You are on page 1of 19

CHAPTER

THE DEVELOPMENT
OF MORPHING AIRCRAFT
BENEFIT ASSESSMENT
3
Fabian Peter, Eike Stumpf
RWTH Aachen University, Aachen, Germany

CHAPTER OUTLINE
1 Experiments as Basis for Morphing Progress .................................................................................... 103
2 The Advent of Transonic Methods .................................................................................................... 105
3 Automated Methods as Enabler for Large Scale Studies .................................................................... 113
4 Reintroduction of Flexible Materials ................................................................................................ 114
5 The Final Step to Industrial Application ........................................................................................... 118
References ........................................................................................................................................ 119

NOMENCLATURE
ACTE adaptive compliant trailing edge
CD drag coefficient
CL lift coefficient
C∗p critical pressure coefficient
FEM finite element method
L/D lift to drag ratio
Ma Mach number
MAW mission adaptive wing
NASA National Aeronautics and Space Administration
RBM root bending moment
Re Reynolds number
VCCTEF variable camber continuous trailing edge flap

1 EXPERIMENTS AS BASIS FOR MORPHING PROGRESS


To understand why performance assessment of morphing aircraft is crucial, it is important to reflect on
the history of morphing. In the early days of aviation, the examples in nature, especially the flight of
birds, were the focus of aircraft designers. The flight of insects, birds, and bats as we know it would not
Morphing Wing Technologies. https://doi.org/10.1016/B978-0-08-100964-2.00003-4
# 2018 Elsevier Ltd. All rights reserved.
103
104 CHAPTER 3 THE DEVELOPMENT OF MORPHING WING RESEARCH
AND TECHNOLOGY

be possible without morphing. The wing planform and camber of birds change dramatically and con-
tinuously with every change in flight conditions [1]. In addition, the structures and materials used in
early aviation (e.g., wood and canvas) were somewhat flexible and allowed for a relatively easy de-
formation. In the 1890s, experiments were conducted to achieve roll control of gliders with wing warp-
ing by Montgomery. The best-known example, which is cited in most publications about morphing as
an early application, is the Wright Flyer and its first powered flight in 1903. Here the Wright brothers
realized roll control by asymmetrical deformation of the outboard wing sections [2]. In a report in 1920,
Parker was one of the first to address the challenge that increasing aircraft cruise speeds posed toward
the requirement of an acceptable landing speeds. He cited three possible approaches for this problem:
the variation of incidence angle of the wing relative to the fuselage, increasing the surface of the wing,
and variable camber. He chose a mechanical concept that he described as variable camber. The intent
would be to deform a surface (i.e., the upper surface of a biplane) in a way that offers little drag and lift
at high speed but high lift with the necessary drag at low speeds. He envisioned a design by which the
surface is beneficially deformed by aerodynamic loads of the different flight attitudes integrating flex-
ible elements into the wing structure (see Fig. 1). Such a design would require no external control
mechanisms. The design was tested but did not find its way into a commercially successful airplane [3].
Another example of early geometry changes was the telescopic wing presented by Makhonine in
1931. The concept was integrated into an airplane called Makhonine Mak-10, which conducted test
flights in 1932 [4].
Since then, morphing has constantly been part of aviation research, but morphing by deformation
shifted out of focus with the appearance of rigid structures and materials, which do not respond well to
large deformations. However, morphing by the deflection or shifting of rigid elements stayed an es-
sential part of aircraft design. Even the deflection of an aileron or the deployment of a movable landing
gear could be interpreted as morphing, making it a part of current civil aviation. However, the general
interpretation of the term morphing is a large deformation of parts of the aircraft or continuous deploy-
ment of control surfaces. The need to change the geometry of the wing increases with the difference in
flight conditions in which an aircraft has to operate.
The most fundamental change is the difference between a high cruise speed to a low landing speed.
For this reason, with the increase in aircraft’s cruise speed, high lift devices were introduced.

FIG. 1
Parker variable camber wing [3].
2 THE ADVENT OF TRANSONIC METHODS 105

FIG. 2
Bell X-5 variable sweep test aircraft [6].

