds.
ratio 共L / D兲max. The experiments of Kesel15 compared the The purpose of the current study is to analyze the 2D
performances of corrugated and filled-smoothed airfoils on flow around a simplified corrugated dragonfly airfoil. First,
finite-wing models 共AR⬍ 4兲. Enhanced performances were the 2D flow around the corrugated airfoil and a known low
measured for the corrugated airfoils. Their efficiency was Reynolds number, airfoil 共Eppler-E61兲 are numerically ana-
assumed to be induced by the vortices residing in the airfoil lyzed. Their performances, such as lift, drag coefficients 共Cl

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-3 Simplified dragonfly airfoil aerodynamics Phys. Fluids 21, 071901 共2009兲

FIG. 1. 共a兲 Simplified corrugated airfoil geometry, 共b兲 Eppler-E61 airfoil FIG. 2. Grid geometry for numerical simulations: 共a兲 The general view of
geometry. The airfoils are plotted at ␣ = 0°, 共c兲 comparison of the two air- the calculation domain and the boundary conditions. 共b兲 The grid in the
foils shapes 共stretched four times in the vertical direction兲, and 共d兲 schematic proximity of the corrugated airfoil surface.
definition of the projected height 共h兲 on the Y axis.

and Cd兲, and lift-to-drag ratio are compared. In parallel, an the region close to the airfoil 共for convenience兲 while it is
experimental validation of the numerical results around the unstructured in the far field 关Fig. 2共b兲兴. The origin of the grid
corrugated airfoil is performed. Additionally, the above two is situated on the lower leading edge point of the airfoil
airfoils, and others, designed on the basis of the flow pattern geometry. The chordwise direction is represented by the X
around the corrugated airfoil, are numerically analyzed in axis and the chord-normal direction is the Z axis 共Z = 0 at the
order to understand the physical phenomena caused by the trailing edge兲. The grid extends up to 12 chords downstream
presence of the corrugations on this class of natural airfoils. from the airfoil trailing edge, 8 chords above and below the
airfoil, and 8 chords upstream of the leading edge 关Fig. 2共a兲兴.
The downstream boundary condition was defined as outlet
IV. A COMPARISON OF CORRUGATED AND
flow. At the other boundaries, the external free-stream veloc-
TRADITIONAL AIRFOILS
ity U⬁ was imposed. With this grid configuration, only a
The corrugated airfoil geometry 关Fig. 1共a兲兴 analyzed in change of 4% U⬁ was calculated at 0.3 chords downstream
this work is based on the simplified dragonfly airfoil of New- 共X兲 and 2.5 chords above 共Z兲 of the airfoil at Re= 8000 and
man’s gliders.8 Its lower surface coordinates are presented in at angle of attack ␣ = 10°. At Re= 8000 and one chord from
Table I. The airfoil thickness is 0.05 mm. Since one of the the leading edge, the characteristic viscous scale, based on
significant parameters of low Reynolds number aerodynam- the momentum thickness of a flat plate laminar boundary
ics is the airfoil thickness, a “classical” airfoil with a layer is of order of 6% c.2 Therefore, a first grid layer height
thickness-to-chord ratio close to that of the corrugated airfoil of 0.002 mm 共0.02% c兲 represents 0.3% of the characteristic
thickness ratio 关5.5% based on the corrugations height, see viscous scale at the highest Reynolds number of the current
Fig. 1共a兲兴 was identified and analyzed. The chosen airfoil is study. Grid convergence studies showed similar flow field
the Eppler-E61, shown in Fig. 1共b兲.20 Furthermore, the rear- solutions for a grid of 256 000 points as compared to those
upper part of the corrugated airfoil 共X / c ⬎ 70%, where X is calculated for a grid of 64 000 points. The more detailed grid
the free-stream direction coordinate and X = 0 at the leading was built by splitting the basic grid cells uniformly. In con-
edge兲 is similar to the aft-upper region surface of the Eppler- trast, a coarser grid of 32 000 points was found to be insuf-
E61 airfoil, emphasizing the role of the leading edge and ficient to describe the unsteady phenomena. Results of the
corrugations’ shape, and minimizing the differences caused grid sensitivity study are shown in Fig. 3. These data provide
by the geometry of the trailing edge region 关Fig. 1共c兲兴. examples of the mean pressure coefficient distribution 共C p兲
The Navier–Stokes equations around these airfoils were for two grids, the time dependent lift coefficient and fre-
numerically solved using the commercial code FLUENT. The quency spectrum at Re= 6000, which is close to the highest
grid which contains 64 000 points 共Fig. 2兲, is structured in Reynolds number currently analyzed. The dimensionless fre-

TABLE I. Coordinates of the lower surface of the simplified corrugated airfoil geometry 共the thickness is t = 0.05 mm兲.

x 共mm兲 0.00 0.90 1.40 1.90 2.40 2.90 3.40 5.00 6.00 7.00 7.50 8.00 8.50 9.00 9.50 10.00
y 共mm兲 0.00 0.00 0.50 0.00 0.50 0.00 0.00 0.20 0.60 0.80 0.85 0.80 0.70 0.55 0.40 0.25

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-4 D.-E. Levy and A. Seifert Phys. Fluids 21, 071901 共2009兲

FIG. 4. Comparison of vorticity in the calculated flow field around the


corrugated and Eppler-E61 airfoil at Re= 6000. The vorticity isolines range
of the gray scale extends −21 000⬍ ␻ ⬍ 21 000 rad/ s 共␻ ⬎ 21 000 rad/ s
are included in the white regions and ␻ ⬍ −21 000 rad/ s are included in the
black regions兲, at Clmean = 0.53 for 共a兲 corrugated airfoil 共␣ = 3.7°兲, 共b兲 E61
共␣ = 1.0°兲, 关共c兲 and 共d兲兴 vorticity at Clmean = 0.82 共␣ = 6.5° and ␣ = 4.0°, for the
corrugated and Eppler-E61 airfoils, respectively兲.

