Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

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

net/publication/268477705

Aeroelastic Flight Control for Subscale UAVs

Conference Paper · April 2007


DOI: 10.2514/6.2007-1706

CITATIONS READS

5 171

3 authors, including:

Vijayakumari Hodigere Siddaramaiah Je Cooper


Jain University University of Bristol
13 PUBLICATIONS   103 CITATIONS    270 PUBLICATIONS   2,804 CITATIONS   

SEE PROFILE SEE PROFILE

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

SLOWD - Sloshing Wing Dynamics View project

CLAReT View project

All content following this page was uploaded by Vijayakumari Hodigere Siddaramaiah on 04 April 2016.

The user has requested enhancement of the downloaded file.


48th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference<br>15th AIAA 2007-1706
23 - 26 April 2007, Honolulu, Hawaii

Aeroelastic Flight Control for Subscale UAVs

Roelof Vos ∗
The University of Kansas, KS
Vijaya Hodigere-Siddaramaiah † Jonathan E. Cooper ‡

University of Manchester, United Kingdom

This paper describes a new type of aeroelastic wing panel which relies on a sweeping
movement of the main spar to alter its flexural axis and consequently change the amount of
wash-out in the wing induced by the aerodynamic forces. A 1.8m (6ft) electrically powered
glider was fitted with two aeroelastically active panels at either side of the wing, each
Downloaded by BRISTOL UNIVERSITY on April 2, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2007-1706

spanning 450mm (17.7in.). By hinging the main spar at the root of each panel it could
sweep over an angle of 14◦ peak-to-peak, covering 60% of the local chord at the wing tip.
The wing structure was entirely composed of balsa wood members and the front and rear
spars were intentionally left flexible to increase the dependency of the local elastic center
on the main spar position. Furthermore, a composite skin was used, which provided enough
torsional flexibility to allow for active wing twisting. The application of the modified wing
panels added 3% to both the operational empty weight and planform area, leaving the
wing loading unchanged. To model the aeroelastic behavior, a finite element model of the
wing was made using MSC.PATRAN/MSC.NASTRAN and coupled to a three dimensional
aeroelastic model (ZAERO). Dynamic bench tests demonstrated a change of 40% in elastic
center position at the tip of the aeroelastic wing panel and showed good agreement in
bending natural frequencies with the finite element model. Furthermore, an increase of
28% in the first coupled natural frequency was demonstrated between the most forward
and most rearward position of the main spar. Wind tunnel tests were carried out to show
wing twist dependency on main spar position and angle of attack. At cruise conditions,
when sweeping the main spar over its maximum range, a change in lift coefficient of 35%
and a change in twist of 3.7◦ were shown. At increased angle of attack, the range of twist
even increased to 9.1◦ peak-to-peak. Static aeroelastic modeling showed to be accurate in
predicting the twist at spar sweep angles ranging from -4.5 to +2.5 degrees. Flight test
proved the effectiveness of aeroelastic roll control without any weight penalties.

Nomenclature
Symbol Description Units
A Aspect Ratio or Amplitude −, m
b Span m
c Chord m
Cd , C D Section and wing drag coefficient −
Cl Section lift coefficient or rolling moment coefficient −, −
CL Wing lift coefficient −
f Frequency Hz
Iy Roll moment of inertia kgm2
p Roll rate deg/s
Re Reynolds number −
S Wing reference surface area m2
t Thickness m
∗ PhD Student, Department of Aerospace Engineering, 2120 Learned Hall, Lawrence, KS 66044, AIAA member
† PhD Student, School of Engineering, Oxford Road, Manchester, M13 9PL, UK
‡ Professor of Engineering, Oxford Road, Manchester, M13 9PL, UK., Associate AIAA member

1 of 12

American Institute for Aeronautics and Astronautics

Copyright © 2007 by Roelof Vos. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
W Weight N
α Angle of attack deg
Λ Main spar sweep angle deg
θ Wing Rotation angle deg
ρ Density kg/m3

Subscripts
B Bending Test
d Divergence
ec Elastic center
f Flutter
L Longitudinal direction
le Leading Edge
r Root
R Radial direction
t Tip or torsion
te Trailing edge
T Tangential direction
Downloaded by BRISTOL UNIVERSITY on April 2, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2007-1706

Abbreviations
AAW Active Aeroelastic Wing
ARTF Almost Ready To Fly
FE Finite Element
HALE High Altitude Long Endurance
OEW Operating Empty Weight
UAV Uninhabited Aerial Vehicle

