Anda di halaman 1dari 12

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

net/publication/268458519

A Reconfigurable Wing for Biomimetic Aircraft

Conference Paper  in  Collection of Technical Papers - AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference · April 2013
DOI: 10.2514/6.2013-1511

CITATIONS READS
6 252

3 authors, including:

Christopher R Marks
Air Force Research Laboratory
37 PUBLICATIONS   126 CITATIONS   

SEE PROFILE

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

Active Flow Control for Low Pressure Turbine View project

All content following this page was uploaded by Christopher R Marks on 31 July 2015.

The user has requested enhancement of the downloaded file.


A reconfigurable wing for biomimetic aircraft

Christopher R. Marks1
University of Dayton Research Institute, Dayton, Ohio 45469-0110

James Joo2, Greg Reich3


U.S. Air Force Research Laboratory, Wright-Patterson Air Force Base, Ohio 45433-7542

Abstract
A biomimetic composite reconfigurable wing for small unmanned air systems has been designed and tested
experimentally in a wind tunnel. Actuation of the wing results in a 30% reduction of wing planform area and
34% decrease in span mimicking wing sweeping motion used by birds for flight control. The wing uses a thin
cambered airfoil along the span. Wing sweep motion of 56° was demonstrated under static aerodynamic
loads at angles of attack between -5 and +15° at 10 m/s. The wing was fabricated using a variety of materials
including carbon fiber and fiberglass composite morphing skins and direct manufactured plastic frame
structures. The large shear displacement required by the skin to achieve large wing sweep angles was
managed by folding large sections of the skin, similar to folding of bird feathers.

Nomenclature
AR = aspect ratio
b = wing span
c = mean aerodynamic chord length
Re = Reynolds number
S = wing planform area
γ = outer wing sweep angle

I. Introduction

B irds possess a level of flight control beyond the current limits of man-made flight enabling flight in highly
spatially constrained environments such as forests and urban settings. Most species of birds have the ability to
fly under powered flight or by gliding, and use both flapping and soaring during any given flight. We focus here on
gliding flight configuration, for either soaring, or approach to perch. Like a modern gliding aircraft the primary
lifting surface is the wings and the primary control surfaces are located on the wings and tail. Additional control is
obtained by shape changing; by variation of wing dihedral, aspect ratio, wing sweep, planform shape, area, and by
moving appendages (such as legs and feet) as control surfaces or to vary the center of gravity. Videler [1] suggests
that in many species all appendages are related to flight. In contrast to modern aircraft the wings and tail of a bird
are significantly more complicated, constructed of multi-degree of freedom mechanisms and covered by feathers,
enabling shape morphing, and passive and active flow control.
Researchers have attempted to understand and mimic some of the advanced flight control abilities of the avian
world. Two active research areas that trend toward biomimicry are avian tails and wings. Bird tails lack the vertical
stabilizer found on airplanes. Birds can twist, warp, change area, and deflect their tails in flight. Sachs [2] showed

1
Research Engineer, Aerospace Mechanics Division, 300 College Park, Senior Member AIAA.
2
Mechanical Research Engineer, Aerospace Systems Directorate, 2130 Eighth St., Senior Member AIAA.
3
Aerospace Engineer, Aerospace Systems Directorate, 2130 Eighth St., Senior Member AIAA.