quency is represented here by the Strouhal number St, cal-


culated from the dominant oscillation frequency 共f兲, the ver-
tical projected length 共h兲 of the corrugated airfoil, and the
free-stream velocity 共U⬁兲. This comparison is followed by
the results of continuity and streamwise velocity conver-
gence history study 关Figs. 3共d兲 and 3共e兲兴 at the converged
state, which shows a similar convergence level. The results
for the 64 000 grid points are well converged, and only a
small decrease in the flow fluctuations dominant frequency
共about 1% smaller for the larger grid兲 can be noted 关Fig.
3共c兲兴. The time step was chosen to provide at least 20 points
per period of the lift/drag time histories. Laminar, 2D, un-
steady calculations were performed at Reynolds numbers
ranging from 2000 to 8000.
At Reynolds numbers ranging from 2000 to 8000, the
calculated flow field is fundamentally unsteady and vortices
are generated over the airfoils. The characteristic results are
similar for all the studied Re numbers and for all the range of
“low amplitude” Cl behavior. Therefore, analysis of repre-
sentative results will be presented only at characteristic flow
conditions, Re= 6000 with Clmean = 0.53 and Clmean = 0.82.
These points were chosen to be in the core of the studied
range of Reynolds numbers and lift coefficients, with suffi-
FIG. 3. Results of the corrugated airfoil numerical convergence study for cient lift to emphasize the phenomena. Figure 4 presents ex-
␣ = 6° and Re= 6000 for two different grids: 共a兲 mean pressure distributions, amples of the calculated vorticity fields at the above-
共b兲 time dependent lift coefficient, 共c兲 normalized frequency spectrum, 共d兲
continuity convergence evolution, and 共e兲 streamwise velocity convergence mentioned flow conditions, where ␻ is the spanwise
evolution 共the peak appear at the beginning of new time step convergence component of vorticity measured in rad/s.
iterations兲. The alternating creation and shedding of vortices into the

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-5 Simplified dragonfly airfoil aerodynamics Phys. Fluids 21, 071901 共2009兲

FIG. 6. Calculated mean lift-incidence curve and their extreme values time
variations at Re= 6000 for 共a兲 the corrugated airfoil and 共b兲 Eppler-E61
airfoil.
FIG. 5. Comparison of calculated lift and drag coefficients vs time at Re
= 6000, 共a兲 Clmean = 0.53 共␣ = 3.7° for the corrugated airfoil and ␣ = 1° for the
Eppler E-61兲 ; 共b兲 Clmean = 0.82 共␣ = 7° for the corrugated airfoil and ␣ = 4°
for the Eppler E-61兲. oscillations of the Eppler-E61 airfoil are gradually increasing
with the angle of attack. Nevertheless, at the low angles of
attack range 共␣ ⬍ −0.7° associated with Cl⬍ 0.31 for Re
wake 共as shown in Fig. 4兲 cause the lift and drag coefficients = 6000兲 a similar sudden transition to a different vortex shed-
to be time periodic with a dominant harmonic content, as ding mode appears 关Figs. 6共b兲 and 7兴, which is associated
shown in Fig. 5. Therefore, the lift and drag coefficients can with flow separation from the lower surface of the Eppler-
be described by their mean values 共Clmean and Cdmean, re- E61 airfoil.
spectively兲 and their standard deviations 关␴共Cl兲 and ␴共Cd兲, The standard deviation of the Cl and Cd time histories
respectively兴. In Fig. 5, the time 共t兲 histories of the lift and are plotted as a function of the mean-lift coefficients at Re
drag coefficients for the two airfoils 共corrugated and Eppler- = 6000 in Figs. 7共a兲 and 7共b兲. These graphs demonstrate the
E61兲 are compared at two lift coefficients and at Re= 6000. sudden transition to a different vortex shedding mode at high
At Clmean = 0.53 关Fig. 5共a兲兴, the amplitudes of the lift signals Cl for the corrugated airfoil, while this process is gradual for
are similar for the two airfoils. However, at Clmean = 0.82 the Eppler-E61 airfoil. Furthermore, the data shown in Fig. 7
关Fig. 5共b兲兴 the amplitudes of the lift fluctuations acting on the clearly demonstrate that one of the advantages of the corru-
Eppler-E61 airfoil are larger 共by a factor of about 5兲 than gated airfoil is its capability to reduce the integral force os-
those acting on the corrugated airfoil. The magnitude of the cillations acting on the airfoil over a wide range of mean-lift
lift fluctuations for the corrugated airfoil actually decreases values. A possible explanation for this phenomenon is illus-
for the larger lift condition 关Fig. 5共b兲兴. The same trend can be trated in Fig. 8 关and Fig. 9共b兲兴, which presents an example of
observed for the drag signals. The dependence of the lift the variations in the pressure distributions around the two
amplitude on ␣ are presented in Fig. 6 for Re= 6000, where airfoils at Re= 6000 and Clmean = 0.82. The data are for times
the mean lift 共thin line兲 and the extreme lift values 共thick corresponding approximately to the minimum 共T0 + 1 / 2T兲
lines兲 are plotted as a function of ␣. In the calculated range and to the maximum 共T0兲 of the corresponding Cl variations,
of ␣, the lift curve of the corrugated airfoil can be separated where T0 is an initial time and T is the duration of one cycle
into two main regions: The first which extends up to an in- of a periodic function. A time harmonic variation of C p ap-
cidence of ␣ ⬇ 7°, where the oscillations of Cl as a function pears around 40% of Eppler-E61 chord and its wavelength
of time are small and nearly constant; and the second range, extends over the rear 60% of the airfoil. The C p over the
above ␣ ⬇ 7°, where Cl oscillations suddenly grow. It is hy- corrugated airfoil is also highly unsteady. However, a travel-
pothesized that this growth is due to a production of a dif- ing wave appears at 50% of the corrugated airfoil chord, but
ferent vortex shedding mode, with larger amplitude and its wavelength is merely 20% of the chord, with a wave
lower frequencies. Note that both extreme lift values are still velocity around 0.27U⬁ in the presented case. The convec-
growing, producing significant lift oscillations while the tive nature of the corrugated airfoil waves and their relatively
mean-lift force are concomitantly grow. In contrast, the Cl short wavelength diminishes the integral force variations