I. Introduction

Ithisnto recent years there is a growing interest in aircraft which are able to change the shape of their wing
1–13
optimize flight performance or to control the rolling motion of the aircraft. One way to achieve
so-called wing morphing is to build aircraft structures that allow aeroelastic deflections to be used in
a beneficial manner and to enhance aerodynamic performance.14–19 This concept has been investigated in
various research programs over the past decade. The Active Flexible Wing20 and Active Aeroelastic Wing
(AAW)21 programs investigated the use of leading and trailing edge control surfaces to control the wing
twist. With very little control surface motion, the AAW techniques employ the energy of the airstream to
achieve the desirable wing twist. The Morphing program22 developed a number of active aeroelastic concepts
based upon smart materials and structures. To improve aileron effectiveness, the leading Russian Aerospace
research institute, TsAGI, successfully demonstrated the aeroelastic application of a small additional control
surface ahead of the wing.23 In Europe, the 3AS (Active Aeroelastic Aircraft Structures) research program24
also developed and demonstrated various active aeroelastic concepts, primarily in the areas of adaptive
attachments, three surface aeroelastic aircraft and novel aeroelastic leading edge and wing tip devices. The
Variable Stiffness Spar approach25, 26 demonstrated the use of rotating spars for roll control.
Aeroelastic phenomena have always had a significant influence on aircraft structural design. It has been
accepted that in order to avoid the occurrence of flutter or divergence anywhere within the flight envelope,
the lifting surfaces must be made stiffer than the quasi-static aerodynamic loads would require. The most
common passive solution to dynamic aeroelastic problems are increased stiffness and mass balance, which
was used as early as 1922.27 This requirement has been termed the “aeroelastic penalty.”28
Rather than suppressing any wing deformation induced by the air loads, Active Aeroelastic Wings utilize
the aerodynamic forces to introduce a twist in the wing structure. The twist in the wing can be used to provide
longitudinal control or to have an optimal wing shape (e.g. maximum lift to drag ratio) at various flight
conditions. Currently, civil and military aircraft are designed to have optimal aerodynamic characteristics

2 of 12

American Institute for Aeronautics and Astronautics


(maximum lift/drag ratio) at one point and fuel condition in the entire flight envelope. However, the fuel
loading and distribution changes continuously throughout the flight, and aircraft often have to fly at non-
optimal flight conditions due to air traffic control restrictions. The consequent sub-optimal performance has
more significance for commercial aircraft as they are more flexible than military aircraft and also have fuel
efficiency as a performance parameter of far greater importance. There is also much recent interest in High
Altitude Long Endurance (HALE) aircraft which are designed to fly for several days at a time and have a
greater proportional of fuel to take-off weight than other aircraft, the resulting changes in the aeroelastic
shape throughout the flight can be substantial. Fuel efficiency is becoming increasingly important for civil
and HALE aircraft, and any approach that enables better aerodynamic performance throughout a flight
needs to be investigated and developed.
The advantages of Active Aeroelastic Wings over conventional aileron control are their higher total energy
efficiency29, 30 and their lack of aileron reversal behavior.19 Other expected benefits of AAW control over
conventional control are increased control power, reduced aerodynamic drag, reduced maneuver loads and
reduced aircraft structural and take-off weight due to relaxed stiffness and hinge moment requirements.31 By
shifting to AAW control it becomes possible to make high aspect ratio wings without having to add weight
to stiffen the wing and reduce aileron reversal. On the other hand, flutter speed and divergence speed can
decrease when the internal structure of the wing is made more flexible.
The center of gravity of AAW research has been on the application of multiple slats and flaps to induce
twist in the wing.19, 31, 32 Research showed that AAW control could increase roll control on an F/A-18A
Downloaded by BRISTOL UNIVERSITY on April 2, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2007-1706

while total weight was decreased.33, 34 An alternative way to allow for aeroelastic wing twisting is by actively
changing the stiffness properties of the wing during flight. This method does not require any flaps or slats
and can be easily integrated in high aspect ratio wings. Part of the 3AS research program investigated the
use of changes in the internal wing structure in order to control the static aeroelastic bending and twisting
behavior of a number of simple wings.35 This paper presents a continuation of this research tailored towards
the integration of hinged spars in a high aspect ratio wing. The design, manufacturing and testing of a
wind tunnel model based on the hinged-spar concept is described. Comparisons are made with the results
obtained from Finite Element simulations.

II. Design and Manufacturing of Prototype Demonstrator


In conventional wing structures, the position of the elastic center is ideally located at the quarter chord
point. Any change in lift (L) induced by a change in angle of attack attaches at this point (c/4), and
consequently does not induce any twist in the wing. When shifting the elastic center behind the quarter
chord point, any change in lift leads to a positive change in wash-out. This in turn leads to a higher local
angle of attack (∆α) and a higher lift. Up to the divergence speed, this process tends to converge to a
fixed combination of wash-out and lift. When shifting the elastic center in front of the quarter chord point
the opposite effect occurs and the wing experiences an additional negative wash-out. The dependence of
the local rotation of a wing section as a function of the elastic center is shown in Figure 1. The torsional
stiffness of the wing (G) is another parameter which has a major influence on the amount twist induced
by the aerodynamic forces. In particular the stiffness/flexibility of the wing skin determines the amount of
twist the wing will experience during aerodynamic loading.