1
American Institute of Aeronautics and Astronautics
that manmade air vehicles of the same size as birds could be made without a vertical tail if the wing took over the
same functions as the airplane vertical tail (i.e provide necessary static and dynamic yaw stability). Chakravarthy et
al. [3] investigated variable wing dihedral and variable wing twist for control of a vertical tailless vehicle during a
perching maneuver. The use of variable dihedral angles to control aircraft yaw and aircraft speed independent of
the angle of attack and flight path angle was demonstrated by Paranjape et al. [4]. Parga et al. [5] experimentally
studied a biomimetric rotating tail, noting that a significant challenge is the coupled effect of the two actuators on
yaw, roll, and pitch.
A notional vehicle designed for perching using rotating wings is described in Lukens et al. [6] and by Siegler et
al. [7]. Rotation of the wing provided an additional degree of freedom for control of lift, drag, and vehicle
orientation. Grant et al. [8] study a morphing wing design with independent inboard and outboard wing section
sweep. They found enhanced turning capability and crosswind rejection.
The work described here was part of a broader research project investigating the perching of bird sized fixed
wing vehicles with a focus on vehicle design [9], aerodynamic modeling[10]-[11], and perching methods
development [12]. This paper describes a new concept for wing shape morphing as part of an effort to develop a
reconfigurable wing that mimics the wing planform area change accomplished by small raptors. The ability to
sweep the wings is useful for both flight control, and for biomimicry in the air, during perching, and during roost.
The planform area change created by wing sweeping is useful for flight control (generating roll), and during extreme
maneuvers such as perching. A four bar mechanism integrated into the structure of the wing imitates the sweeping
motion of a bird and allows for both forward and aft wing sweep with one actuator. The wing design achieves large
wing sweep angles and planform area changes while using a cambered airfoil profile rather than a simple flat plate
section profile.

II. Wing Design


Birds have the ability to sweep the hand-wing forward or
backward. When a bird sweeps its hand-wing (outer portion
of wing consisting of the primary feathers) rearward the arm-
wing (inboard portion containing the secondary feathers)
sweeps in the opposite direction (See Fig. 1). In many raptor
size birds, the arm-wing moves forward with the wrist joint,
while the hand-wing is swept to the rear. The wing design
described in this work sought to mimic the complex sweeping
motion of avian wing using a single degree of freedom
mechanism integrated in a cambered airfoil.
A four bar mechanism was chosen for the wing design.
This mechanism enables forward or rearward hand-wing Figure 1. The American Kestrel in flight is an
sweep while mimicking overall planform shape change created example of a raptor wing planform shape.
when the arm-wing sweeps forward and the hand-wing is (image courtesy of Earl Reinink ©)
swept rearward (see Figure 2b). The four bar mechanism itself
is primarily located inside the inboard portion of the wing and is shown overlaid on the final wing in Figure 2a. The
wing geometry of a variety of bird species can be mimicked using this four bar mechanism by varying the length of
the outer wing member.
The choice of wing cross-sectional shape is integral to wing and vehicle design. In the case of a reconfigurable
wing, the choice of cross sectional shape must be considered mutually with the mechanism and structural design.
Large planform area changes require management of significant surface shear displacement by either stretching or
shrinking of skins, or by hiding areas inside the wing or fuselage (Weisshaar [12]). A single surface membrane or
flat plate cross section could be considered as the baseline, leading to easier management of arm and hand-wing
planform surface variation, but results in limited aerodynamic performance. The use of an airfoil cross section
results in higher lift, but curved surfaces complicates wing sweep. A thicker cross section is required to embed a
mechanism and corresponding structure inside the wing. In our approach we use a thin cambered airfoil with a
single surface at the trailing edge as a pattern for the wing. The wing cross section is not uniform across the span.
The arm-wing cross section differs from the hand-wing cross section on the pressure side of the airfoil, which is
open on the arm-wing, and closed for the hand-wing. The area of the wrist joint includes a fairing on the pressure
side that covers the joint structure and bearing surface,

2
American Institute of Aeronautics and Astronautics
Forward Sweep

+γ Backward
Sweep

a.)

b.) c.)
Figure 2 a.) Four bar mechanism geometry integrated into wing structure overlaid on top of final wing design
showing forward and backward wing sweep; b.) Reconfigurable wing design overlaid with a Pheasant wing
showing mimicry of actual avian wing sweep and planform area change; c.) Wing mounted on a notional
“birdlike” body.