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-6 D.-E. Levy and A. Seifert Phys. Fluids 21, 071901 共2009兲

FIG. 9. An example of corrugated airfoil high angles 共␣ ⬎ 8°兲 vortex shed-


ding mode, at Re= 6000, ␣ = 9° 共a兲 C p distribution variations in half period
共T / 2兲 increment, 共b兲 dependence of the lift coefficients on time at ␣ = 9° 共up
graph兲 and ␣ = 6.5° 共low graph兲.

FIG. 7. Comparison between the standard deviations of the calculated Cl


and Cd fluctuations vs the mean-lift coefficients for the corrugated and the
Eppler-E61 airfoils at Re= 6000. around the airfoil. Such a traveling pressure wave also allows
the reattachment 共partial at least兲 of the flow over the aft part
of the corrugated airfoil and consequently, a stronger pres-
sure recovery.
At high angles of attack 共e.g., ␣ ⬎ 7°兲 the vortex shed-
ding that develops in the airfoil wake has greater magni-
tudes, translated to larger amplitudes of the Cl oscillations
and a decreasing in the dominant frequency. These differ-
ences are emphasized in Fig. 9共b兲, in what the Cl oscillations
behavior at ␣ = 9° is compared to this at ␣ = 6.5°. Character-
istic pressure distributions and integrated lift coefficients are
time dependent, and are presented in Fig. 9共a兲 for this vortex
shedding mode, at ␣ = 9°. In this case, the time harmonic
variation of C p develops closer to the leading edge 关note the
significantly enhanced C p variations at X / c ⬇ 0.5 in Fig. 9共a兲
as compared to Fig. 8共b兲兴. Furthermore, its wavelength in-
creases significantly from about 20% to about 40% of the
corrugated airfoil chord. In addition, the amplitude of the
pressure wave grows at least threefold, leading also to in-
creased mean-lift coefficient, as previously seen in Fig. 6.
Lift and drag polar plots of the different airfoils were
calculated for Re= 6000 and the mean values are presented in
Fig. 10. The mean-lift curve of the Eppler-E61 airfoil is ap-
proximately linear in the range of −0.7° ⬍ ␣ ⬍ 8° with a re-
duction in the mean-lift slope at the higher angles of attack.
The features of this lift curve are similar to those of known
airfoils at higher Reynolds numbers. However, the mean-lift
curve of the corrugated airfoil is not linear, with a range of
FIG. 8. Calculated C p distributions time variations at two times 共T is the
dominant period兲, Re= 6000, Clmean = 0.82, 共a兲 Eppler-E61, and 共b兲 corru- slope larger than 2␲ 共e.g., 2.6␲ rad−1 at 0 ° ⬍ ␣ ⬍ 2°兲, con-
gated airfoil. trary to that of the Eppler 61, which is lower than 2␲ rad−1

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-7 Simplified dragonfly airfoil aerodynamics Phys. Fluids 21, 071901 共2009兲

FIG. 10. Calculated mean-lift curves and mean lift-drag polar comparisons
at Re= 6000.

共0.92␲ rad−1 at 0 ° ⬍ ␣ ⬍ 2°兲 in all the “low amplitude” lift


fluctuations range. The slope of the lift curve 共Cl␣兲 can ex-
ceed the potential limit of 2␲ rad−1 at very low Reynolds
numbers, when local detachments, as laminar separation
bubble, appear on the airfoil surface and change the effective
airfoil camber as previously shown, for example, for the
NACA 0012.21 Currently the larger than 2␲ rad−1 slope is
caused by the separation bubbles that progressively grow FIG. 11. The mean boundary layer shape 共based on isovelocity contour of
50% of the free stream velocity兲 for the corrugated and the Eppler-E61
with ␣ in the forward part of the corrugated airfoil, between
airfoils at ␣ = 1° for 共a兲 Re= 2000, 共b兲 Re= 8000, and at ␣ = 5°, 共c兲 Re
and downstream of the corrugations increasing the effective = 2000, and 共d兲 Re= 8000.
airfoil camber 共Fig. 11兲. A similar phenomenon was identi-
fied on a corrugated airfoil by Buckholtz.13 The lift-drag po-
lar comparisons show lower mean-drag values for the corru-
gated airfoil for mean-lift coefficients between 0.0 and 1.1.
The data of Fig. 11 clearly indicate the reattachment of the this transition is abrupt for the corrugated airfoil and gradual
separated shear layer to the curved aft region 共arc兲 of the for the Eppler-E61 airfoil, over the entire Reynolds number
corrugated airfoil. range currently studied. The development of the Cl ampli-
tude of both airfoils is expressed by the Cl standard devia-
tions that are presented in figure Fig. 14. It can be seen that
V. THE EFFECT OF REYNOLDS NUMBER
at Re= 2000 and low angles of attack the flow around the two
The effect of Reynolds numbers, between 2000ⱕ Re airfoil is steady; indicated by ␴共Cl兲 ⬇ 0. Increasing the Rey-
ⱕ 8000, on mean-lift curves and mean-drag polars are sum- nolds number causes the Cl amplitude to grow; however the
marized in Fig. 12. Similar to high Reynolds number aero- growth rate decreases with increasing Re. In all the calcu-
dynamics, the mean drag increases when Re number de- lated Reynolds numbers cases, the range of the incidence
creases. Moreover, in this range of Reynolds numbers the angles for which low amplitude Cl oscillations existed over
zero-lift coefficient increases with Re. The thinning of the the corrugated airfoil is wider than the incidence range of
boundary layers around the airfoil 共as shown in Fig. 11兲 which low amplitude oscillations was observed over the
seems to cause an effective increase in the airfoil camber Eppler-E61 airfoil. As previously noted, in this range 共0
with ␣, resulting in overall higher lifting capability. ⱕ Clⱕ 0.9兲, the Cl amplitudes are nearly constant for the
The transition from one mode of vortex shedding to a corrugated airfoil, contrary to what was observed for the cor-
second mode at high angles of attack, is illustrated for the responding Cl range of the Eppler-E61 airfoil for which the
two extreme Reynolds numbers that were studied, Re lift oscillations gradually, but persistently, increase with
= 2000 and Re= 8000, in figure Fig. 13. As previously shown, Clmean.