Figure 1. Schematic Representation of Section Rotation as a Function of Elastic Center Position

To demonstrate the dependence of wing twist on the position of the flexural axis, a new wing design
was made for a 1.8m (6ft) span sub-scale Uninhabited Aerial Vehicle (UAV). An off-the-shelf ARTF glider
with only lateral and directional control was used as the baseline aircraft. The outboard 0.45m (18in) of its

3 of 12

American Institute for Aeronautics and Astronautics


wing was positioned at a polyhedral angle of 10◦ . For convenience, it was determined to modify the internal
structure of only this part of the wing. A design was made for two aeroelastic panels that would be positioned
instead of the original polyhedral panels. At the polyhedral point the main spar was hinged, such that it
would be able to sweep backwards and forwards in chordwise direction throughout the aeroelastic panel. The
sweep angle, Λ, was defined positive when the spar was sweeping backwards. A schematic representation
of the wing design is shown in Figure 2. A simple push-pull mechanism controlled by an electromechanical
servoactuator was designed to induce the sweeping motion of the main spar.
Downloaded by BRISTOL UNIVERSITY on April 2, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2007-1706

Figure 2. Schematic Wing Design with Hinged Spars

The structure of the AAW panel was manufactured entirely out of Balsa wood. Both the front and the
rear spar were intentionally down-sized with respect to the original design, such as to have the main spar
transfer the bulk of all the aerodynamic loads. To increase the range of sweep angles near the trailing edge,
the main spar was vertically tapered towards the tip. Here, the maximum travel of the spar spanned 60%
of the local wing chord. Rectangular cut-outs in the ribs functioned as free sliding joints and allowed the
main spar to sweep through the wing panel. Figure 3 shows the Balsa structure of the starboard wing panel,
fitted with the hinged spar in different positions. By sweeping the main spar, the flexural axis of the wing
could be shifted, allowing for active wing twisting during flight.

Figure 3. Balsa Wood Structure of Active Aeroelastic Panel; Most Aft Spar Position (Left), Neutral Postion
(Center) and Most Forward Position (Right)

The skin of the wing was not uniform over the entire wing. At the bottom side of the airfoil and over

4 of 12

American Institute for Aeronautics and Astronautics


the first 30% of the top side, Balsa wood panels with a thickness of 0.6mm (0.025in) were bonded to the
ribs. To air-tighten the wing, a thin plastic foil covered the entire wing. On top of that, a low-density elastic
polyamide fabric was shrink fitted which allowed for wing twisting and at the same time ensured the airfoil
shape at the top side of the wing during aerodynamic loading. Figure 4 shows the AAW panel.
Downloaded by BRISTOL UNIVERSITY on April 2, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2007-1706

Figure 4. AAW Panel with Polyamide Skin

The total weight of one panel amounted to 59 grams, which was lower than the weight of the original
panel (65g). The overall increase in OEW due to the fitting of the aeroelastic panels amounted to 3%.
The original outboard wing panels exhibited a taper ratio of 0.85. The AAW panels that replaced these
panels were straight. Consequently, the total wing area was increased with 3%, leaving the wing loading
(W
S ) unchanged.

III. Conceptual Modeling


A. Airfoil Analysis
A simple airfoil analysis was carried out, using a two dimensional panel method based on viscous potential
flow. The airfoil that was used in the original wing and in the modified AAW panel was a Clark Y airfoil,
which had a maximum thickness of 11% and a flat lower side. Figures 5(a) and 5(b) show the lift and
drag characteristics of the airfoil at a simulated Reynolds number of 150,000. It can be seen that stall
onset occurred at an angle of attack of approximately 16◦ . Furthermore, the maximum lift to drag ratio
((Cl /Cd )max ) was found between an angle of attack of 4◦ and 6◦ .

