You are on page 1of 23

47th AIAA Aerospace Sciences Meeting Including The New Horizons Forum and Aerospace Exposition AIAA 2009-1291

5 - 8 January 2009, Orlando, Florida

Design Limitations of Deployable Wings


for Small Low Altitude UAVs

Jamey D. Jacob∗
Aerospace Engineering, Oklahoma State University, Stillwater, OK 74078, U.S.A
Suzanne W. Smith†
Mechanical Engineering, University of Kentucky, Lexington, KY 40506, U.S.A

Deployable wing technology can offer many benefits to specific UAV missions, most im-
portantly low-volume storage of the wings when the vehicle is not in-flight. Deployable
wing designs enable aircraft to be more effective by providing larger numbers in volume
restricted applications, such as air, submarine, or hand launches. As platforms become
smaller, new avenues for deployable wings arise. This paper discusses issues related to de-
ployable wing technology for small UAVs and some of the limitations for certain designs. In
particular, the differences between rigid and inflatable solutions are discussed and example
applications are presented. Like morphing aircraft technology, there is not a single solution
for aircraft requiring deployable wings. The suitability of low-volume storage on an aircraft
will be dependent on weight, span, storage requirements, and the flight envelope.

I. Introduction

Small UAS platforms are becoming increasingly important on and off the battlefield where situational aware-
ness can be dramatically increased by rapid deployment of one or more surveillance UAVs. With small UAVs,
the issue of weight fraction is not only crtical, but absolute weight value as well, where technology avail-
able at a larger scale can ground a small scale vehicle if the technology weight does not scale down faster
than the size. This is particularly true for hand launched UAVs and MAVs, where differences measured
in ounces can have a dramatic impact on vehicle performance. When packing is added as a requirement
whereby the wings must be stowed to save volume, these restrictions become particularly difficult to design
around. However, packing can add tremendous benefit to certain missions, such as tube-launched UAVs
where more vehicles can be deployed if the wings can be stowed when not in-flight. Other aircraft designs
that benefit from highly packable wings include backpackable UAVs ISR applications both on and off the
battlefield,1–3 balloon- or rocket-assisted deployment of scientific payloads4 and extraterrestrial aircraft for
flight in non-Earth atmospheres such as Mars, Venus, and Titan.5–10

Requirements for deployment comes from many avenues, but much like morphing inspiration has primarily
been biological in nature.11–13 The two basic approaches to designing a deployable wing include (i) mechanical
folding and (ii) inflatable structures. Each bring their own design issues to the table. Whereas folding wings
are typically volume limited, inflatable wings are characteristically mass limited. Thus, the use of inflatable
wings over folding wings can be examined in the vehicle design trade study depending on whether one desires
to minimize mass or minimize volume. Additional rigid designs are also possible, including nesting, extending
and flexible folding, and will also be examined.
∗ Associate Professor; Associate Fellow AIAA; jdjacob@okstate.edu.
† Donald and Gertrude Lester Professor; Associate Fellow AIAA; ssmith@engr.uky.edu.

1 of 23
Copyright © 2009 by the authors. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.

American Institute of Aeronautics and Astronautics


II. Deployable Wing in Design Concepts

A. Overview

A broad array of UAVs with deployable wing technology has been developed and some are discussed here.
Since specific vehicles are discussed throughout the article in relation to their deployment concept, a rigorous
survey is not provided in this section. Strictly speaking, our interest is not in morphing aircraft but in wings
that deploy from one fixed state to another in a short amount of time. Morphing aircraft have been reviewed
elsewhere so will not be discussed in detail here.12 The deployment is usually but not always irreversible in-
flight. As such, variable wing geometry (VG) refers to rigid wings that change simple geometric properties in
flight, such as sweep or area. A wing that is variable geomety but not stowable is not considered deployable
wing technology. Thus, a deployable system is not necessarily variable geometry, such as a wing that deploys
from a stowed state prior to flight to a fixed geometry during flight.

Figure 1. NRL VRS timeline.14

NRL has been at the forefront of deployable technology and has developed several pioneering concepts that
have been successfully field tested.14 These include several folding wing models such as FINDER, ALICE
and FLYRT as well as Sail-a-Plane whose folding wings served as both horizontal and vertical lifting surfaces
in flight and sailboat modes, respectively (Fig. 1). ALICE is a good example of the diverse capabilities. It
was an unmanned platform that could be air launched from a tactical aircraft at speeds up to 0.8 Mach and
altitudes up to 45,000 ft. After launched from the tactical aircraft, ALICE glides using tail control surfaces
until it reaches a speed of approximately 250 knots. The rigid wing and propeller then deploy and the heavy
fuel engine starts. ALICE cruised approximately 200 nautical miles in one hour before the outer rigid wing
panels deploy for loiter. In the loiter mode, it operated at 65 knots with a two hour endurance carrying a
25 lb payload. Research efforts included development of the polymorphic wing, a JP-8 fueled rotary engine,
a high efficiency starter/generator, a folding variable pitch propeller, and an advanced EW payload.

Similarly, Raytheon’s Morphing Program proposed an adaptive wing structure for their Tomahawk cruise
missile that could travel to the target area at high speed, loiter and then move to another target area, at
speed changes from 0.3 Mach and 3.0 Mach. A Multi-Purpose Unmanned Aerial Vehicle (MPUAV) was also
briefly considered by Lockheed Martin. The idea is that the drone could handle all-weather reconnaissance,
battle damage assessment, or specialized mission support for the submarine. This UAV could be launched
from a submarine, then perform wing shaping for a different mission and return to the boat. Many more
platforms in a similar vein have been developed since, some examples of which are discussed below.

2 of 23

American Institute of Aeronautics and Astronautics


B. Design Concepts

Their are several options available when designing deployable wings and an equal number of ways to categorize
the design types, such as flexible versus rigid or root deployment versus span extenders. The options can
be crudely binned into one of the following types: Folding, Inflating, Nesting, or Extending (FINE). While
this categorization is rudimentary, it serves its purpose in breaking down the wing designs based upon
type of deployment, albeit with some limitations such as placement of compliant deployment methods not
discussed in detail here. (Landon provides a thorough categorization of deployment methods for compliant
mechanisms.15 ) Note that each category may have multiple designs associated with it. For example, folding
can be either rigid or flexible folding and the fold may be along the span (“dihedral”) or along the chord
(“sweep”). We will discuss in turn various design concepts and their practical limitations for rigid mechanical
designs, inflatable designs, and hybrid concepts that integrate both rigid and hybrid elements.

1. Folding

Folding wings can be either rigid where the folds are at discrete hinge locations or flexible where the fold is
continuous along the span. Some examples are shown in Fig. 2. When discussing rigid folding wings herein,
we refer to lifting surfaces that are stowable and thus do not include variable geometry or morphing wings
as these are designed to operate during all phases of flight and will not satisfy the stringent deployment
requirements needed for a small UAS. Design limitations for morphing are discussed elsewhere.16 While
folding wing designs have received little attention in the literature due to their limited use, they do have
a rich history in both manned and unmanned vehicles. Folding wings are common among manned naval
aircraft, but also include some land based aircraft such as the Su-47. Though uncommon, some aircraft have
incorporated folding elements in the vertical tails to reduce height to fit into hangars.