Figure 3 Pressure side of reconfigurable wing highlighting the folding panel section.

3
American Institute of Aeronautics and Astronautics
and blends into the hand-wing cross sectional profile. Both passive and active wing surfaces are used to achieve up
to a 30% reduction of planform area and 34% reduction in span with 56° of sweep. The active wing surfaces consist
of curved surface panels that fold like a fan, and then into a passive area located at the wing root. Figure 3 shows
the active and passive wing sections. Areas marked with an a and e are passive portions of the wing with no moving
parts. The folding fan panels are marked with a b. Each panel is made of carbon fiber and fiberglass composite
formed on a custom mold. The out of plane stiffness near the pivot point is increased by “hand cast” carbon fiber
stiffeners mounted to each fan (labeled c Figure 3 and shown in detail of Figure 4). Each fan stiffener has an
integrated bearing surface used for the fan pinned joint. The fan panels are connected to each other and to the
inboard passive piece near the trailing edge by a pin and cap passing through slots in each panel. Each panel
required enough flexibility near the trailing edge to allow some deformation as the curved surfaces folded over each
other when the wing was swept rearward.
The skin is made of carbon fiber and fiberglass composite. The inner wing leading edge area is carbon fiber and
fiberglass weave skin over a direct manufactured plastic internal frame structure that extends from the wrist joint to
the root joint. The internal frame forms an internal ‘D’ box structure to carry the wing out of plane load. The wrist
joint is a pinned joint integrated in the composite skin (see Figure 5). The design of the internal frame and wrist
joint enable the wing to sweep through its full range of motion while maintaining a streamlined leading edge surface
at the joint.

Carbon fiber
stiffener

Composite fan
panel

Figure 4: Detailed view of fan blade design.

•Integrated into complex airfoil shape


•Maintain streamline leading edge during sweep

Figure 5: The wrist joint consists of a pinned joint integrated into wing.

4
American Institute of Aeronautics and Astronautics
The internal frame structure of the both the arm and hand-wing includes the wrist joint bearing surface. Metal
inserts bonded to the internal frames with form the outer surface of a plain bearing at the wrist joint. A tubular metal
axle forms the inner surface of the plain bearing. The passive portion of the hand-wing section is comprised of a
CNC machined foam core in the outboard portion of the wing, and a direct manufactured internal frame structure at
the wrist joint. The skin near the joint is primarily carbon fiber weave, fiberglass near the trailing edge, and carbon
fiber unidirectional tow reinforcement along the skin oriented like spar and ribs. In most areas only one or two
layers of lightweight (2 – 5.7 oz.) fabric was used for the skin to assure a lightweight structure. The folding fan
pinned joint was attached to the passive portion of the hand-wing by a small screw passing directly through the
composite skin. The final wing weight was 76 grams which includes internal plastic structure, metal screws, and
carbon fiber/Nomex honeycomb pass through structure. The outer wing passive area and folding fan panels are the
two areas in which additional weight savings could be obtained.
Full span wing parameters are listed in Table 1. The maximum wing loading tested is calculated from wind
tunnel testing described in the next section. The change in span, planform area, and mean chord versus sweep angle
are listed in Figure 6.
Table 1. Full Span Wing Parameters at Zero Sweep Angle
Wing Span Mean Aerodynamic Planform Area Aspect Ratio Weight Max Wing Loading
(cm) Chord (cm) (cm2) (gram) Tested (kg/m2)
59.7 12.87 743.3 4.8 76 4.4

30
Span
20 Planform Area
Cx

10
% Change

-10

-20

-30

-40
0 10 20 30 40 50 60
 (deg)

Figure 6 Change in geometry versus sweep angle (γ).