B. Structural Analysis
A finite element (FE) model was constructed to investigate the static and dynamic aeroelastic features of
the wing and compare with experimental results. Only the beam elements (ribs and spars) in the wing were
considered for convenience. Since the skin of the wing had a comparatively low stiffness (Eskin ≪ EBalsa ),
the skin was not considered in the FE analysis. This section presents the FE model and its properties.
Standardized material parameters for Balsa wood36 (12% moisture content) were obtained:

EL EL GLR GLT GRT


= 0.015 = 0.046 = 0.054 = 0.037 = 0.005
ET ER EL EL EL
νLR = 0.229 νLT = 0.488 νRT = 0.467 νT R = 0.231

where L denotes longitudinal grain direction. T and R are the tangential and radial direction, respectively,
with respect to the growth rings of the wood, as defined in Figure 6. E is the Young’s modulus, G is the

5 of 12

American Institute for Aeronautics and Astronautics


(a) Relation Between Lift Coefficient, Cl , and Angle of Attack, (b) Relation Between Lift to Drag Ratio, Cl /Cd , and Angle of
α Attack, α

Figure 5. Results for Two-dimensional Analysis of Clark Y Airfoil


Downloaded by BRISTOL UNIVERSITY on April 2, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2007-1706

shear modulus and ν is Poisson’s ratio. A simple bend test showed that the bend stiffness of the material,
EB = 3.4 GP a, resulting in a longitudinal stiffness of EL = 1.1 · EB = 3.74 GP a (accounts for the amount
of shear deflection during bend test). Using these relations the material properties of the balsa wood are
presented in Figure 7.

Figure 7. Balsa Properties

Symbol Amount Units


ρ 170.7 kg/m3
EL 3.74 GP a
ER 0.172 GP a
ET 0.0561 GP a
GLR 0.202 GP a
GLT 0.138 GP a
GRT 0.0187 GP a
Figure 6. Principal Axes of Wood w.r.t Grain Direction
and Growth Rings36

The wing geometry was created using MSC.PATRAN pre-processor because it can detail numerous things
in the model, and properties are easy to define. In MSC.PATRAN, the model was meshed by using a shell
elements and structural properties were assigned. Ribs, leading-edge spar, trailing-edge spar and main spar
were modeled using CTRIA3 and CQUAD4 elements. Boundary conditions prevented any rotational or
translational motion of the root section.
Once the model was meshed, a dynamic load analysis was selected and MSC.PATRAN built the structural
input file for MSC.NASTRAN. Results were post-processed in MSC.PATRAN. Figure 8 shows the shapes
of the first torsional and bending mode for minimum and maximum sweep angle.

C. Aeroelastic Modeling
For static aeroelasticity analysis ZAERO was employed. Since this software system is based on the finite
element method and has features to generate high reliability geometry it is suitable for the preliminary design
tasks that need high reliability modeling and rapid analysis.37
Since the ZAERO system does not provide the structural finite element solutions, it imports externally
computed the mode solutions generated by another structural finite element code (in this case MSC.NASTRAN).
The ZAERO system processes the finite element output file to obtain the structural grid point locations,

6 of 12

American Institute for Aeronautics and Astronautics


(a) First Bending Mode for Λ = −4.5◦ (b) First Torsional Mode for Λ = −4.5◦
Downloaded by BRISTOL UNIVERSITY on April 2, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2007-1706

(c) First Bending Mode for Λ = 9.5◦ (d) First Torsional Mode for Λ = 9.5◦

Figure 8. Mode Shapes

the coordinate transformations for relating the local and global coordinates to the basic coordinate system,
the modes, the natural frequencies, and the generalized mass matrices and aerodynamic moment center of
the structural finite element models. The stiffness of the structure is represented by the global stiffness ma-
trix. The stiffness matrix elements are based on the material type, the element properties, and the element
geometry.
To explore the static aeroelastic properties of the wing configurations successfully, MATLAB was exploited
to generate aerodynamic models as flow parameters change during trim analysis (because of fluid-structure
interaction). Geometric and flight condition variables, e.g. mass matrices and aerodynamic center were
obtained from the MSC.NASTRAN output file. Together with Mach number, angle of attack and other
flow parameters they were parameterized and coded in a MATLAB script file that generated the required
input file for ZAERO. Results from the static aeroelastic analysis were summarized using a separate code in
MATLAB.

IV. Experimental Testing and Results


A. Bench Testing
Dynamic bench testing was carried out to obtain the shift in natural frequency as a function of the main spar
position. It also served to verify the FE model that was made of the wing. During the tests the AAW panel
was positioned vertically and clamped on the top side. To induce a vibration it was excited perpendicular
to the surface at the trailing edge at the lower tip of the wing panel. During the vibration, the absolute
position of the trailing edge tip and the leading edge tip were recorded by two laser displacement probes
with an accuracy of 0.5mm at a sampling rate of 100Hz.
From the displacement response of the tip, the elastic center at the tip position could be established. It
could be seen that the leading and trailing edge displacement at the tip were in phase at all spar positions.
From the difference in amplitude between leading and trailing edge displacement, the position of the rotating