Figure 2. Rigid wing concepts including SLAM-ER (scissor fold), Coyote (tandem scissor fold) and ALICE
(swing wing with outboard hinge).

The most straightforward concept for stowing aircraft wings via folding employ hinged rigid wing sections
that fold to reduce parked area on carrier decks or in other limited-area circumstances. Examples are included
in Fig. 3. Single-hinge concepts are seen in early military aircraft, but also in contemporary designs. As
the images illustrate, the spanwise length of the aircraft is reduced sometimes by as much as 2/3, while
the vertical height required for the stowed aircraft may be increased by a factor of 2 to 3 times that of the
deployed configuration.

Increased stowed height is addressed through the use of complex folding mechanisms that enable the wings
to rotate while folding into place. From the 1944 image of the Grumman F6F-3 ”Hellcat” fighters in Fig. 4,
to the private folding-wing designs and roadable aircraft concepts in Fig. 5, low-profile folded rigid wings
typically include the folding mechanism at the wing root, rather than further out along the span.19 These
folding configurations are often constrained by fuselage length and tail interference, but successfully eliminate
unusable stowed volume between the folded wings in the single-hinge designs.

An overview of requirements for folding wings on UAVs was presented by Foch from the perspective of the

3 of 23

American Institute of Aeronautics and Astronautics


Figure 3. de Havilland Sea Vixen (top); Hawker Sea Fury, Hawker Sea Hawk, Douglas Skyraider (middle);
Super Hornet Strike Fighter (middle).

NRL.17 Part of the design requirements that fit into the overall vehicle design considerations and mission
requirements include the package volume, package shape, package mass, and airborne deployment. Vehicles
such as FLYRT and FINDER demonstrated the ability to deploy from stowed volumes that were fraction-
ally larger than the fuselage to relatively high aspect ratios. Through experience, NRL learned that the
aircraft configuration has a major impact on deployment and packaging (shape, mass, volume) but has a
minor impact on the vehicle aerodynamic performance. In terms of mechanical design, hinges and pivots
proved more reliable than sliding or telescoping elements. System reliability is inversely proportional to
the number of deployment events, thus these were limited as much as possible when reliability was a key
performance requirement. In many of their designs, both aerodynamic and inertial loads were exploited
to aid deployment, eliminating the need for complex mechanical, electromechanical or chemical deployment
systems. Importantly, packaging folding rigid-structure aircraft becomes volume limited before mass limited,
thus the vehicles are not optimized for weight, but restricted in design due to the volume that the stowed
wings take.

Low-profile stowed volumes have been achieved for small unmanned aircraft that deploy in-flight using
rotating wings with unfolding sections, such as those on the FLYRT, FINDER and ALICE .18 Initiated in
the early 1990s, FLYRT demonstrated viability of the concept, while FINDER continues in use today. A
rotating centerspan section is length-limited; unfolding wing tips approximately double the wing span from
that of the rotating section as shown later. A similar two-stage deployment concept is seen in the air-launched
countermeasure ALICE, which includes a rotating cruise wing section with extending loiter wing panels.

4 of 23

American Institute of Aeronautics and Astronautics


Figure 4. Grumman F6F-3 ”Hellcat” fighters landing and stowed on the USS Enterprise, 1944.

Figure 5. Stowed configuration concepts for storing or towing personal aircraft: examples include the Stits
Playmate, EuroFox two-seat, and SYNERGY roadable aircraft hydraulic wing.

A number of patents have been granted on folding wing designs, particularly for roadable vehicles (though
shockingly, no successful flying car yet exists).19 For example, folding wings can be taken to extreme designs
by utilizing multiple articulated elements. One design uses a frame with a pair of span-wise beams at the
leading and trailing edge spaced by one or more ribs along a flexible covering material over the frame that
defines the shape of the wing. Each segment is pivotally connected to an adjacent segment at a joint and a
flexible actuation line extends past each joint and is attached to a portion of the beam. Tension in each line
keeps the wing in its extended configuration while stops limit rotation. While stowed volume is small, this
is essentially a collapsible hang glider wing, thus shapes are limited to thin airfoils with associated fabric
deformations with the tension lines creating an unacceptable amount of drag for optimal aerodynamics.

The aeroelastic response of rigid folding hinged wings has been examined by Radcliffe and Cesnik.20, 21 The
wing semi-span sections were allowed to rotate about the hinges such that aerodynamic loads kept the wing
straight. It was found that adding hinges to a wing significantly altered the speed at which instabilities first
appear and the stiffness of the hinges plays a significant role in the flutter speeds. Reducing hinge stiffness

5 of 23

American Institute of Aeronautics and Astronautics


(and by association probably hinge weight) increase first bending and torsion instability speeds, but values
of the parameters exist where second instability modes appear at much lower speeds than the first mode.
The wings enter chaotic or limit cycle oscillation at speeds near the flutter speeds which limites the energy
added to the system while it is in an unstable state.

Compliant mechanisms and their role in deployable air vehicles was the subject of Landon’s recent thesis.15
Overall wing compliance and the resulting wrapped-fuselage packing option is seen in the Florida bendable-
wing MAV. The entire class of micro-aerial vehicles (MAVs) has generated new packaging strategies, making
use of bendable wings such as the University of Florida MAV shown in Fig. 6.22, 23

Figure 6. UF bendable-wing endurance MAV23

An alternative to developing a compact (low-profile) stowed volume for individual aircraft in a multi-unit
mission is to employ a tessellating packing strategy to used volume available from one stowed unit for
adjacent systems.24 The NRL CICADA employs such a strategy to package large numbers of expendable
micro UAVs (Fig. 7). Additional designs leverage folding mechanisms to provide non-flight benefits being
storage, such as power generation using solar an expanded array of solar cells in DII’s MAV concept (Fig. 8).

Figure 7. NRL CICADA packing, prototype and in-flight illustration.

(a) Closed - flight mode. (b) Open - perch mode.

Figure 8. DII MAV solar cell clam shell wings.

6 of 23

American Institute of Aeronautics and Astronautics


2. Inflating

The use of inflatable structures to realize a significant aircraft packing advantage has a long heritage stemming
from the 1950s and the Goodyear Inflatoplane.25 The aircraft seen in Fig. 9 was packaged into a crate that
measured 6-feet long by 3-feet wide by 2-feet high a stowed volume of 36 cubic feet. From the scale drawings,
an estimate of the inflated structure deployed volume is 202 cubic feet, resulting in a stowed-to-deployed
volume ratio of 0.178 (i.e., 17.8%). This estimate does not account for the motor and other subsystems that
are in the crate, so this ratio is larger than a more accurate estimate would produce. This small stowed-to-
deployed ratio is because the Goodyear Inflatoplane had only two rigid structures: the motor mount and the
landing-gear interface.26 Note that wing bracing was incorporated in the design to supplement the structural
stiffness of the inflatable wing.

Figure 9. Goodyear Inflatoplane dimensions and in-flight.