Subsequently, the swept wing enabled speeds approaching and exceeding the sound barrier. The dif-
ficult low-speed handling qualities of highly swept wings of military supersonic aircraft called for a
change of sweep angle during flight.
The first example for this technology was the Bell X-5 in 1951. Based on a captured Messerschmitt
P.1101, whose wing sweep could be altered on the ground, the Bell X-5 could vary its wing sweep angle
from 20 to 60 degrees (see Fig. 2). Because the center of pressure moved with variation of the sweep,
the wings were mounted on rails in order to move forward, resulting in a complex mechanism. The X-5
had a successful test career and provided important data, building the basis for the introduction of var-
iable sweep in military aircraft [5].
Variable sweep wings found application in several western and eastern military production designs,
one of the first being the General Dynamics F-111 Aardvark. With the feasibility of high speeds and the
emerging possibilities of transonic wind tunnels and theoretical methods for transonic and supersonic
flow calculations, more research was focused on the economics of airfoils at different speeds. This led
to large-scale research projects.

2 THE ADVENT OF TRANSONIC METHODS


The F-111 served as a testbed for the National Aeronautics and Space Administration (NASA) research
programs for almost 20 years. The first program started in the 1970s and was named Transonic Aircraft
Technology (TACT). It focused on the integration of a supercritical wing airfoil and the validation of
calculation methods and wind tunnel tests [7]. In the 1980s, the Advanced Fighter Technology Inte-
gration (AFTI) program was introduced. Its aim was to increase the efficiency in variable mission and
106 CHAPTER 3 THE DEVELOPMENT OF MORPHING WING RESEARCH
AND TECHNOLOGY

FIG. 3
General Dynamics F-111 Aardvark with the mission adaptive wing [10].

flight conditions by integration of the mission adaptive wing (MAW) as shown in Fig. 3. For this pur-
pose, the wing of the F-111 was equipped with a smooth variable camber leading and trailing edge [8].
The forward 20% of the wing chord could be uniformly deflected 5 degrees up and 30 degrees down.
The aft 40% of chord could be deflected from 7.5 degrees to 25 degrees with a restriction of 1 degrees/
foot of span. [9].
The goal of the variable camber application was to achieve lift to drag ratio (L/D) improvement,
integral roll control, maneuver load control, direct lift mode (increased lift without increased angle of
attack), and gust alleviation. Flight test results comparing the TACT with the AFTI are shown in Fig. 4.
For a lift coefficient (CL) of 0.5, as indicated by the dashed line, a reduction of trimmed drag coefficient
(CD) of more than 8% compared to the TACT drag was achieved. The increase in weight for the in-
tegration of the MAW type leading and trailing edge was stated to be 94 lb. (42.6 kg). The F-111’s
original wing, which was equipped with a Krueger flap and a single slotted Fowler flap, weighed
9068 lb. (4113.2 kg), showing an increase in weight of about 1%. [11].
Variable sweep wings are still in operation in current military aircraft, such as the Panavia Tornado
from 1974. With complex and heavy hinge mechanisms, variable sweep wings never found their way

1.20
AFTI/F-111
1.00
Trimmed lift coefficient

TACT/F-111
0.80

0.60

0.40

0.20

0.00
0.00 0.05 0.10 0.15
Trimmed drag coefficient
FIG. 4
Lift to drag improvement for mission adaptive wing [11].
2 THE ADVENT OF TRANSONIC METHODS 107