7 of 12

American Institute for Aeronautics and Astronautics


point was established as ³ ´
³x´ Ale
Ate
=³ ´ . (1)
c ec, t Ale
+1
Ate

Figure 9(a) shows that there was a linear relation between the sweep angle of the main spar and the relative
position of the elastic center at the tip of the panel. Furthermore it demonstrates that by sweeping the main
spar the elastic center at the tip could be shifted with as much as 40% of the local chord, between the 30%
chord and the 70% chord. The shift in elastic axis through the AAW panel is drawn schematically in Figure
9(b).
By exciting the structure at the tip, the wing panel exhibited a combined motion of torsion and bending.
By averaging the displacement of the leading and trailing edge at the tip of the panel, the first bending
eigen mode could be established. Analogously, by subtracting the position data for leading and trailing
edge, the first torsional eigen mode could be extracted. From these eigen modes, the natural frequencies
were determined. Figure 10 shows how the natural frequencies changed with the spar position. Apart from
Λ = 6 the torsional and bending natural frequencies were close together, with a maximum mutual difference
of 7% at Λ = 3. Accordingly, the response to the excitation demonstrated the first coupled eigen mode,
where torsion and bending were in phase. The FE model showed good correlation to the experimentally
obtained bending natural frequency. The FE prediction of torsional natural frequency showed to be off by an
Downloaded by BRISTOL UNIVERSITY on April 2, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2007-1706

average of 25%. This difference is thought to originate from the coupled behavior of the bending and twisting
during the test, while these motions were modeled as being uncoupled. Increasing the sweep angle between
−4.5◦ and 9.5◦ demonstrated a linear increase of 28% in the coupled natural frequency. This translated in
an increase in (torsional) stiffness with increasing sweep angle. At Λ = 6 the wing panel eigen mode was
different from the other positions. The natural frequency of the torsional frequency was three times the
natural frequency of the bending natural frequency.

(a) Shift in Elastic Center at the Tip Due to Main Spar Sweep- (b) Schematic Representation of Elastic Axis Sweep Through
ing AAW Panel Due to Main Spar Sweeping

Figure 9. Elastic Shift Through AAW Panel

B. Wind Tunnel Testing


Wind tunnel measurements were done on the AAW panel in one of the low speed wind tunnels at The
University of Manchester. The wing panel was clamped vertically from the wind tunnel ceiling. A six-axis
balance system could measure all the aerodynamic forces. At the tip of the wing, two lasers measured the
position of the leading and trailing edge of the wing, during the measurement. The lasers were mounted on a
metal frame which was suspended from the wind tunnel ceiling and rotated together with the angle of attack
of the wing. Accordingly, the laser beam was always perpendicular to the surface of the panel regardless of
the wing’s angle of attack.

8 of 12

American Institute for Aeronautics and Astronautics


Downloaded by BRISTOL UNIVERSITY on April 2, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2007-1706

(a) First Bending Natural Frequency (b) First Torsional Natural Frequency

Figure 10. Bench Test Results

Based on the operating weight (OEW=1400g) and the wing area (S = 0.0846m2 , preliminary calculations
showed that the cruise velocity for the aircraft would be 10m/s. Consequently, tests were conducted at a
velocity of 10m/s (Re=130,000) at angles of attack between 0◦ and 20◦ at an interval of 4◦ . Five different
positions of the main spar were considered ranging from -4.5◦ to +9.5◦ at an interval of 3.5◦ . At each
combination of spar position and angle of attack, the lift (L) was measured as well as the absolute position
of the leading and trailing edge at the wing tip. Furthermore, the wing was excited at the trailing edge wing
tip and the response was recorded so as to determine first bending and first torsional natural frequencies.
The results of the lift measurements are displayed in Figure 11(a). This figure proves quantitatively that
when the position of the main spar was altered, the lift coefficient changed. For example, at an angle of
attack of 4◦ ((Cl /Cd )max ) there was a lift increase of 35% between the most forward swept position to the
most backwards swept position. The graph also demonstrates that there was a substantial increase (20%) in
lift coefficient between Λ = 2.5◦ and Λ = 6◦ , while the increase between any other subsequent positions was
comparatively small. This showed that sweeping the main spar between these two positions was most efficient
in terms of lift change. The ZAERO results (denoted in Figure 11(a) as ‘FEM’) show similar trends to what
was measured in the wind tunnel. At 0◦ and 12◦ angle of attack the change in lift w.r.t spar position was
within 12% of what was measured. However, at intermediate angles of attack, the theory under-predicted
the aeroelastic effect and the difference between theoretical and experimental dCL /dΛ amounted to 61%
(α = 4) and 84% (α = 8). The offset of the experimental values w.r.t. the predicted values might have been
caused by inaccurate determination of angle of attack during wind tunnel testing.
From the position measurements at the wing tip, the twist of the wing could be determined. Figure 11(b)
shows that the wing twist was influenced by the angle of attack, as well as by the position of the main spar.
The flexibility of the wing caused it to twist as a result of the aerodynamic loads. As was shown earlier in
Figure 5(a), stall onset started at an angle of attack of 16◦ . During the test, stall could be observed due to
the vibration of the wing (stall flutter). The angle of attack at the tip, αt was defined as
αt = αr + θ. (2)
In Figure 11(b) a line is drawn that represents the tip angle of attack at which stall occurred. It can be seen
that the data points to the left side of this line show a greater dependency on the angle of attack than the
data points on the right side of this line. This behavior was assigned to stall behavior at the tip of the wing.
Good agreement between theory and experiment was achieved at main spar seep angles between -4.5 and
+2.5 degrees up until stall. At higher sweep angles, the experimental data showed a greater dependency of
wing twist on spar position than was anticipated by the static aeroelastic model.