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-8 D.-E. Levy and A. Seifert Phys. Fluids 21, 071901 共2009兲

FIG. 13. The extreme values of lift coefficients as a function of the angle of
attack for 共a兲 the corrugated airfoil and 共b兲 the Eppler-E61 airfoil.

共normalized by the wing area兲 at Re= 3000 are considered.


The body drag coefficient, CDb, normalized by the char-
acteristic wing area, was assumed to vary as a function of Re
as indicated in Table II. The 3D lift coefficient 共CL兲 and the

FIG. 12. The effect of Reynolds number on lift curves and lift-drag polars of
the 共a兲 corrugated and 共b兲 Eppler-E61 airfoils.

VI. EVALUATION OF THREE-DIMENSIONAL


PERFORMANCE
Practically, the performances of the corrugated airfoil
must be distinguishable also in a wing-body configuration. A
rectangular wing with a constant airfoil section and a repre-
sentative dragonfly body were considered to evaluate this
effect. The 3D dragonfly configuration performances are
evaluated, by adding 3D aerodynamic effects, such as lift
induced drag and body parasitic drag, to the 2D airfoil per-
formances. For this purpose a 3D configuration, including
characteristic dragonfly geometry of wing and body, were
defined based on published data.14 Specifically, a wing with FIG. 14. The effect of Reynolds number on standard deviations of Cl vs
AR= 11 and a body with a drag coefficient of CDb = 0.05 Clmean for the 共a兲 corrugated and 共b兲 Eppler-E61 airfoils.

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-9 Simplified dragonfly airfoil aerodynamics Phys. Fluids 21, 071901 共2009兲

TABLE II. The assumed drag variation of a characteristic dragonfly body as


function of Re and the Reynolds number based on body length Reb.

Re Reb CDb

2000 16 000 0.055


4000 32 000 0.050
6000 48 000 0.045
8000 64 000 0.040

total 3D drag coefficient 共CD兲 of a characteristic dragonfly


configuration were evaluated by the following formula:22
Clmean
CL = , 共1兲
1 + Cl␣/␲AR

CL2
CD = Cdmean + CDb + , 共2兲
␲eAR
where the value of the induced drag efficiency, e = 0.85, was
arbitrarily chosen, as a conservative value. The 3D lift and
drag estimations are principally derived using potential flow
considerations. Therefore, the viscous phenomena, even if
important, are not supposed to affect the validity of these
formula, and are in all the cases taken into account in the
changes that they cause to the lift curve slope 共Cl␣兲.
The range and endurance performances of these kinds of
vehicles with wings comprised of two different airfoils 共cor- FIG. 15. 共a兲 Calculated mean-lift to mean-drag ratio and 共b兲 endurance
rugated and Eppler-E61兲 were compared in the range of Cl in coefficients comparisons for a 3D flying configuration at different Reynolds
which the vortex shedding amplitudes are low 共Fig. 15兲 for a numbers.
range of Reynolds numbers. The range and endurance per-
formances of the corrugated airfoil configuration are superior
a vertical, open-loop water tunnel with a cross section of
for the CL and Reynolds ranges analyzed as compared to the
22⫻ 21 cm2 and a 70 cm long test section.23 The experi-
configuration with the Eppler-E61 airfoil. However, the dif-
ments were performed on a wing model with a 2D corru-
ference between the two wings is more important at the in-
gated airfoil shape, with a chord of 10 cm and a thickness of
termediate Reynolds number range studied. If at Re= 8000
0.05 cm, mounted between the test section sidewalls. For the
and CL⬇ 0.75 the wing with the corrugated airfoil has an
flow visualization study, two holes, for dye injection, were
advantage of 7.5% in range and 6.0% in endurance, these
placed on the center line of the wing: one on the summit of
advantages increase to 10.0% in both range and endurance at
the first corrugation and the second on the rear arc at 78%c
Re= 4000. From an engineering point of view, the lower drag
from the leading edge. In the cross-flow 共Y兲 direction, two
of the corrugated airfoil wing becomes more advantageous
additional dye injections holes were placed on the peak of
as the body and induced drag decrease. Therefore a better
the first corrugation at a distance of 2.5 cm from each side of
designed body or wing with higher AR will enhance the
the center line 共⫾0.25c, Fig. 16兲. In all the experiments, the
presently calculated performances. These findings demon-
strate the advantages of the fixed, rigid corrugated airfoil for
flight at Reynolds numbers between 2000 and 8000 as com-
pared to conventional airfoils.

VII. EXPERIMENTAL VALIDATIONS


To validate the numerical results, an experimental study
was performed. The above discussed calculations are based
on two main assumptions: the first is that the flow is laminar
around the airfoil and in the near wake; the second is that the
perturbations caused by the corrugations do not disturb the
2D nature of the flow on a sufficiently large aspect ratio wing
at Re= 2000 to Re= 8000. In order to qualitatively validate
these assumptions, a flow visualization study was carried out
in the water tunnel of the Faculty of Aerospace Engineering FIG. 16. A sketch of the wing plan form configuration for water tunnel flow
at the Technion, Israel Institute of Technology, Haifa. This is visualization experiments. Flow from top to bottom.