into civil aviation and were replaced by delta wings in most military applications. However, they were
essential for enabling high speeds. In civil aviation, the methods developed for transonic speed regimes
showed the potential of fixed camber change for higher Mach numbers. These methods were applied
during the development of the Airbus A310, and then were used to identify potential improvements for
the Airbus A300-B4. The A300-B4’s wing was developed with theoretical subsonic tools, extensive
wind tunnel testing, and experiences made during the development of the Trident and HS 125.
A certain amount of rear loading was already applied in the A300-B4’s original wing, however, the
methods available in the 1980s allowed for further optimization of the airfoil. It was found that an in-
crease in camber of the trailing edge moved the shock aft and increased the area of supersonic flow.
Consequently, the camber of the A300-600 was increased. The increase was stronger inboard and smal-
ler outboard. The result on the drag can be described by a rotation of the polar in a CL versus CD di-
agram similar to Fig. 4. The increase in camber results in a rotation of the polar around a specific point
situated on the polar. In order for the change to be beneficial, the point has to be at lower CL than the
intended cruise CL, thereby shifting the polar points of higher CL values to smaller CD values. The
position of the rotation point is dependent on the Reynolds number (Re) and moves to higher CL values
with increasing Re. This was confirmed by flight tests with a modified A300-B4. In total, the aerody-
namic improvements, which also included other changes such as the elimination of boundary layer
fences and measurements for parasite drag decrease, amounted to an L/D improvement of 6% [12].
Furthermore, the stronger camber increase inboard changed lift distribution and reduced wing root
bending moment (RBM) by 2.4% [13].
The advances in 2D airfoil analysis were also beneficial for the improvements for helicopter blades.
Furthermore, the application of camber variation is very attractive for helicopters. This is explained in
more detail in a publication of Awani from 1982 [14]. He describes the inherent problem of very dif-
ferent flow conditions for advancing and retreating blade and the resulting necessary compromise for a
fixed airfoil selection (see Fig. 5). For the advancing blade, compressibility effects dominate the design
choice; for the retreating blade, blade stall is the aspect with the strongest impact on the airfoil choice.
The study by Awani presents an integration of a simple flap geometry in the Vertol VR-7 rotor
blade. The flap extents from 50% chord to the trailing edge of the blade and is able to deflect 20 degrees
downward. The chord length is a compromise between small deflections and large flap chords with the
goal of achieving the desired camber change (Fig. 6).
The analysis was conducted with computational methods and yielded a reduction of necessary
power in hovering flight of 2.6%. Helicopter blades are still an area of research for morphing, incor-
porating piezo elements for actuation and with the long-term goal of reducing the complex and main-
tenance intensive rotor kinematics and decreasing vibrations and noise (e.g., [16]). In the following, the
scope is set on fixed-wing aircraft.
Another NASA program including variable camber was the X-29 (see Fig. 7), which had its first
flight in 1984 and the second in 1989. While the most obvious design features were the planes forward-
swept main wings and its canards, it also featured discrete variable camber flaperons (see Fig. 8). In
comparison to the AFTI smooth variable camber, the discrete variable camber was introduced to the
X-29 because of its lower cost. While offering a smooth upper surface at design conditions, its penalties
at off-design conditions are described to be minor. In addition to the performance optimization over-
large CL regimes, an additional goal for the variable camber application in the X-29 was the usage for
roll control [17]. Because of the many highly innovative design features of the X-29, it was difficult to
allocate performance improvements to specific technologies [18].
108 CHAPTER 3 THE DEVELOPMENT OF MORPHING WING RESEARCH
AND TECHNOLOGY

1.6 Region III


Forward
flight

1.2

Retreating
blade stall
Region II
Lift coefficient, CL

0.8 Hover

0.4 Region I
Forward
flight

0
0.2 0.4 0.6 0.8 1.0
Mach number, Ma

Compressibility
–0.4 effects on
advancing blade
FIG. 5
Influence of lift coefficient and Mach number on rotor aerodynamics [15].

Blade forward C
section 0.9R .50C df Radius = .1060C
line
e
ing Undeflected flap
ph
Fla Flap to control
rotor lift (d f = 20°)
Chord Hinge d f = 20°
Pilot input to flap deflection line axis Deflected
(standard control motions flap
with swashplate mechanisms)
FIG. 6
Helicopter blade variable camber concept [14].