9 of 12

American Institute for Aeronautics and Astronautics


(a) Relation Between Lift Coefficient, CL , and Angle of At- (b) Relation Between Wing Twist, θ, and Angle of Attack, α,
tack, α, at Various Main Spar Sweep Angles, Λ at Various Main Spar Sweep Angles, Λ
Downloaded by BRISTOL UNIVERSITY on April 2, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2007-1706

Figure 11. Static Results from Wind Tunnel Tests

The dependency of the wing twist, θ, on the main spar sweep angle, Λ, was consistent throughout the
angle of attack range. At α = 4◦ the sweeping of the main spar changed the twist of the wing with 3.7◦ .
A higher angle of attack increased the twist range that could be achieved by sweeping the main spar. For
example, at an angle of attack of 12◦ the change in twist amounted to 9.1◦ . Furthermore, Figure 11(b)
demonstrates that when the center of pressure of the wing was behind the flexural axis of the wing, a
negative twist was generated (confer Λ = −4.5), while as the center of pressure was in front of the flexural
axis, the wing displayed positive twist (Λ = 2.5 to Λ = 9.5). At Λ = −1 the aerodynamic center was virtually
on the flexural axis, since the dependence of tip rotation on angle of attack change was relatively small.

C. Flight Testing
After successful wind tunnel tests a flight test was conducted under good visibility conditions, a wind speed
of 4m/s (7.8kts) and a pressure of 1027mbar. Figure 12 shows the aircraft in full configuration. During
flight, when commanding a sweeping motion of both the spars, the aircraft showed a rolling motion. It was
shown that by giving the opposite command, the aircraft was capable of rolling back to level flight. The roll
rate was considerably lower than for aileron operated aircraft.
To increase the roll rate, a number of (partial) solutions exist. The stiffness of the AAW panels can be
further decreased to enhance torsional flexibility and consequently controllability. The span of the AAW
panels can be increased, which will result in a higher twist angle at the tip and a larger control surface,
which both increase controllability. Furthermore, the sliding joint between the main spar and the ribs could
be improved such that the main spar does not carry any of the torsional loads on the wing panel. Each of
these solutions should be investigated to further enhance the aeroelastic performance.

V. Conclusions
Two aeroelastic active wing panels, which rely on a sweeping movement of the main spar to alter their
flexural axes were designed and built to fit on a 1.8m (71in.) span UAV. The application of the modified
wing panels added 3% to both the operational empty weight and planform area, leaving the wing loading
unchanged. Dynamic bench tests demonstrated a shift in elastic center position of 40% at the tip of the
panel. Furthermore, an increase of 28% in the first coupled natural frequency was demonstrated with good
agreement with the bending natural frequency as predicted by the finite element model. Wind tunnel tests
showed at cruise condition, the peak-to-peak change in twist amounted to 3.7◦ , resulting in a change in lift

10 of 12

American Institute for Aeronautics and Astronautics


Figure 12. UAV with AAW Panels

coefficient of 35%. At higher angles of attack this increased to a maximum of 9.1◦ peak-to-peak twist. Static
aeroelastic modeling showed to be accurate in predicting the twist at spar sweep angles ranging from -4.5
to +2.5 degrees. Flight testing proved that sweeping the main spar could induce a rolling motion to the
Downloaded by BRISTOL UNIVERSITY on April 2, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2007-1706

aircraft and bring it back to level flight. However, the relatively low demonstrated roll rates leave room for
improvement.

Acknowledgements
This research was sponsored by the University of Manchester. The authors would like to acknowledge
the outstanding contribution of Mr. John Davidson and Mr. Ger May.