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-10 D.-E. Levy and A. Seifert Phys. Fluids 21, 071901 共2009兲

FIG. 19. Visualization of flow fields around the corrugated airfoil at Re


FIG. 17. Visualization of flow fields around the corrugated airfoil at Re = 8000 for ␣ = 0°, ␣ = 4°, and ␣ = 8° 共top down, respectively兲.
= 4000, for ␣ = 0°, ␣ = 4°, and ␣ = 8° 共top down, respectively兲.

dye was injected over the upper surface of the corrugated over and in the wake of the airfoil is laminar. The flow vi-
airfoil and photographed with a 10 Mpixel digital camera sualization images show that increasing the Reynolds num-
with an 18–200 mm lens. The airfoil was lighted by a pro- ber leads to a reduction in the wavelength of the shed vorti-
jector from a direction perpendicular to the wing plane in ces, i.e., increasing its Strouhal number. In Fig. 18, the dye
order to accentuate the dye color contrast from the ambient streak line structures formed in the flow downstream of the
medium. first corrugation are compared to the streak lines computed
At the beginning of the flow visualization study, the vor- from the numerical simulation data, with simulated “tracer”
tices created by the corrugations were visualized, by photo- released from the first corrugation peak at Re= 6000. Good
graphing the injected dye in the flow field perpendicular to agreement between experiment and calculations can be ob-
the wing section plane. Figures 17–19 present the flow streak served in this case too. Vortices are clearly seen to be shed
lines at Re= 4000, 6000, and 8000 for angles of attacks 0°, and reattach to the curved region. Most of these vortices
4°, and 8°. The unsteady nature of the flow past the corruga- travel downstream but a fraction travel upstream 共or at least
tions can be identified in all the images. Furthermore, these the associated dye兲 into the separated region between first
pictures clearly confirm that the vortex system developed corrugation and the rear arc.
The 2D character of the flow in the water tunnel was
visualized by photographing the streak lines perpendicular to
the wing span and these images are shown in the Fig. 20. The
three parallel streak lines exiting from the first peak corruga-
tion holes demonstrate the 2D character of the flow over the
wing at Re= 2000, 4000, and 8000 and ␣ = 0° and 8°. The
only exception, perhaps, is for the case Re= 8000 and ␣
= 8° in which the flow seems to develop mild three-
dimensionality downstream of 60% c. Overall, these flow
visualization experiments justify the use of laminar, 2D flow
simulations for all the current range of Reynolds numbers
and angles of attack.
A quantitative validation of the numerical results was
conducted at the low-speed, low-turbulence wind tunnel at
Tel-Aviv University. The closed-loop wind tunnel has test
section dimensions of 61 cm width, 91 cm height, and 6 m
long. 2D experiments were performed on a wing model with
a 2D corrugated airfoil shape. The chord length was 10 cm
and its thickness of 0.5% 共0.5 mm兲 of the chord. The model
was mounted between the wind tunnel sidewalls. Data were
acquired by one or two single velocity component hot-wire
sensors, mounted on a three dimensional traversing system.
To demonstrate that the flow in the wind tunnel is two
dimensional, mean velocity profiles at a distance of 18 cm
downstream of the airfoil trailing edge were measured at
three spanwise stations 共Y = 4.1c, 3.3c, and 2.7c from the side
of the tunnel兲. The Reynolds numbers varied from 2000 to
12 000, and the angle of attack varied between 0° and 10°.
FIG. 18. Visualization of the flow fields around the corrugated airfoil at
Re= 6000, for ␣ = 0°, ␣ = 4°, and ␣ = 8° 共top down, respectively兲. Compari- The representative measurements for ␣ = 6°, that are pre-
son of visualizations to computed streak lines. sented in Fig. 21, show a good correlation between the mean

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-11 Simplified dragonfly airfoil aerodynamics Phys. Fluids 21, 071901 共2009兲

FIG. 21. Experimental 2D flow validations: velocity profiles measured in


FIG. 20. Flow field visualization showing the 2D character of the flow over the wake 共1.8c from the trailing edge兲 of the corrugated airfoil at three
the wing. Re= 2000 共a兲 a = 0°, 共b兲 a = 8°; Re= 4000 共c兲 a = 0°, 共d兲 a = 8°; spanwise stations at ␣ = 6°, for 共a兲 Re= 2300, 共b兲 Re= 6200, and 共c兲 Re
Re= 8000, 共e兲 a = 0°, and 共f兲 a = 8°. = 12 200.

velocity profiles along the spanwise direction over the entire mentum deficit along the vertical axis was performed for all
Reynolds number range that was tested. This finding indi- the experimental and calculated data. The integration of each
cates the 2D nature of the mean flow in the wake. Clearly, if velocity profile will result in the momentum thickness calcu-
the wake flow is two dimensional, so would be the flow over lation in the streamwise flow direction normalized by c, at
the airfoil. the corresponding Reynolds number and angle of attack, ac-
Thereafter, these mean wake velocity profiles, as a func- cording to:
tion of the vertical position in the wake, Z, measured as
above at 1.8c 共18 cm兲 downstream from the airfoil trailing ␪=
1
c
冕冉
Z

0
1− 冊
U U
U⬁ U⬁
dZ. 共3兲
edge, were compared to the results obtained from the nu-
merical simulations 共Fig. 22兲. In Fig. 22, the origin of the In the above, c is the chord length, Z is the vertical coordi-
vertical Z axis is the vertical position of the airfoil trailing nate, U is the local streamwise velocity, U⬁ is the free-stream
edge. A good correlation exists between the numerical and velocity, and ␪ is the momentum thickness normalized by c.
the experimental results for the mean velocity profiles. To The normalized momentum thickness comparison of ex-
strengthen the comparison, integration of the streamwise mo- periment and simulation is presented in Fig. 23. A good cor-

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-12 D.-E. Levy and A. Seifert Phys. Fluids 21, 071901 共2009兲

FIG. 22. Comparison of numerical and experimental wake velocities mea-


sured at x / c = 2.8 from the leading edge in the wake of corrugated airfoil:
␣ = 6°, 共a兲 Re= 4000, 共b兲 Re= 6000, and 共c兲 Re= 8000.