Inflatable wings maintain structural stiffness by using internal pressure to withstand the applied loads. The
primary consideration for failure in an inflatable structure is the maximum sustainable bending moment.
For an inflatable wing solution to be practical for a HALE vehicle, the inflation pressures required to sustain
the root bending and other associated forces on the wing (such as twist) must be small enough that the wing
design does not impose severe inflation and material requirements. This limit is somewhat arbitrary since it
is dependent on the wing design and the inflation system. In reality, this limit must be balanced with other
constraints, such as endurance requirements and available weight. Naturally, higher pressures will increase
mission risk by reducing the reliability of inflatable components through increased chances of leaks. While
leakage effects can be countered with make-up gas and onboard compressors, these must be factored in as
increased weight during the trade study portion of the vehicle design.

To determine the load carrying capability of an inflatable wing design, we begin with the well known Euler-
Bernoulli beam equation that relates the beam deflection with applied moment and material properties on
a cantilever beam;27
d2 y M (x)
2
= (1)
dx Ew I
where M is the applied moment, Ew is the Young’s modulus of the material, and I is the cross-sectional
moment of inertia. Main et al. modified this with respect to an inflated fabric tube to develop a relation for
the bending moment equation for a single inflated fabric spar for space based inflated structures.28, 29

d2 y M πpr3
= for M < (1 − 2νt ) (2)
dx2 El πr3 2

d2 y M − 2νt pr3 sin θ0 πpr3


2
= 3
for M > (1 − 2νt ) (3)
dx El r [(π − θ0 ) + sin θ0 cos θ0 ] 2
where M is the bending moment. El is the longitudinal fabric tensile modulus, r is the beam radius, and
θ0 is the wrinkle angle. The relation includes the impact of wrinkling and accounts for the biaxial stress in

7 of 23

American Institute of Aeronautics and Astronautics


the beam fabric and the impact that it has on the wrinkling threshold of the beam as well as the beam’s
postwrinking bending behavior. The model was well validated against experimental data of fabric beams
under loads. As the load increases, the beam deflects in a linear manner. Once the wrinkling threshold is
reached, the relation becomes non-linear. Soon after, the beam buckles. This will scale depending upon the
type of structure involved.While buckling is the failure mode, the onset of wrinkling indicates the maximum
design load and will be used for the design limit. It should be noted that unlike metal or composite rigid
structures that will either plasticly deform or crack, respectively, once the yield stress is reached, the inflatable
beam will bend, but then will return undamaged to its original state after the load is reduced or removed.
Inflatable sections can either be made out of plastic or fabric material with fabric tubes either woven or
braided, the latter typically referred to as “airbeams.”30

In terms of required inflation pressure for a given bending moment, the equation can be written as
2M0
P = (4)
πr3
While P is linear with M0 , doubling the tube radius reduces P by a factor of 8. Thus, larger diameter tubes
are extremely beneficial when used on wings, taking advantage of the current design of thicker HALE airfoil
profiles.

The required pressure can be reduced by 1/2 by using a braided beam that groups the active axial fibers to
double the moment of inertia in the bending direction.30
M0
P = (5)
πr3 (1 − 2 tan2 β)
where β is the angle of the bias braid. Further increases in the allowable bending load can be made by
applying the principle of tensairity where tension and compression elements are designed integrally with the
airbeam.31 While not extensively tested, it has the potential to increase the allowable load by an order of
magnitude. Other methods to increase allowable bending moment include increasing the material’s elastic
modulus and the wing’s cross-sectional moment of inertia. Multiple spars or baffles accomplish the latter.

The simplest inflatable wing design consists of a single inflatable tube used as the main wing spar about
which the rest of the airfoil is supported.32, 33 Crushable foam is used as the remainder of the wing support,
which must support the chordwise dynamic pressure loads. Multiple spars can be used to increase the wing
loading and serve as the chord wise wing box.34 Use of cylindrical spars, though a simple design, results in
a requirement for very high inflation gas pressure while still allowing for substantial deflections at the tip as
shown below. These can be reduced by using a tensioned skin, but the gas weights are still non-negligible
and thus deleterious for small UAVs or where weight constraints are tight.

Using a multi-spar approach, wing stiffness is dictated by internal pressure and the modulus of elasticity of
the restraint material, and can be elevated through the use of higher strength and modulus materials that
can withstand higher internal pressures.1 The cross-section of this type of airfoil reveals that the geometry of
the wing is defined by a series of intersecting cylinders. The use of internal spars to separate the upper and
lower external restraints yields a wing that optimizes the cross-sectional moment of inertia of the wing by
maximizing internal pressurized area. This yields the lowest possible internal pressure required for the wing
which in-turn yields lower potential for leakage, lower inflation system mass, and a lower packed volume. The
design allows for inflatable wings that can be constructed for any practical wing loading (W/S) requirement.
The important limits include the maximum bending moment due to the aerodynamic loads and inflation
pressure. Note that the relations may be non-linear depending upon the wing design. Further increases
in wing loading are possible using rigid tension and compression elements, but some of the advantages of
packing are lost.1 Rigidization of the wings post-deployment using epoxies impregnated within the wing
fabric are also possible and have been demonstrated should extra strength and a one-time use be required,
but this also creates an increase in the wing weight.

Proven wing designs are shown in Fig. 10 based around a NACA 4412 profile. These include a single
cylindrical spar (Fig. 10a), multiple cylindrical spars (Fig. 10b), baffled without a skin (Fig. 10c), and

8 of 23

American Institute of Aeronautics and Astronautics


Figure 10. Examples of possible inflatable wing configurations for a NACA 4412 profile. (a) Single spar. (b)
Multiple spar. (c) Baffled. (d) Baffled with skin.

baffled with a skin (Fig. 10d). Each of these concepts has been demonstrated in flight in some form or
another.1, 4, 30, 35 In each design, wing stiffness is determined by the internal pressure and modulus of elasticity
of the material. Since the nature of an inflatable softgood is to assume a cylindrical shape, the supporting
inflatable spars need to be integrated into the airfoil. In the cylindrical spar designs ((Figs. 10a and b), the
surrounding space is typically filled with crushable foam. This material only needs to sustain the chord wise
loads while the main spar sustains the bending and torsion loads. The main spar diameter is determined by
the airfoil’s maximum thickness. By using multiple spars (Fig. 10b), the profile of the airfoil can be outlined
requiring a lower amount of fill material and also reducing the required inflation pressure. The baffle designs
of Figs. 10c and d are an extension of Fig. 10b, where the individual baffles are defined by larger diameter
intersecting cylinders. The use of internal spars to separate the upper and lower external restraints yields
a wing that optimizes the cross-sectional moment of inertia of the wing by maximizing internal pressurized
area and also minimizes the amount of fill material required so it increases the possible packing volume
ratio.1

NASA’s I2000 was a small-scale unpowered inflatable wing aircraft tested and evaluated at Dryden in 2001.34
The inflatable wing used in this program was designed and fabricated by Vertigo, Inc. for the Navy. This
64 inch span inflatable wing contains five high pressure inflatable cylindrical spars than run span wise from
tip to tip and the chord is 7.25 inch. Between the spars and to the trailing edge of the wing is open-cell
foam bonded to the spars and to a rip-stop nylon outer skin. Additionally, a rib at each tip rigidly connects
all the spars to establish wing torsional stiffness. Thermally activated adhesives are used to bond the spars,
foam, and the nylon skin into a contiguous wing structure. The airfoil profile is a NACA 0021 and does
not contain any control surfaces, therefore, full axis control was affected only by the tail. This inflatable
wing consists of a manifold at the center of the wing to hold the wing spars in position. A small commercial
off-the- shelf tank (COTS) with a volume of approximately 35 cubic inches was selected as the high pressure
tank source. Nitrogen gas was selected as the gas source at 1800 psig and was reduced to a wing pressure
between 150 psig to 300 psig using an adjustable pressure regulator. By using this high pressure inflation
system, the inflatable wing was rapidly deployed in-flight within less than a second.