III. Wing Testing


A half span, full scale version of the wing was tested the University of Dayton Low Speed Wind Tunnel (UD-
LSWT) to verify wing sweep under aerodynamic load at notional flight speed, and to obtain static force
measurements. The UD-LSWT is powered by a 60 HP motor, has a 30 inch x 30 inch test section, and 120 ft/s
maximum tunnel velocity. There are six anti-turbulence screens on the tunnel inlet providing a turbulence intensity
of less than 0.1% (Lego et al. [14]). A small symmetric body was installed around the wing root area to conceal a
servo used to sweep the wing. A hobby grade digital servo was used to sweep the wing (see Table 2). The servo
was driven by a Parallax servo controller. Wing sweep angle achieved in the wind tunnel was calculated
geometrically from video.
Forces and moments at fixed sweep angles were measured using an ATI Gamma F/T sensor. The maximum
force and torque that can be applied to the sensor along with maximum precision is listed in Table 3. The signal was
low pass filtered, and then averaged over five seconds.

5
American Institute of Aeronautics and Astronautics
Table 2 Servo specifications. Table 3 ATI Gamma F/T transducer maximum
force, torque, and accuracy at maximum load.
Max Force Max Torque Accuracy
Direction
(N) (N-m) (N)
x 33.4 2.8 0.0018
y 33.4 2.8 0.0018
z 33.4 2.8 0.0018

Wing sweep from γ = 0° to 56° and then back to 0° under aerodynamic loading was verified multiple times at 10
m/s inlet velocity (Rec = 8.5x104) at angles of attack between -5 and +15°. Full wing sweep to γ = 56° was not
possible at angles of attack above +10° using 4.8V and corresponding lower servo torque; full wing sweep was
achieved with 6.0V at α = +10° to +15°.

Figure 7. Still photos extracted from video of wing sweeping under aerodynamic load.
Static force measurements were made at five different sweep angles. The wing angle was locked at a given
angle by replacing the servo with a bracket that set the wing at the correct angle. A photo and geometric parameters
at each position are shown in Table 4.

Table 4 Geometric parameters and photos of the five different wing sweep angles settings at which static force
and moment data was measured.

a.) γ = 0o b.) γ = 14o c.) γ = 34o d.) γ =47o e.) γ = 57o


S = 743.3 cm2 S = 698.4 cm2 S = 629.1 cm2 S = 577.8 cm2 S = 524.8 cm2
c = 12.87 cm c = 12.55 cm c = 13.24 cm c = 13.97 cm c = 15.90 cm
b = 59.7 cm b = 59.6 cm b = 53.8 cm b = 47.4 cm b = 40.4 cm
AR = 4.8 AR = 5.1 AR = 4.6 AR = 3.9 AR = 3.1

6
American Institute of Aeronautics and Astronautics
The wing static lift and drag coefficients measured in the wind tunnel are shown in Figure 7. The pitch moment
about the quarter chord is shown in Figure 9. The Reynolds number based on mean chord was 8.5e4, with local
Reynolds number varying from 1.0e5 at the root to 5e4 at the chord. Examining the lift coefficient data indicates
that CL,0 occurs at negative angle of attack due to the cambered airfoil sections. An increased lift slope measured for
the wing that is swept to γ = 14o compared to zero wing sweep. This may be due to slightly higher aspect ratio,
change in lift along the span, and change in streamwise section profiles. As wing sweep angle is increased the slope
gradually decreases as expected primarily due to reduction in aspect ratio. Examining the drag coefficient indicates
that the drag increases when the wing is swept at low angle of attack. The highest drag coefficient at angles of
attack near zero occurs when the wing is fully swept. Figure 8c shows an increase in drag at values of C L near zero
(a flattening and even inverted area of the curve), which could be attributed to low Reynolds number boundary layer
separation phenomena. Understanding the increase in drag coefficient with sweep angle and angle of attack is
difficult since there are many surface and profile changes occurring when the wing is swept. Detailed flow
visualization or boundary layer measurements would certainly help understand the flow behavior, but were not
possible due to project constraints.