FIG. 23. Comparison of computed and measured momentum thickness of


the corrugated airfoil at 共a兲 Re= 2000, 共b兲 Re= 4000, 共c兲 Re= 6000, and 共d兲
Re= 8000.
relation between tested and simulated normalized momen-
tum thickness was generally obtained in the range of
fh fc sin共␣兲
Reynolds numbers and angles of attack measured. St = ⬎ . 共4兲
The instantaneous velocities were also recorded across U⬁ U⬁
the airfoil wake 共at X = 1.8c downstream from the trailing In Eq. 共4兲, h is the airfoil projected height 关which is greater
edge and between 0.1c ⬍ Z ⬍ 0.3c above it, depending on the than c sin共␣兲 due to the corrugations height兴 on the Z axis, as
Reynolds number and the angle of attack兲. Fourier analysis defined in Fig. 1共d兲.
of these velocity signals was performed and the typical Figures 24–27 present the comparisons between the ex-
dominant frequencies that develop in the wake were ob- perimental results and numerically obtained Strouhal num-
tained. The dominant nondimensional frequencies 共St兲 as a bers at three different Reynolds numbers, shown on the left
function of ␣ measured in the experiment were compared to side of the figures 关marked by 共a兲兴. The lower parts of these
those computed by Fluent, where the Strouhal number is figures 关marked by 共b兲 and 共c兲兴 present a comparison of the
defined as measured and computed spectral components, at representa-

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-13 Simplified dragonfly airfoil aerodynamics Phys. Fluids 21, 071901 共2009兲

FIG. 25. A comparison of numerical and experimental wake velocities spec-


tral content for corrugated airfoil at X / c = 1.8: Re= 4000. 共a兲 Frequencies
FIG. 24. A comparison of numerical and experimental wake velocities spec- comparison. Spectra comparison at 共b兲 ␣ = 6° and 共c兲 ␣ = 2°.
tral content for corrugated airfoil at X / c = 1.8: Re= 2000. 共a兲 Frequencies
comparison. Spectra comparison at 共b兲 ␣ = 6° and 共c兲 ␣ = 2°.

and ␣, see Figs. 28共b兲 and 29共a兲兴. This concept increases the
tive incidence angles of ␣ = 2°, and 6°. These graphs demon- maximum thickness of the airfoil in comparison to the cor-
strate good agreement between the experimental and the cal- rugated airfoil 共measured by the corrugation height兲 and
culated spectral content of the wake velocity fluctuations. smoothes its surface gradients. Consequently, a drag increase
Moreover, note that the angle of attack of the transition from is expected since thin airfoils are more efficient at low Rey-
one vortex shedding mode, at the low angles of attack, to the nolds numbers.14 Therefore, a second family of airfoils was
second mode, at the higher angle of attack, indicated by the designed and analyzed. This type of airfoils is based on the
sudden decrease in St number with increasing ␣, was effec- contours created by the collection of points with a mean
tively reproduced by the computations in all the analyzed streamwise velocity equal to zero, encompassing the sepa-
cases 关Figs. 24共a兲, 25共a兲, 26共a兲, and 27共a兲兴. rated flow regions as well 关Figs. 28共a兲 and 29共b兲兴 and called
“v0-lined” airfoil. It should be noted that the results of this
design concept are incidence and Re dependent.
VIII. ANALYSIS OF CORRUGATED AND SMOOTH
Lift-drag performance for the four airfoil types were cal-
AIRFOILS
culated at Re= 6000 and the polar curves are presented in
In Sec. VII, the corrugated airfoil was compared to the Fig. 30. The airfoils with better performances were chosen
classical Eppler-E61 airfoil and shown to be more efficient. for presentation here. In this case too, the polar comparisons
The goal of this section is to try to understand the flow clearly show a superior lift-to-drag ratio for the corrugated
mechanisms which are responsible to this higher efficiency. airfoil 关Fig. 30共b兲兴. Note that the performance of the v0-lined
For this purpose, a comparison of the corrugated airfoil airfoil is quite similar to performances calculated for the
properties with those of an airfoil designed for Reynolds Eppler-E61. As for the Eppler-E61 flow behavior, long trav-
numbers of the order of 100 000 is not sufficient. It was eling waves 共beginning at about 0.3c with a length of order
hypothesized that an efficient, smooth airfoil will result from of 0.7c兲 are created over the airfoils without corrugations
a closed flow streamline, created by the corrugated airfoil at 共Fig. 31兲. This finding indicates again that the vortices pro-
the same flow conditions. Therefore, two families of smooth duced by the corrugations are responsible for the enhanced
airfoils based on the flow field around the corrugated airfoil efficiency, allowing intermittent flow reattachment to the
were designed and analyzed: The first is based on the closest curved aft region while diminishing in magnitude when in-
streamline that circles the corrugated airfoil 关for a specific Re tegrated over the entire chord 共as seen in Fig. 8兲. More spe-

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-14 D.-E. Levy and A. Seifert Phys. Fluids 21, 071901 共2009兲

FIG. 26. A comparison of numerical and experimental wake velocities spec- FIG. 27. A comparison of numerical and experimental wake velocities spec-
tral content for corrugated airfoil at X / c = 1.8: Re= 6000. 共a兲 Frequencies tral content for corrugated airfoil at X / c = 1.8: Re= 8000. 共a兲 Frequencies
comparison. Spectra comparison at 共b兲 ␣ = 6° and 共c兲 ␣ = 2°. comparisons. Spectra comparison at 共b兲 ␣ = 6° and 共c兲 ␣ = 2°.