(a) t =0 s. (b) t = 2 s. (c) t = 5 s. (d) t = 9 s.

Figure 11. Inflatable wing with deployment control α = 12◦ .

Rapid deployment may be neither possible nor desirable in some situations. By balancing some external
restraint force against the internal pressure, deployment can be controlled, however. Fig. 11 show the
controlled deployment of an inflatable wing under aerodynamic loads using deployment control mechanisms,
in this case, Velcro.36 Fig. 12 reveals the gradual and controlled nature of the deployment, with lift gradual
increasing during the constant pressure phase to its maximum value at full deployment.

9 of 23

American Institute of Aeronautics and Astronautics


(a) Lift coefficient. (b) Pressure history.

Figure 12. Inflatable wing with deployment control.

Examples of designs for small UAVs using inflatable wings are shown in Fig. 13. A design of a backpackable
UAV using inflatable wings is shown in Fig. 13a. This prototype UAV allows a 10+ lb vehicle with a 5
foot span to be stored in under a 6 inch wide space. The deployment involves 2 steps: deployment of the
tail (mechanical or inflatable) and inflation of the wings using either an onboard compressed air tank or
a separate compressed air tank or compressor. The latter option saves weight, obviously, but also allows
multiple UAVs to be deployed using a single “launch” system. In the design shown here, the tail is rigid,
but this can also be constructed of inflatable components to further reduce the space required by the tail
while also increasing the robustness of the empennage. This concept is extended to in-flight deployment for
aspect ratio morphing in the In-Flight Inflatable (IFI) show in Fig. 13b.36

(a) Backpackable UAV. (b) In-flight inflatable UAV.

Figure 13. Inflatable wing concept vehicles.

3. Nesting

Nested wings are a subset of deployable wing designs, falling somewhere between a deployable and morphing
design. As such and due to their limited design space, they will not be discussed in depth here. Multiple
wings are nested within one another such that a single wing is formed by the separate structures, detaching
when needed. A proposed closed or nested wing configuration airfoil designed by Mueller and Noffke is
shown in Fig. 14a.37 Two separate wings are designed such to conform in high speed flight, then separate
vertically in a biplane configuration for low speed flight. Tests performed at 3 different Reynolds numbers
in 2-D and at Re = 200,000 for 3-D on the nested design are shown in Fig. 14b.

The embedded wing concept (Fig. 15) is a variant of the nested wing design, except a smaller wing (usually

10 of 23

American Institute of Aeronautics and Astronautics


(a) Wing design (b) Drag polar.

Figure 14. Nested wing design.37

for dash) is now completely internal to the larger loiter wing. This allows greater flexibility in designing both
wing profiles and planforms with the constraint that the dash wing must be completely embedded within the
loiter wing. Loads from the loiter wing are transferred to the skin and hard points to the dash wing. Some
complexity is added since the loiter wing must be ejected outwards towards the wing-tip and make a clean
deployment from the dash wing. This is a mechanical design issue rather than a limitation of the concept.

Figure 15. Embedded wing design.

4. Extending

The extending or telescoping method of deployment has been investigated, at least conceptually, for a long
period of time as shown by the numerous patents for roadable vehicles employing this wing deployment
concept (Fig. 16a).19 The most common approach uses pressurized or motorized telescoping spars driving
either external or internal wing segments (usually hollow shells). The major limitation of the design is that
the smallest undeployed volume is limited by the largest wing segment size. Thus, the minimum size for a
2 segment wing is b/2, b/3 for a 3 segment wing, b/4 for a 4 segment wing, etc. As the segment number
increases, so does wing weight due to the increase in telescoping components, though the addition is normally
not linear. Chord is also limited such that ci < co . Care in the design must be taken to ensure that the
telescoping elements do not bind under load when deployed and it the toughest mechanical design hurdle.

The University of Maryland is currently developing a small scale morphing aspect ratio wing using an
inflatable telescopic spar that may be possible to develop a UAV with variable aspect ratio wings.38 The
telescopic wing consists of three concentric circular aluminum tubes of decreasing diameter and increasing
length that deploy under pressure to produce various wingspan configurations. This wing could extend or
retract its span from 7 inches to 15 inches. While both binding and friction are issues, these can be overcome

11 of 23

American Institute of Aeronautics and Astronautics


for in-flight deploymemt as shown by the Akaieg Stuttgart FS-29 sailplane that used a telescoping wing
powered by a hand crank allowing a change in span from 13.3 m to 19 m in flight (AR from 21 to 28.5) as
shown in Fig. 16b.

(a) AFA telescoping spar design for flying car. (b) FS-29 sailplane with extending wings.

Figure 16. Extending wing deployment method.

5. Hybrid Design Concepts

Hybrid concepts combine elements of multiple designs, such as rigid and inflatable. Four wing concepts
have been investigated to achieve increased high aspect ratio while maximizing the low wing volume storage
capability and providing additional structural stiffness of the inflatable wing structure. Currently, some of
the concepts have been successfully fabricated in the lab for wind tunnel and flight testing, while others are
still in the conceptual design stage. Each concept will be introduced separately.

Figure 17. Inflatable wing attached to the rigid wing tip.

The simplest hybrid consists of two sections: a rigid wing inboard section and inflatable wing outboard section
as shown in Fig. 17. The rigid wing has the identical airfoil profile as the inflatable wing. Aerodynamically
designed lightweight-external pods are required to store the rolled up inflatable wing while reducing drag
during flight. Upon deployment, the pod cover will be separated from the rigid wing. However, unsteady

12 of 23

American Institute of Aeronautics and Astronautics


flight may occur upon deployment if using an onboard inflation system that produces slow flow rate. Based
on wind tunnel deployment test, we observed that the inflatable wing was pushed backward by air stream
upon deployment.39 The inflatable wing was folded in half and laying flat on either top or bottom of the
rigid wing while the inflatable wing is slowly being inflated to its final shape. Therefore, a fast inflation
system that could fill the wing within a fraction of a second is needed to overcome or minimize the unsteady
flight issues upon deployment. Additionally, deployment can be controlled by using restraints integrated
into the wing as mentioned previously.36 Fig. 13b shows the incorporation of these concepts into a flight test
vehicle An additional concept is simply a combination of a rigid folding wing integrated with an inflatable
wing. This method helps reduce the rigid wing span by replacing part of the wing with an inflatable section.
This combination helps minimize the storage volume and increase the wing span after deployment and can
optimize both mass and volume limitations. This can be accomplished in a number of ways, such as show
in Fig. 18 where an inflatable wing is integrated with a rigid folding wing.