Re = 1.0 x 105 Re = 1.0 x 105


0.7 0.35

0.6
0.3
0.5

0.4 0.25

0.3
0.2
CD
CL

0.2
0.15
0.1 
 = 0 =0
0  = 14

0.1  = 14
-0.1  = 34

 = 34
 
 = 47 0.05  = 47
-0.2  
 = 57  = 57
-0.3 0
-10 -5 0 5 10 15 20 -10 -5 0 5 10 15 20
 (Degrees)  (Degrees)

a.) b.)
Re = 1.0 x 10 5 Re = 1.0 x 105
0.7 6

0.6 5

0.5 4

0.4 3

0.3 2
L/D
CL

0.2 1

0.1  0 
=0 =0
0  = 14  -1  = 14

-0.1  = 34 
-2  = 34
 
 = 47  = 47
-0.2 -3 
  = 57
 = 57
-0.3 -4
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 -10 -5 0 5 10 15 20
CD  (Degrees)

d.)
c.)

Figure 8. Lift and drag coefficient measurements at fixed sweep positions.

7
American Institute of Aeronautics and Astronautics
Re = 1.0 x 105
0

-0.01

-0.02

-0.03

-0.04

M -0.05
C

-0.06

-0.07  = 0
 = 14
 = 34
-0.08

-0.09  = 47
 = 57
-0.1
-10 -5 0 5 10 15 20
 (Degrees)

Figure 9. Pitch moment coefficient about ¼ chord.

The most significant outcome of the force measurements is the difference between the predicted wing force
coefficients versus experiment. The well known 2D design tool XFOIL [15] was used to predict 2D airfoil
performance with corrections for a finite wing based on thin airfoil theory [16]. Predictions from XFOIL compared
with experiment are shown in Figure 10. Wind tunnel measurements of lift coefficient with the wing extended were
significantly lower than our predictions. Some difference in prediction versus experiment was expected owing to
the considerable variation between the wing cross-sectional profiles at the root compared to the baseline profile at
the wing tip. In addition to cross-sectional variation, the wing is made of many different panels, some overlap
irregularly, and some panels have slots that certainly have an effect on aerodynamic forces.

2
Experiment
AR = 4.8
1.75 Theory (AR = 4.8)
XFOIL 2D Re = 1,000,000
1.5
XFOIL 2D Re = 85,000
1.25 XFOIL 2D Re = 50,000
Lift Coefficient

1
0.75
γ = 0°
0.5
0.25
0
-0.25
-0.5
-0.75
-1
-15 -10 -5 0 5 10 15 20 25
Angle of Attack (deg)

Figure 10. Lift measurements compared to prediction with wing extended (zero sweep).

8
American Institute of Aeronautics and Astronautics
Figure 10 contains 2D predictions from XFOIL of the baseline airfoil at three different Reynolds numbers
including the Reynolds number based on mean chord, the local Reynolds number at the wing tip, and an arbitrarily
high Reynolds number at which viscous forces are much smaller than in the experiments. As Reynolds number
decreases, both stall angle of attack, and Cl,max decrease. At low Reynolds number, and low angles of attack viscous
forces have a larger effect on airfoil performance which results in lower predicted lift coefficients at Re = 50,000
and 85,000 compared to Re = 1 million. XFOIL predicts a long laminar separation bubble on the suction surface of
our airfoil at small, positive angles of attack. As angle of attack is increased the separation bubble length decreases
up to an angle of attack of six degrees at which point it has a smaller effect on lift coefficient. At low Reynolds
number and low angles off attack the pressure side of the airfoil is separated with the separation length decreasing as
angle of attack is increased.
The reason for the large deviation of measured lift slope versus prediction is unknown. It is clear that the source
of the discrepancy remains for all sweep angles tested. Three different theories which could explain the large
difference in actual lift coefficient versus predicted have been formulated:
1. Wing twist or TE deflection under aerodynamic load
2. Diminished aerodynamic performance due to slots, holes, and steps in geometry
3. Boundary layer separation at low AOA due to low Reynolds number or three dimensional effects
not captured in prediction method.
The first theory was investigated using XFOIL by varying the shape of the section geometry to mimic the shape
of a deformed wing. Both wing twist and trailing edge deflection were investigated. Wing twist would create a
lower effective section angle of attack. Analysis showed that the wing would have to twist as much as nine degrees
to account for the reduced lift coefficient measured in experiment. A large degree of wing twist would have been
noticeable during experiment. Images from high speed camera video acquired during wind tunnel testing are shown
in Figure 11. No noticeable wing twist was observed during experiment. The images at α = 15° indicates that the
trailing edge deflects a small amount in the area of the folding fan. Analysis indicated that the deflection observed
would not have caused the low lift coefficient measured in experiment. The second and third theories have not yet
been exhaustively investigated, but are the likely sources of the discrepancy between predicted aerodynamic forces
and those measured experimentally. Experimental force measurements by Ananda et al. [17] of a moderate aspect
ratio (AR=4) wing at low Reynolds number found a critical Reynolds number below which a significant decrease in
lift and drag ratio was observed due to laminar boundary layer separation without reattachment. Below the critical
Reynolds number the lift slope deviated from theoretical, even at small angles of attack.