cifically, the combination of near-leading-edge vortex forma- sis shows that a minimum limit of absolute value of 兩␻兩
tion with a relatively short traveling wavelength = 500 rad/ s 共nearly 2% of the maximum vorticity at 5% c
共significantly shorter than the chord length兲 is one of the from the trailing edge兲 is suitable to obtain converged circu-
major differences between the corrugated airfoil and its lation values 共⌫兲.
smooth counterparts. Resulting from these differences, the The data presented in Fig. 33 show the maximum “posi-
smooth airfoils are associated with an increase in the integral tive” vortices’ circulation as a function of the downstream
forces fluctuations 共Fig. 32兲. distance at two mean-lift coefficients for a representative
The reason for the smaller drag of the corrugated airfoil Reynolds number of 6000. It is noted that the circulation
is of fundamental and applicable importance, and as such, it
is important to understand its origin. An analysis of vortices’
circulation is capable of showing how much energy is lost in
the process of vortex formation and shedding over the differ-
ent airfoils investigated currently. Moreover, the vortices
shed into the wake, induce unsteady variations in the airfoil
incidence, affecting its unsteady loads as well.
An analysis of the vortices’ circulation evolution in the
wake of the different airfoils at a given Clmean is presented in
Fig. 33. The peak circulation magnitude 共⌫ = /␻dS兲 at each
S
point in the wake was calculated from the vorticity 关␻
ជ 兲兩, where U
= 兩rot共U ជ is the flow velocity vector兴 field com-
puted by Fluent, over the area of all the vortices in the do-
main which extends from 0.5c to 1.5c downstream from the
trailing edge.
FIG. 28. The flow lines chosen for the calculated “smoothed” airfoils fami-
In the calculation of ⌫, S represents the vortex area en- lies superimposed on the background of corrugated airfoil velocity vector
closed by an arbitrary limit value of ␻. A convergence analy- field at Re= 6000 and ␣ = 6°, 共a兲 v0-lined airfoil and 共b兲 streamlined airfoil.

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-15 Simplified dragonfly airfoil aerodynamics Phys. Fluids 21, 071901 共2009兲

FIG. 29. 共a兲 “Streamlined” family of airfoils, for Re= 6000, ␣ = 3° 共gray
line兲, 6° 共black line兲; 共b兲 v0-lined family, for Re= 6000, ␣ = 3° 共gray line兲,
and 6° 共black line兲.

magnitude of the vortices in the wake of the corrugated air-


foil is lower, which also has a lower drag coefficient 关Fig.
30共a兲兴. To investigate the universality of the above finding
we shall examine the circulation values of the positive vor-
tices in the wake 共2c from the leading edge; i.e., at 1c down-
stream of the trailing edge兲 of each airfoil as a function of the
lift coefficient, as presented in Fig. 34. The circulation of the
vortices in the corrugated airfoil wake is lower than the cor-
responding magnitudes of the Eppler-E61 wake over the en-
tire range of investigated lift coefficients. Furthermore, the
general behavior of these curves is quite similar to the be-
havior of the lift-drag polars 共Fig. 10兲, where lower circula-
tion levels correspond to lower drag values and also to lower
lift and drag fluctuations. These findings result from lower
magnitude vortices shed into the wake of the corrugated air- FIG. 31. Calculated C p distribution variations at half period 共T / 2兲 incre-
foil 共Fig. 7兲. ments, Re= 6000, Clmean = 0.82 for 共a兲 the streamlined airfoil 共b兲 the v0-lined
Therefore, it can be concluded that the vorticity fluctua- airfoil.
tions produced on the upper part of the corrugated airfoil
enable energy transport from the free-stream into the bound-
ary layer, in a manner similar to that produced by transition

FIG. 30. 共a兲 Comparison of calculated mean-lift-drag polar and 共b兲 mean-lift
to mean-drag ratio vs mean-lift coefficient of the different airfoils at Re FIG. 32. Comparison of calculated lift and drag coefficients’ oscillations as
= 6000. a function of time at Re= 6000, 共a兲 Clmean = 0.45, and 共b兲 Clmean = 0.77.

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-16 D.-E. Levy and A. Seifert Phys. Fluids 21, 071901 共2009兲

FIG. 34. Vortex circulation at 2ⴱc from the leading edge in the wake of the
Eppler-E61 and corrugated airfoils in function of the mean-lift coefficient