Figure 18. Hybrid folding rigid wing with inflatable wing section

(a) Hybrid wing stowed. (b) Hybrid wing inflated.

Figure 19. Hybrid Wing Design

Other combinations of inflatables and rigid elements are straightforward and offer many options. As shown
in Fig. 19 above, a hybrid wing design is a combination of three sections: a leading edge rigid section, rigid
spar and inflatable wing trailing edge. The carbon fiber spar has an I-beam shape with a hollow space
in-between that allows air hoses and servo wires to go through along the entire span of the wing. The spar
also provides storage space to store the deflated inflatable wing. To enhance the design of this hybrid wing

13 of 23

American Institute of Aeronautics and Astronautics


leading edge section, the whole foam and spar section can be made into one piece using carbon fiber sheets
using a male mold method to shape its contour. This enhanced design can provide more storage space to
store the deflated inflatable wing while maintaining its airfoil leading edge shape and providing structural
strength.

6. True Morphing

While wing deployment can be considered a type of morphing, wing morphing in its truest sense changes the
shape of the wing continuously in-flight. However, many of the current morphing concepts under investiga-
tion display similar deployment methods. Armando provides a thorough survey of current wing morphing
technology.12 As wing morphing technology has gained interest due to the increased use and required per-
formance of unmanned aircraft vehicle and military aircraft, DARPA’s Morphing Aircraft Structure (MAS)
Program has funded proof-of-concept projects with Lockheed-Martin, Raytheon Missile System and NextGen
Aeronautics.

Lockheed-Martin Skunk Works’ Agile Hunter used a folding ‘z-wing’ configuration that has the ability to
significantly alter its rigid wing shape to modify the flight performance characteristics to accommodate a
variety of missions including combat and reconnaissance. While no longer under active development, it could
perform a 150% change in surface area through the morphing process transitioning from full span extension
to a position resembling that of a bird’s folded wing. NextGen Aeronautics developed and flight tested
a sliding skin concept that demonstrated a wing area change of 145%, aspect ratio changes of 440% and
thickness/chord changes of 280% on a modified Ryan Firebee Drone as shown in Fig. 20 below. Both of
these made substantial use of compliant mechanisms and materials.

(a) Lockheed-Martin. (b) NextGen Aeronautics.

Figure 20. Morphing concepts.

III. Design Limitations and Requirements

There will never be a single solution for deployable wing aircraft. The suitability of a particular type of
deployable technology that is integrated into any given UAV will be dependent on weight, span, storage
requirements, and the flight envelope. Integration requirements include matters such as weight (mass)
limitations, pre- and post-deployment span requirements, stowed volume limitations, and in-plane and out-
of-plane loads. Some of these are discussed below. While the focus is on small low altitude UAVs, the
concerns are similar for other vehicle classifications though with stricter requirements such as higher span
(HALE) and higher wing loading (cruise missiles). These primarily impact weight fractions when sizing
actuators.

14 of 23

American Institute of Aeronautics and Astronautics


A. Aerodynamic and Structural Load Requirements

A quick survey of small UAVs show that the majority are designed for long endurance (or at least as long
as possible). These small UAVs have unique requirements that include high aerodynamic efficiency coupled
with operation at low Reynolds numbers due to the low speeds and chord lengths. The present focus is on
aircraft that optimize aerodynamic efficiency to maximize vehicle endurance and have a reasonably useful
payload fraction.

The most obvious characteristics of long endurance UAV wings are their high AR and wing profile to achieve
high L/D.. Both of those attributes work against use of most deployable wing technology, including folding
and inflatable wings, but that will be discussed in detail below. Regardless, the only passive way to achieve
high L/D is to use a high lift airfoil coupled with a high aspect ratio. L/D can be determined via the
required thrust during cruise condition, such that
1
L/D = (6)
qCDo /(W/S) + (W/S)/(qπARe)

which can approximated by r


πeAR
L/D ≈ (7)
4KCDo

Irrespective of design, all deployable wings on aircraft increase span, thus the wing deployment will affect
the drag and lift characteristics, particularly induced drag. This section here will briefly discuss the impact
of variable planform on drag at low speed flight. The analysis presented here will follow Inman et. al’s
discussion of variable span effects for morphing UAVs.40 Consider the un-swept planform geometry shown
in Fig. 21. The planform area can be written as

S = S0 + 2∆bc (8)

The wing planform is low aspect ratio when the inflatable wing is stowed, thus the rigid wing planform alone
is used for lift. This simplification makes ∆b = 0 and c is the chord length of the outer wing section. The
reference area Sref will be defined as S0 . For the rigid wing planform shown in Fig. 21, the profile drag can
be written as
2∆bc
Dp = qSref Cdp + 2∆bc(1 + ) (9)
Sref
where q is the dynamic pressure and Cdp may be considered a 2D profile drag coefficient which is a function
of the Reynolds number. The induced drag for an elliptic load distribution can be written as

W2
Di = (10)
πqb2

where W = L and b is the total wing span (span efficiency is neglected). The total wing span for the unswept
configuration can be written as
b = b0 + 2∆b (11)
where b0 is the total span of the unextended-span, un-swept configuration. The total drag can be written
by combining Eq. 9 with Eq. 11 as

2∆bc L2
D = qSref Cdp + 2∆bc(1 + )+ (12)
Sref πqb2

This equation may be used to determine the ∆b required that achieves a given L with minimum D. This
is done by taking the derivative of Eq. 12 with respect to ∆b and setting it equal to zero. The resulting
equation for ∆b is as follows " #
1 2W 2 1
∆b = ) 3 − b0 (13)
2 πq 2 Cdp c

15 of 23

American Institute of Aeronautics and Astronautics


Figure 21. Illustrated planform parameter for an unswept configuration.40

2
This equation shows that for minimum drag, ∆b varies as W 3 . Note that Cdp is in the denominator of Eq. 13.
This indicated the expected result that as Cdp decreases, meaning the penalty for wetted area decreases, the
∆b for minimum drag increases.

Once the aerodynamic loads have been determined, the structural components can be sized. The wing must
be able to withstand both shear loads and bending moments generated by the aerodynamic forces. This
is the direct determinant of wing weight. As shown in Fig. 22a, both of these drop significantly outwards
toward the wing-tip. This will have an impact on sizing (hence, weight) of actuators, hinges, latches and
other mechanisms of the deployable design. Moving the actuation from the root to mid-chord can reduce the
bending moment by over a factor of 3, hence this drives concepts outboard. For small UAVs, gust loading
is a greater concern than with larger aircraft since gust velocities may be of comparable magnitude with the
flight speed. The load path along the wing will include transfer points along the deployable structure. Unlike
like a conventional wing that can carry much of the aerodynamic load in the stressed skin, many deployable
designs have to pass loads along the span through discrete points in the load path. This is particularly true
with folding and telescoping designs where the load must be transferred directly through hinges and spar
joints, respectively, increasing their weight.