References
1 Vos, R., Barrett, R., Krakers, L., and van Tooren, M., “Post-buckled Precompressed (PBP) piezoelectric actuators for

UAV flight control,” SPIE No. 6173-15, San Diego, 26 February-2 March 2006.
2 Vos, R., DeBreuker, R., Barrett, R., and Tiso, P., “Morphing wing flight control via Post-Buckled Precompressed

Piezoelectric Actuators,” Journal of Aircraft, accepted for publication 13 June 2006.


3 Gano, S. E. and Renaud, J. E., “Optimized unmanned aerial vehicle with wing morphing for extended range and

endurance,” Proceedings of the 9th AIAA/ISSMO Symposium and Exhibit on Multidisciplinary Analysis and Optimization,
2002.
4 Gano, S. E., Renaud, J. E., Batill, S. M., and Tovar, A., “Shape optimization for conforming airfoils,” Proceedings of

44th AIAA/ASME/ASCE/AHS Structures, Structural Dynamics, and Materials Conference, Norfolk, VA, 7-10 April 2003.
5 Gern, F. H., Inman, D. J., and Kapania, R. K., “Structural and Aeroelastic Modeling of General Planform Wings with

Morphing Airfoils,” AIAA Journal, Vol. 40, No. 4, 2002, pp. 628–637.
6 Jha, A. K. and Kudva, J. N., “Morphing Aircraft Concepts, Classifications, and Challenges,” Proceedings of SPIE Smart

Structures and Materials, Vol. 5388, 2004, pp. 213–224.


7 Kota, S., Hetrick, J., Osborn, R., Paul, D., Pendleton, E., Flick, P., and Tilmann, C., “Design and application of

compliant mechanisms for morphing aircraft structures,” Proceedings of SPIE Smart Structures and Matererials, Vol. 5054,
March 2003, pp. 24–33.
8 Kudva, J. N., “Overview of the DARPA Smart Wing Project,” Journal of Intelligent Material Systems and Structures,”

Journal of Intelligent Material Systems and Structures, Vol. 15, No. 4, April 2004, pp. 261–267.
9 Lind, R., Abdulrahim, M., Boothe, K., and Ifju, P., “Morphing for Flight Control of Micro Air Vehicles,” Proceedings of

AIAA Guidance, Navigation, and Control Conference and Exhibit, Austin, TX, 11-14 August 2003.
10 McGowan, A.-M. R., Washburn, A. E., Horta, L. G., Bryant, R. G., Cox, D. E., Siochi, E. J., Padula, S. L., and Holloway,

N. M., “Recent Results from NASA’s Morphing Project,” Proceedings of SPIE Smart Structures and Materials, Vol. 4698, 2002,
pp. 97–111.
11 Rusnell, M. T., Gano, S. E., Perez, V. M., Renaud, J. E., and Batill, S. M., “Morphing UAV Pareto Curve Shift

for Enhanced Performance,” Proceedings of the 45th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and
Materials Conference, Palm Springs, CA, 19 - 22 April 2004.
12 Sanders, B., Eastep, F. E., and Forster, E., “Aerodynamic and Aeroelastic Characteristics of Wings with Conformal

Control Surfaces for Morphing Aircraft,” Journal of Aircraft, Vol. 40, No. 1, January-February 2003, pp. 94–99.
13 Sanders, B., Crowe, R., and Garcia, E., “Defense Advanced Research Projects Agency - Smart Materials and Structures

Demonstration Program Overview,” Journal of Intelligent Material Systems and Structures, Vol. 15, No. 4, April 2004, pp. 227–
233.
14 Amprikidis, M. and Cooper, J. E., “Adaptive Internal Structures for Active Aeroelastic Control,” International Forum

on Aeroelasticity and Structural Dynamics, Amsterdam, 2003.

11 of 12

American Institute for Aeronautics and Astronautics


15 Flick, P. and Love, M., “The Impact of Active Aeroelastic Wing Technology on Conceptual Aircraft Design,” Proceedings

of the Applied Vehicle Technology Panel, NATO Research and Technology Agency, Ottawa, Canada, October 1999.
16 Pendleton, E., “Back to the Future - How Active Aeroelastic Wings are a Return to Aviation’s Beginnings and a Small

Step to Future Bird-like Wings,” NATO RTO Meeting on Active Control Technology for Enhanced Performance Operational
Capabilities of Military Aircraft, Land Vehicles and Sea Vehicles, Braunschweig, Germany, 8 May 2000.
17 Phillips, W. F., “New Twist on an Old Wing Theory,” Aerospace America, January 2005, pp. 27–30.
18 Phillips, W. F., Fugal, S. R., and Spall, R. E., “Minimizing Induced Drag with Geometric and Aerodynamic Twist, CFD

Validation,” Presented at the 43rd AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV, January 2005.
19 Zink, P. S., Mavris, D. N., and Raveh, D. E., “Maneuver trim optimization techniques for active aeroelastic wings,”

Journal of Aircraft, Vol. 38, No. 6, 2001, pp. 1139–1146.