No significant twist
Small deflection of fan at wing tip
section near TE

α = 10 α = 15
a.) b.)
Figure 11. Photographs extracted from high speed video show no significant wing twist under aerodynamic
loading.
While our wing design produces enough lift to satisfy our notional vehicle design, the performance would be less
than predicted. The difference in experimentally measured lift forces on our wing compared to prediction justifies
reiterating an earlier point: In the case of a reconfigurable wing design, the choice of cross sectional shape must be
considered mutually with the mechanism and structural design in order to achieve desired performance. In our

9
American Institute of Aeronautics and Astronautics
case, relatively low fidelity aerodynamic prediction methods were not sufficient to capture the aerodynamic
performance of our complex reconfigurable wing geometry.

IV. Conclusion
A reconfigurable wing that mimics the wing sweeping motion of raptors was designed and tested in a wind
tunnel. This will enable the vehicle to fold its wings away upon landing, enabling it to hide-in-plain-sight by
appearing similar to a bird. It will also make the vehicle much less susceptible to wind gusts while perched. Wing
folding could also be used in flight to make changes to the wing area such as for roll control and might also be
useful for perching maneuvers. Forward and rearward wing sweep of birds was achieved by embedding a four bar
mechanism inside the wing structure. The folding wing uses a cambered airfoil along the span and was fabricated
using a variety of materials including carbon fiber and fiberglass composite skins and 3D printed plastic frame
structures. The large shear displacement required by the skin to achieve large wing sweep angles was mitigated by
folding large sections of the skin, similar to folding of bird feathers. The wing shape change was tested in a wind
tunnel under static aerodynamic loads at angles of attack between -5 and +15° at 10 m/s and demonstrated the full
range of motion under aerodynamic load with lightweight servos of the size that would typically be used for aileron
or elevator control on a small remote controlled aircraft. The lift slope obtained with the wing was however
significantly below the theoretical value. Analysis of photographs taken during the experiments indicated that
reduced lift slope was not due to wing deformation under aerodynamic load. The exact cause reduced aerodynamic
performance compared to theoretical is still under investigation.

Acknowledgements
The authors would like to acknowledge the assistance of Lauren Zientarski who fabricated and assembled key
pieces of the reconfigurable wing during development and testing, Nick Jones for assistance with fabrication, and
Zach Lego for assistance with wind tunnel setup and testing.