(a) Aerodynamic shear load and bending moment on wing. (b) Deployment weight fraction with increasing n.

Figure 22. Load and weight requirements.

Following Weisshaar’s discussion of morphing requirements, we can demonstrate the weight requirements
of deployment technology as follows.11 The gross vehicle weight WG is the sum of the empty weight, fuel
weight, and payload weight
WT OGW = WE + Wf + Wp
The empty weight is given by
WE = mWG + ∆Wdeployment

16 of 23

American Institute of Aeronautics and Astronautics


where m is the empty weight fraction based upon vehicle design and technology level. Similarly, the deploy-
ment weight fraction is defined as
∆Wdeployment
fdeployment =
WT OGW
Since we are concerned here with deployment, the deployment weight can be estimated by
∆S
∆Wdeployment = n
S
where n is a function of parameters such as wing geometry (taper, t/c), wing loading and deployment method.
Assuming that n is constant with ∆S for a given design, as n increases, deployment weight fraction can be
significant and detrimental, as seen in Fig. 22b. This also drives deployment outboard and to lower values
of deployable span fraction.

B. Stowed Volume Considerations

Low volume storage requirement is a prime factor in determination of deployable concept. When rigid wing
sections are deployed, that the stowed volume cannot ever be less than 100% of the deployed wing volume.
Packing an aircraft into a specific constrained volume whether a cylindrical-shaped volume for a missile-
launched UAV or a cone-shaped volume for a Mars airplane will result in unused volume among the rigid
sections. The best that could be achieved with folded rigid sections would be to minimize the volume of
unused space, so that the packed volume is minimized to be as close to 100% of the deployed volume as
possible. A rectangular geometric envelope is a useful construct for understanding and comparing stowed
volume attributes of different deployable wing concepts. Consider, for example, the situation of enclosing the
single-hinge folded wings seen in Figure x. Each wing in its deployed configuration would have a rectangular
envelope defined by the total span length, b = b0 + 2∆b, the chord length, c, and the airfoil thickness, t, and
consequently a deployed volume of Vd = bct. If the folding sections of the wing, each with length ∆b > b0 /2,
fold into the triangular stowed positions seen in many of the images, the rectangular geometric envelop will
have height h = [(∆b)2 (b0 /2)2 ]1/2 . The resulting rectangular envelope will have a wing stowed volume of
Vs =√b0 c(2t + h). As an illustrative example, if t = 0.2c and AR = 10, with b0 = 1/3b, and ∆b = 1/3b, then
h = 3/2b and the resulting stowed volume is more than 5 times that of the deployed wings (i.e., Vs /Vd > 5).
The wing span length Ld = b is decreased by a factor of three to the stowed span length Ls = b0 = b/3 , so
if spanwise packing reduction is the only goal, then the single-hinge folding wing may provide the required
stowed efficiency (i.e.,Ls /Ld = 1/3).

Note that if ∆b > b0 /2 and h = 0 and the fold is 180◦ , so that only double the wing thickness remains as
the vertical dimension of the envelope, then the stowed volume essentially equals the deployed volume: half
the span and twice as thick. This is essentially the result achieved for the NRL FINDER, with the rigid
centerspan and unfolding wing tips, except that the stowed wing volume is also reoriented 90 degrees to
align with the fuselage due to the rotating center section.

Consideration of the overall volume ratio, Vs /Vd , with respect to the span ratio of rigid inner section length
to deployed length, b0 /b, is plotted in Fig. 25. In contrast with the large total stowed volume required
by hinged rigid wings on carrier decks, the total stowed volume of the FINDER aircraft is seen to include
a minimal amount of unused volume associated with the wing packing design. This low-profile concept
produces an overall stowed aircraft geometry that enables side-by-side packaging of multiple units while
minimizing unused volume.

Figure 26 shows comparisons of the volumes of different ILC inflatable wings in both packed and inflated
states. Three different types of inflatable wing construction methods are shown that include a range of wing
loadings, from low to high. Folding the wing results in lower packed volumes than rolling. As the wing
loading increases, the packed volume to inflated volume ratio increases as well due to the increase in required
material thickness to withstand the higher wing loading. The packed volumes range from less than 2% (50:1)
of the inflated volume for the lightest material to just over 10% (10:1) of the inflated volume for the heaviest.

17 of 23

American Institute of Aeronautics and Astronautics


Figure 23. Comparison of deployed volume envelope to stowed volume envelope for a single hinged fold of a
rigid wing with ∆b > b0 /2.

Figure 24. NRL FINDER with deployed volume envelope, with comparison of deployed volume envelope to
stowed volume envelope.

Overall compliance packaging (or flexible folding) has been more recently demonstrated for MAV-class sys-
tems, but for larger air vehicles, the concept of overall compliance and the significant packing advantage
that comes with this has been seen in numerous UAVs with inflatable wings.

Current inflatable-wing aircraft include missile-launched concepts as well as designs used to test the feasibility
of inflatable wings for Mars exploration (Fig. 27). Laboratory tests of current inflatable wing concepts realize
stowed-to-deployed volumes of 5-10% without vacuum packing and 2-5% with vacuum packing. Baffled
inflatable-wings can be rolled and segmented in addition to z-folding. Segmented folding results in a thin
compliant stowed volume that can be wrapped around the fuselage similar to the bendable-wing MAV
packaging.

18 of 23

American Institute of Aeronautics and Astronautics


Figure 25. Volume ratio versus span length ratio.

Figure 26. Comparison of packed versus inflated volumes for 3 different wings.

C. Deployment Time

Deployment time requirements are dependent upon the specific deployment scenario. There are cases where
deployment time may not be a concern; parachute or balloon deployment, for example. In-flight deployments
will probably need to be more rapid, unless the deployment is carefully controlled and symmetric. Regardless,
deployment time requirements may be play a role in the deployment system weight. For example, inflatable
wings can use pressure vessel reservoirs, compressors, or chemical systems, each with different deployment
times and various levels of cost, complexity and reliability.

Time to deploy the wings may also be of concern during the flight. Unless the vehicle is suspended beneath a
balloon or parachute, the wings will encounter aerodynamic loads during flight. These may be used to enable
in-flight deployment, such as was used successfully on FLYRT as shown in Fig. 28. Drag forces force rotation
of the inboard wing section while lift forces drive the outboard sections into their flight positions. Leveraging
the aerodynamic loads not only reduce weight and complexity, but provide the correct deployment time
scale, albeit at increased risk since deployment is dependent upon correct aircraft attitude. Inflatable wing
deployment itself occurs when the internal wing pressure is greater than the external pressure, typically using
an onboard storage tank but possibly chemical means. Deployment time is a function of many parameters,
including internal wing volume, inflation pressure, and the inflation mechanism. Since wing stiffness is
directly related to the inflation pressure, each wing is designed with an upper wing loading in mind. An
increase in wing loading requirement (through an increase in vehicle weight or dynamic pressure) requires
an increase in inflation pressure and possibly an increase in wing material thickness. Regardless, inflation

19 of 23

American Institute of Aeronautics and Astronautics


(a) Packed wings rolled (left) and folded (b) Deployed for flight.
(right).