due to the local flow separations and captured vortices


produced by the corrugated regions and the intermittent
reattachment of the separated flow to the curved region
at the aft-upper surface, promoted by the vortices shed
from the corrugations.
共3兲 The vortices formed on the corrugated airfoil and shed
into its wake produce lower integral force fluctuations
due to their upstream pinch-off location and shorter
wavelength, leading to lower mean drag and lower inte-
gral forces fluctuations, as compared to smooth airfoils.
It was also found that the vortices’ magnitude in the
wake of the corrugated airfoil is lower, leading to
weaker induced velocities on the airfoil and therefore
lower fluctuating lift.
共4兲 Consequently, the calculated maximum range coefficient
FIG. 33. Vortex circulation development: maximum vorticity of the airfoil 共L / D兲max and the maximum endurance coefficient
wake vortices, for positive vorticity only, as a function of the distance from 共L1.5 / D兲max of specific flying configuration with corru-
the leading edge 共XLE / c. = 0兲, Re= 6000, 共a兲 Clmean = 0.41, 共b兲 Clmean = 0.78
for the corrugated, Eppler-E61 airfoils. gated airfoil should be superior to a similar configura-
tion with the smooth airfoils examined. In parallel, the
ability of the corrugated airfoil to reduce drag allows
to turbulence at higher Reynolds numbers, significantly re- reaching greater velocities.
ducing the wake width and consequently the pressure drag of
the airfoil. Moreover, the vortices shed over the airfoil and The numerical analysis of the corrugated airfoil was sup-
into the wake of the corrugated airfoil are weaker, resulting ported by experimental measurements, mainly for validation
in weaker fluctuations of the integral parameters. purposes. However, a more detailed experimental validation
should be performed. Some important parameters’ validation,
as the velocities around the airfoil should be performed, pref-
IX. SUMMARY AND CONCLUSIONS
erably using particle image velocimetry interrogation of the
A numerical and experimental study of a corrugated air- flow field. A parameter study of the corrugated airfoil shape
foil aerodynamics, based on a simplified dragonfly wing will be performed to draw engineering guidelines for wing
cross section, was performed at Reynolds numbers between design at this Reynolds numbers range. An implementation
2000 and 8000. The results of the corrugated airfoil analysis of active flow control for further improvement and hingeless
were compared to the performance of a “traditional” low maneuverability of MAVs is planned as well.
Reynolds number airfoil, the Eppler-E61, and also to a fam-
ily of smooth airfoils designed on the basis of the flow field ACKNOWLEDGMENTS
around the corrugated airfoil.
The authors would like to acknowledge valuable sugges-
The findings are as follows.
tions made by G. Iosilevskii 共Technion兲 and G. Zilman
共1兲 The assumption of laminar and 2D flow around the cor- 共TAU兲. Assistance by all TAU Meadow Aerolab members is
rugated airfoil was confirmed experimentally by both appreciated, especially by O. Stalnov. The technical assis-
qualitative flow visualizations and quantitative wake ve- tance of T. Bachar, S. Pasteur, and M. Goldberg is greatly
locity data in the range of 2000ⱕ Reⱕ 8000. This con- appreciated. The water tunnel experiments were enabled by
firms the relevance of the 2D laminar numerical simula- the kind support of M. Ringel and J. Sasson from the wind
tions. tunnel center of the Technion, Israel Institute of Technology,
共2兲 The lift-incidence slope is in some cases larger than 2␲ Haifa.

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
071901-17 Simplified dragonfly airfoil aerodynamics Phys. Fluids 21, 071901 共2009兲

1
T. J. Mueller and J. D. DeLaurier, “An overview of micro air vehicles smooth aerofoil compared,” Nature 共London兲 258, 141 共1975兲.
aerodynamics,” Prog. Astronaut. Aeronaut. 195, 1 共2001兲. 13
R. H. Buckholtz, “The functional role of wing corrugations in living sys-
2
H. Schlichting and K. Gersten, Boundary Layer Theory, 8th ed. 共Springer- tems,” ASME J. Fluids Eng. 108, 93 共1986兲.
Verlag, Berlin, 1999兲, p. 31. 14
M. Okamoto, K. Yasuda, and A. Azuma, “Aerodynamic characteristics of
3
J. D. Anderson, Fundamentals of Aerodynamics, 3rd ed. 共McGraw-Hill, the wings and body of a dragonfly,” J. Exp. Biol. 199, 281 共1996兲.
Blacklick, Ohio, 2000兲, pp. 307–308. 15
4
A. B. Kesel, “Aerodynamic characteristics of dragonfly wing sections with
G. B. McCulllough and D. E. Gault, “Examples of three representative technical aerofoils,” J. Exp. Biol. 203, 3125 共2000兲.
types of airfoil section stall at low-speed,” NACA Report No. TN 2502, 16
S. Sunada, A. Sakaguchi, and K. Kawachi, “Airfoil section characteristics
1951. at a low Reynolds number,” ASME J. Fluids Eng. 119, 129 共1997兲.
5
I. Tani, “Low-speed flows involving bubble separations,” Prog. Aeronaut. 17
S. Sunada, T. Yasuda, K. Yasuda, and K. Kawachi, “Comparison of wing
Sci. 5, 70 共1964兲. characteristics at an ultralow Reynolds number,” J. Aircr. 39, 331 共2002兲.
6
R. F. Huang and C. L. Lin, “Vortex shedding and shear-layer instability of 18
H. Gao, H. Hu, and Z. J. Wang, “Computational study of unsteady flows
wing at low Reynolds numbers,” AIAA J. 33, 1398 共1995兲.
7 around dragonfly and smooth airfoils at low Reynolds numbers,” 46th
Z. J. Wang, “The role of drag in insect hovering,” J. Exp. Biol. 207, 4147
AIAA Aerospace Sciences Meeting and Exhibit 共AIAA, Reno, NV, 2008兲,
共2004兲.
8
B. G. Newman, S. B. Savage, and D. Schouella, “Model test on a wing Paper No. 2008-385.
19
section of a dragonfly,” in Scale Effects in Animal Locomotion, edited by A. Vargas, R. Mittal, and H. Dong, “A computational study of the aero-
T. J. Pedley 共Academic, London, 1977兲, pp. 445–477. dynamic performance of a dragonfly wing section in gliding flight,” Bio-
9
J. M. Wakeling and C. P. Ellington, “Dragonfly flight,” J. Exp. Biol. 200, inspir. Biomim. 3, 026004 共2008兲.
20
543 共1997兲. R. Eppler, Airfoil Design and Data 共Springer-Verlag, Berlin, 1990兲.
21
10
R. J. Wooton, K. E. Evans, R. Herbert, and W. Smith, “Morphology and E. V. Laitone, “Aerodynamic lift at Reynolds numbers below 7 ⫻ 104,”
operation of the locust hind wing,” J. Exp. Biol. 203, 2921 共2000兲. AIAA J. 34, 1941 共1996兲.
11 22
A. B. Kesel, U. Philippi, and W. Nachtigall, “Biomechanical aspects of B. W. McCormick, Aerodynamics, Aeronautics, and Flight Mechanics
insects’ wings: An analysis using the finite element method,” Comput. 共Wiley, New York, 1979兲, p. 137.
23
Biol. Med. 28, 423 共1998兲. S. Brokman and D. Levin, “A flow visualization study of the flow in a 2 D
12
C. J. C. Rees, “Aerodynamic properties of an insect wing section and a array of fins,” Exp. Fluids 14, 241 共1993兲.

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
View publication stats

You might also like