References

[1] Videler, J. J., Avian flight. Oxford University Press, Oxford, United Kingdom, 2005.
[2] Sachs, G., “Why Birds and Miniscale Airplanes Need No Vertical Tail,” J. of Aircraft, Vol. 44, pp. 1159-1167, 2007.
[3] Chakravarthy, A., Paranjape, A, Chung, S. J., “Control Law Design for Perching an Agile MAV with Articulated Wings,”
AIAA Flight Mechanics Conference, Toronto, Canada, 2-5 August 2010, AIAA Paper 2010-7934.
[4] Paranjape, A. A., Chung, S. J., Selig, M., “Flight Mechanics of a Tailless Articulated Wing Aircraft,” Bioinspiration &
Biomimetics, Vol. 6, pp. 1-20, 2011.
[5] Parga, J. R., Reeder, M. F., Leveron, T., Blackburn, K., “Experimental Study of a Micro Air Vehicle with a Rotatable Tail,”
J. of Aircraft, Vol 44, pp. 1761-1768, 2007.
[6] Lukens, J. M., Reich, G. W., Sanders, B., “Wing Mechanization Design and Analysis for a Perching Micro Air Vehicle,” 49th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, Schaumburg, IL, 7-10 April,
2008.
[7] Siegler, T., M., Lubbers, J. L., Reich, G. W., “Perch Landing Maneuvers for a Rotating Wing MAV,” 51 st
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, Orlando, Florida, 12-15 April
2010, AIAA Paper 2010-2826.
[8] Grant, D. T., Abdulrahim, M., Lind, R., “Flight Dynamics of a Morphing Aircraft Utilizing Independent Multiple-Joint
Wing Sweep,” AIAA Atmospheric Flight Mechanics Conference, Keystone, Colorado, 21-24 August, 2006, AIAA Paper
2006-6505.
[9] Robertson, D.K. & Reich, G.W., "Design and Perching Experiments of Bird-like Remote Controlled Planes,” 54th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, Boston Massachusetts, 8-11
April, 2013.

10
American Institute of Aeronautics and Astronautics
[10] Robertson, D.K. & Reich, G.W., 3-D Vortex Particle Aerodynamic Modeling and Trajectory Optimization of Perching
Maneuvers,” 54th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, Boston
Massachusetts, 8-11 April, 2013.
[11] Meckstroth, C., Reich, G., "Aerodynamic Modeling of Small UAV for Perching Experiments," AIAA Applied
Aerodynamics Conference, San Diego, CA, 24-27 June, 2013 (submitted for publication).
[12] Meckstroth, C., Reich, G., "Near-Optimal Perching Trajectory Selection using Bézier Curve Interpolation," AIAA
Guidance, Navigation, and Control Conference, Boston, MA, 19-22 August, 2013 (submitted for publication).
[13] Weisshaar, T. A., Morphing Aircraft Technology – New Shapes for Aircraft Design. In Multifunctional Structures /
Integration of Sensors and Antennas (pp. O1-1 - O1-20). Meeting Proceedings RTO-MP-AVT-141, Overview 1. Neuilly-
sur-Seine, France: RTO, 2006.
[14] Lego, Z., Altman, A., Gunasekaran, S., “Experimental and Computational Analysis of High Angle of Attack Perching
Maneuvers,” 30th AIAA Applied Aerodynamics Conference, New Orleans, Louisiana, 25-28 June 2012, AIAA Paper 2012-
2666.
[15] Drela, M., and Youngren, H., “XFOIL 6.9 User Primer,” http://web.mit.edu/drela/Public/web/xfoil_doc.txt, 2001.
[16] Anderson, J. D. Jr., Fundamentals of Aerodynamics, 5th Edition, McGraw-Hill, New York, 2011.
[17] Ananda, G. K., Sukumar, P. P., Selig, M. S., “Low-to-Moderate Aspect Ratio Wings Tested at Low Reynolds number,” 30th
AIAA Applied Aerodynamics Conference, 25-28 June, 2012, New Orleans, Louisiana, AIAA 2012-3026.

11
American Institute of Aeronautics and Astronautics

View publication stats

Anda mungkin juga menyukai