Figure 27. UK/OSU/ILC BIG BLUE inflatable wing UAV.

time can be very rapid, such as shown in Fig. 29 for both ground and flight tests for two different designs,
where the entire inflation process takes less than 1/4 second using compressed air.34 Regardless, even with
slow inflation times at design flight speeds, it has been shown that the wings will deploy to the completely
deployed state under aerodynamic loads.36, 41 The deployment behavior has been shown to be invariant of
the folding method and flight conditions.

Figure 28. NRL’s FLYRT in flight.18

Deployment time requirements for in-flight deployment depend on vehicle response time which is related to
roll rate. The total rolling moment can be determined from
rb pb 1
Clβ + Clδr δr + Clδa δa + Clδr + Clp = Ixx ṗ
2U 2U qSb

where p is the roll rate in radians/sec, Ixx is the moment of inertia in roll, Clp is the roll damping coefficient,
Clδa is the control power coefficient, and δa is the effective aileron deflection in radians.42 Considering only
a single degree of freedom and neglecting roll due to rudder deflection and sideslip, this is simplified to
Ixx pb
ṗ − Clp − Clδa δa = 0
qSb 2U
For a given flight, flight parameters include q, S, b, U and Ixx while measurable variables include p and ṗ.

20 of 23

American Institute of Aeronautics and Astronautics


(a) Rapid round deployment of ILC FASM (<1/4 s).

(b) In flight deployment of NASA Dryden I2000.

Figure 29. Rapid deployment sequence of small scale inflatable wings (<1/4 s).

At initial lateral control input, we can write


Clδa δa qSb
ṗ =
Ixx
while at steady state we have
pb −Clδa δa
=
2U Clp
The dimensionless roll-rate requirement, pb/2U , is approximately 0.25. The roll response time for the rolling
mode is approximately given by
Sb2 Cl Cn
1/τR ≈ −ρV 0 Clp [1 − β p ]
4Ix Cnβ Clp
which is proportional to V /q.43 For larger UAVs, the response times are on the order of 1 s but smaller
vehicles can be 0.1 s or lower requiring rapid deployment or slow deployments that are highly symmetric and
controllable.36 For in-flight aspect ratio changing, stability analysis shows that it is feasible to use deployable
wings on UAVs because the longitudinal and lateral motions remain stable for values of aspect ratio ranging
from 3 to 10.44

IV. Summary

Deployable wing technology can offer many benefits to specific UAV missions, most importantly low-volume
storage of the wings when the vehicle is not in-flight. Deployable wing technologies enable aircraft to be
more effective by providing larger numbers in volume restricted applications, such as air, submarine, or hand
launches. As platforms have become smaller, new avenues for deployable wing design have appeared, many
biomimetic. Like morphing aircraft technology, there is not a single solution for aircraft requiring deployable
wings and only a portion of the solutions and their inherent limitations have been discussed herein. The
suitability of low-volume storage on an aircraft is dependent on weight, span, storage requirements, and the
vehicle flight envelope so each solution must be examined as part of the design process. Table 1 summarizes
some of the concepts discussed in this paper along with their advantages and disadvantages. TRL values
are not listed since they will vary greatly among specific designs. The continued development of miniature
smart actuators, low weight high strength materials and compliant mechanisms will have a direct impact on

21 of 23

American Institute of Aeronautics and Astronautics


deployable wing technology, allowing new designs to achieve greater spans, packing and deployment scenarios
in the years to come.

Table 1. Summary of deployable concepts

Type Limitations Advantages Disadvantages


Folding
Rigid Volume limited Large spans allowable Complex deployment
Simple construction Low packing ratio
Flexible Volume limited Light weight Simple deployment
High packing ratio Scalability
Flexible (wing warping) Low strength
Inflating Mass limited High packing ratio Leakage
Robust High loads require high pressure
Scalable No fuel storage
Flexible (wing warping) Weight
Nesting Volume limited Optimizes volume Complex deployment
Profile tailoring Limited design space
Limited fuel storage
Extending Volume limited Controllable actuation Friction leads to binding
Large span allowable Limited fuel storage
Limited number of span extensions
Weight

V. Acknowledgements

The authors would like to acknowledge many helpful discussions on the topic of folding wings with Rick
Foch of the Naval Research Laboratory Vehicle Research Systems branch and on inflatable structures with
Dave Cadogan, Steve Scarborough, and Dan Gleeson at ILC Dover, Inc.

References
1 Cadogan, D., Smith, T., R., Scarborough, S., and Graziosi, D. “Inflatable and Rigidizable Wing Components for Unmanned
Aerial Vehicles,” AIAA 20036630, 44th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Con-
ference, Norfolk, VA, April 2003.
2 Jacob, J., Smith, S., Cadogan, D., and Scarborough, S. “Inflatable Wings as a UAV Design Alternative,” AUVSI North

America, Washington, DC, August, 2007.


3 Jacob, J.D. and Smith, S. “Design of HALE Aircraft Using Inflatable Wings,” AIAA 2008-0167, 46th AIAA Aerospace Science

Meeting & Exhibit, Reno, NV, Jan.8–11, 2008.


4 Murray, J., Moes, T., Norlin, K., Bauer, J., Geenen, R., Moulton, B., and Hoang, S. “Piloted Simulation Study of a Balloon-

Assisted Deployment of an Aircraft at High Altitude,” NASA Technical Memorandum 104245, January, 1992.
5 Hall, D., Parks, R., and Morris, S. “Airplane for Mars Exploration: Conceptual Design of the Full-Scale Vehicle Design,

Construction and Test of Performance and Deployment Models,” Final Report, May, 1997.
6 Smith, S., Hahn, A., Johnson, W.,.Kinnery, D., Pollitt, J., and Reuther, J. “The Design of the Canyon Flyer, an Airplane for

Mars Exploration,” AIAA Paper 2000-0514, 38th AIAA Aerospace Sciences Meeting, Reno, NV, Jan. 2000.
7 Landis, G.A., A. Colozza, and C.M. LaMarre, “Atmospheric Flight on Venus,” AIAA Paper 2002-0819, 40th AIAA Aerospace

Sciences Meeting, Reno, NV, January 2002.


8 http://marsairplane.larc.nasa.gov/, December 28, 2007.
9 Chandler, J. and Jacob, J., “Design and Flight Testing of a Mars Aircraft Prototype Using Inflatable Wings,” 58th International

Astronautical Congress, Hyderabad, India, September 27, 2007.


10 Smith, S., Smith, W., Siegler, M. and Jacob, J. “Multi-Disciplinary Multi-University Design of a High- Altitude Inflatable-

Wing Aircraft with Systems Engineering for Aerospace Workforce Development,” 46th AIAA Aerospace Sciences Meeting &
Exhibit, Reno, NV, Jan. 7–10, 2008.