20 Perry, R., Cole, S., and Miller, G., “Summary of an Active Flexible Wing Program,” Journal of Aircraft, Vol. 32, No. 1,

February 1995, pp. 10–15.


21 Pendleton, E., Bessette, D., Field, P., Miller, G., and Griffen, K., “The Active Aeroelastic Wing Flight Research Program -

Technical program and model analytical development,” 39th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics,
and Materials Conference and Exhibit, Long Beach, CA, 20-23 April 1998.
22 Wlezien, R. W., Horner, G. C., McGowan, A. R., Padula, S. L., Scott, M. A., and Silcox, R. J., “The Aircraft Morphing

Program,” proceedings of SPIE Smart Structures and Materials, 1998, pp. 176–187.
23 Kusmina, S. and et al., “Review and Outlook for Active and Passive Aeroelastic Design Concepts for Future Aircraft,”

Proceedings of the International Forum on Aeroelasticity and Structural Dynamics, Vol. 2, Spain, 2001, pp. 475–486.
24 Schweiger, J. and Suleman, A., “The European Research Project Active Aeroelastic Structures,” CEAS International

Forum on Aeroelasticity and Structural Dynamics, 2003.


25 Florance, J., Heeg, J., Spain, C., Ivanco, T., Wieseman, C., and Lively, P., “Variable Stiffness Spar Wind-Tunnel Model
Downloaded by BRISTOL UNIVERSITY on April 2, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2007-1706

Development and Testing,” 45th AIAA Structures, Structural Dynamics and Materials Conference, 2004.
26 Chen, P., Sarjaddo, D., Jha, R., Lui, D., Griffin, K., and Yurkovich, R., “Variable Stiffness Spar Approach for Aircraft

Manoeuver Enhancement Using ASTROS,” Journal of Aircraft, Vol. 37, No. 5, 2000, pp. 865–871.
27 Costa, A. P. F., Novel concepts for piezoelectric actuated adaptive aeroelastic aircraft structures, Phd. thesis, Universidade

Tecnica de Lisboa, December 2001.


28 Kusmina, S. and et al., “Some Applications of Active Aeroelastic Concepts to Aircraft Design,” ICAS , 2002.
29 Johnston, C. O., Neal, D. A., Wiggins, L. D., Robertshaw, H., Mason, W., and Inman, D., “A Model to Compare the

Flight Control Energy Requirements of Morphing and Conventionally Actuated Wings,” 11th AIAA/ASME/AHS Adaptive
Structures Conference, Norfolk, VA, 7-10 April 2003.
30 Prock, B. C., Weisshaar, T. A., and Crossley, W. A., “Morphing Airfoil Shape Change Optimization with Minimum Ac-

tuator Energy as an Objective,” 9th AIAA/ISSMO Symposium on Multidisciplinary Analysis and Optimization, 4-6 September
2002.
31 Pendleton, E. W., Bessette, D., Field, P. B., Miller, G. D., and Griffin, K. E., “Active Aeroelastic Wing Flight Research

Program: Technical Program and Model Analytical Development,” Journal of Aircraft, Vol. 37, No. 4, 2000, pp. 554–561.
32 Andersen, G., Forster, E., , Kolonay, R., and Eastep, F., “Multiple Control Surface Utilization in Active Aeroelastic

Wing Technology,” Journal of Aircraft, Vol. 34, No. 4, July August 1997, pp. 552–557.
33 Wilson, J., “Active aeroelastic wing: a new/old twist on flight,” Aerospace America, September 2002.
34 Wilson, J., “A twisted approach to wing warping,” Aerospace America, October 2005.
35 Amprikidis, M. and Cooper, J., “Development of smart spars for active aeroelastic structures,” 44th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, Norfolk, VA, 7-10 April 2003.
36 Green, D. W., Winandy, J. E., and Kretschmann, D. E., Wood handbook: wood as an engineering material, USDA Forest

Service, Forest Products Laboratory, Madison, WI, 1999, General technical report FPL; GTR-113.
37 Anon., ZAERO User’s Manuals Version 7.4 , ZONA Technology Inc., 9489 E. Ironwood Square Drive, Scottsdale, AZ,

15th ed., July 2006.


38 Rodden, W. P. and Johnson, E. H., MSC.NASTRAN Version 68 Aeroelastic Analysis User Guide, MSC.Software, 2006.

12 of 12

American Institute for Aeronautics and Astronautics

View publication stats

You might also like