22 of 23

American Institute of Aeronautics and Astronautics


11 Weisshaar, T. “ Morphing Aircraft Technology – New Shapes for Aircraft Design,” Report RTO-MP-AVT-141, October, 2006.
12 Armando R. R. ”Morphing aircraft technology survey.” AIAA 2007-1258, 45th AIAA Aerospace Sciences Meeting and Exhibit,
Reno, NV, Jan. 2007.
13 Abdulrahim, M. and Lind, R. “Modeling and Control of Micro Air Vehicles with Biologically-Inspired Morphing,” Proceedings

of the 2006 American Control Conference, Minneapolis, Minnesota, USA, June 14-16, 2006.
14 NRL Vehicle Research Section, http://www.nrl.navy.mil/vrs.
15 Landon, S. “Development of Deployable Wings for Unmanned Aerial Vehicles Using Compliant Mechanisms,” MS Thesis,

Brigham Young University, August, 2007.


16 Skillin, M. and Crossley, W. “Modeling and Optimization for Morphing Wing Concept Generation,” NASA/CR-2007-214860,

March 2007.
17 Foch, R. “Deployable Aircraft Structures,” DARPA Rapid Eye Industry Day Presentation, July 25, 2007.
18 Bovais, C.r S., Davidson, P. T. “Flight testing the Flying Radar Target (FLYRT),” AIAA-1994-2144, AIAA Biennial Flight

Test Conference, 7th, Colorado Springs, CO, June 20-23, 1994, Washington, DC.
19 Stiles, P. Roadable Aircraft From Wheels to Wings: A Flying Auto & Roadable Aircraft Patent Search, Custom Creativity,

1994.
20 Radcliffe, T. O., Cesnik, C. E. S., “Aeroelastic Response of Multi-Segmented Hinged Wings,” AIAA-2001-1371

AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference and Exhibit, 42nd, Seattle, WA,
Apr. 16-19, 2001.
21 Radcliffe, T.. O. “Aeroelastic Study of a Multi-Hinged Wing,” PhD Dissertation, MIT, February, 2003.
22 Ifju, P. G., Jenkins, D. A., Ettinger, S., Lian, Y., Shyy, W., and Waszak, M. R., “Flexible-Wing-Based Micro Air Vehicles,”

AIAA 2002-0705, AIAA Aerospace Science Meeting & Exhibit, Reno, NV, January 2002.
23 Johnson, B., Claxton, D., Stanford, B., Jagdale, V. and Ifju, P. “Development of a Composite Bendable-Wing Micro Air

Vehicle,” AIAA-2007-1044, 45th AIAA Aerospace Sciences Meeting and Exhibit, Reno Nevada, 8-11 January, 2007.
24 Kahn, A.D. and Foch, R.L. “High Packing Efficiency EAV for Local Area Seeding,” www.nrl.navy.mil/vrs/pdfs/06-1226-

1085.pdf.
25 Berthe, C. J., Wyckoff, P. B. and Baals, J. R. “Longitudinal Stability of the Goodyear Inflatoplane,” Princeton University

Department of Aerospace and Mechanical Sciences, Report No. 689, May 1964.
26 Baals, J. R., Personal communication, July 2008.
27 Comer, R. and Levy, S. “Deflections of an Inflated Circular Cylindrical Cantilever Beam,” AIAA Journal, 1, No. 7, 1963, pp.

1652–1655.
28 Main, J., Strauss, A. and Peterson, S. “Load-Deflection Behavior of Space-Based Inflatable Fabric Beams,” Journal of

Aerospace Engineering, 7, No. 2, April 1994, pp. 225–238.


29 Main, J., Peterson, S., and Strauss, A. “Beam-Type Bending of Space-Based Inflated Membrane Structures,” Journal of

Aerospace Engineering, 8, No. 2, April 1995, pp. 120–125.


30 Brown, G., Haggard, R., and Norton, B. “Inflatable Structures for Deployable Wings,” AIAA 2001-2068, AIAA Aerodynamic

Decelerator Systems Technology Conference and Seminar, Boston, MA, May 2001.
31 Breuer, J., Ockels, W., and Luchsinger, R. “An Inflatable Wing Using the Principle of Tensairity,” AIAA-2007-2117, 48th

AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, April, 2007.


32 Priddy, United States Patent Office, Patent Number 1,905,298, 1933.
33 Veldman, S. L., “Design and Analysis Methodologies for Inflated Beams,” Ph.D. Dissertation, Delft, 2005.
34 Murray, J., Pahle, J., Thornton, S., Frackowiak, T., Mello, J., and Norton, B. “Ground and Flight Evaluation of a Small

Scale Inflatable-Winged Aircraft,” AIAA 2002-0820, 40th AIAA Aerospace Sciences Meeting & Exhibit, Reno, NV, Jan. 14–17,
2005.
35 Simpson, A. “Design and Performance of UAVs with Inflatable Wings,” Ph.D. Dissertation, University of Kentucky, January,

2008.
36 Loh, W. K, and Jacob, J.D. “Behavior and Control of Dynamic Deployment of Inflatable Wings,” AIAA 2008-1291, 47th

AIAA Aerospace Science Meeting & Exhibit, Orlando, FL, Jan. 5–8, 2009.
37 Yanof, J. and Crowley, D. “Glide and Wind Tunnel Testing and Mathematical Analysis of Wing Configurations For Morphing

Wing Aircraft,” AIAA Regional Student Paper Competition, Paducah, KY, 2003.
38 Bondeau J., Richeson J., and Pines D.J. “Design, Development and Testing of a Morphing Aspect Ration with using an

Inflatable Telescope Spar.” AIAA 2003-1718, 44th AIAA/ASME/ASCE/AHS Structure, Structural Dynamics, and Materials
Conferences, Norfolk, VA, April. 2003
39 Loh, W. K, and Jacob, J.D. “In Flight Aspect Ratio Morphing Using Inflatable Wings,” AIAA 2008-0425, 46th AIAA

Aerospace Science Meeting & Exhibit, Reno, NV, Jan. 8–11, 2008.
40 Inman, D. J., Good, M. G., Johnston, C. O., Robertshaw, H. H. Mason., W. H. and Neal, D. A. “Design and Wind-Tunnel

Analysis of a Fully Adaptive Aircraft Configuration,” AIAA 2004-1727, AIAA Aerospace Sciences Meeting.
41 Loh, W. K. “Deployment Dynamics of Inflatable Wing,” MS Thesis, Mechanical & Aerospace Engineering, Oklahoma State

University, Dec. 5, 2008.


42 Kimberlin, R., Flight Testing of Fixed-Wing Aircraft, AIAA, Reston, 2003.
43 Stevens, B. and Lewis, F. Aircraft Control and Simulation, Wiley-IEEE, 2003.
44 Abdullah, E. J., Ajir, M. R., Ahmad, M. T., and Filipski, M. N. “Effect of Aspect Ratios on Longitudianal and Lateral

Motions of Unmanned Aerial Vehicles,” International Journal of Science and Technology, Vol. 3, 2006.

Paper v. 1.0, January 1, 2009.

23 of 23

American Institute of Aeronautics and Astronautics

You might also like