The knowledge gained by research for transonic airfoils and NASA research aircraft featuring var-
iable airfoils showed the potential of variable camber. Consequently, the focus of research in the field
of morphing for civil aviation shifted to the topic of variable camber. Fundamental experiments were
conducted to improve the understanding of the camber influence and the design of specific variable
camber airfoils. The basic phenomenon is that, for constant lift, the minimum drag is situated at higher
CL with increasing camber. This can be explained by the increase in rear loading, which reduces local
velocities and hence skin friction at the leading edge. Therefore, the relatively sharp optimum of a fixed
2 THE ADVENT OF TRANSONIC METHODS 109

FIG. 7
Grumman X-29 [17].

Full up
Forward
position

Cruise
position

Maneuver
position

Full down
position
FIG. 8
The X-29’s discrete variable camber flaperon [17].

wing L/D versus CL graph can be changed to a rather flat optimum for a variable camber wing. If the
improvements are located at higher CL, which is the case in most studies, the buffet onset is also
delayed to higher CL. The increase in camber is sensible until trailing edge separation occurs. Also
taking into account the relaxation of the design compromise of a fixed wing and restraints for maneuver
loads, Szodruch states L/D improvements of 3%–9%, a buffet boundary increase of 12% and a block
fuel reduction of 5%. The studies also showed that a continuous camber change is not necessary be-
cause most of the improvements are already reached when the camber setting is changed two times
during a short-range mission [19].
Greff published an extensive study about the opportunities for variable camber in civil aircraft fam-
ilies in 1990 [20]. The primary motivation for this publication was the application of variable camber
on a wing that can be used for medium- and long-range aircraft. The wing of an aircraft family has to be
110 CHAPTER 3 THE DEVELOPMENT OF MORPHING WING RESEARCH
AND TECHNOLOGY

large enough to meet the requirements of the largest family member, which is usually the most
stretched version. The specific design tasks in the focus of the Airbus engineers in that period was
the common wing for the Airbus A330 and the Airbus A340. For smaller planes of the family, this
wing means a design compromise. This compromise is accepted since the wing contributes more than
50% of the development cost, 45% of structural weight, and 60% of the drag. Hence, a common wing as
basis for an aircraft family offers large cost savings. The concept of the MAW is too expensive and
complex for civil transport aircraft, which spend most of their cruise flight in patterns of relatively
similar speed and altitudes. Furthermore, the need for lift enhancement during take-off and landing
is larger for civil transport aircraft than for fighter planes, thereby an application of the MAW concept
would conflict with high-lift devices needed for transport aircraft. For this reason, variable camber re-
search focused on the use or modification of conventional trailing edge components in order to min-
imize development effort. The concept introduced by Hilbig [21] and analyzed by Greff makes use of
an additional rotation track that enables the Fowler flap to rotate and translate for a limited amount
during which no gap occurs between flap and spoiler (see Fig. 9) [20].

FIG. 9
Greff and Hilbig variable camber concept [20].
2 THE ADVENT OF TRANSONIC METHODS 111

VC-design
Dragrise in
Conventional Max. Reduced super- off-design
transonic design mach-no. sonic region CP
−1.2

Stable shock position CP*


−1.0
X/C
Low (dcp/dx) ‫٭‬
−0.8
Other
CP* wise
−0.6

−0.4 Moderate
CP

Subsonic
−0.2 recompression
(separation)
0

0.2

0.4 T.E. thickness


Reduced degree structural weight
of rear loading behind rear spar
0.6 trim drag
0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.7 1.0
X/C X/C
FIG. 10
Greff variable camber airfoil design [20].

Greff states that special attention has been paid to the airfoil design for variable camber application.
Highly loaded airfoils with large supersonic flow regions are prone to trailing edge separation, espe-
cially if the hinge position of the varied surface is close to the shock position. Also, the regions around
the critical pressure coefficient (Cp∗) should exhibit small gradients to guarantee a stable shock for off-
design conditions. The rear loading of the lower surface should be decreased to counter the increase in
pitching moment. These goals are illustrated in Fig. 10.
A crossover point (where a benefit of the variable camber polars exists) below the design CL would
be achieved only if these aspects are taken into account. For this reason, an airfoil was designed spe-
cifically for the VC application. The basis for comparison was a highly loaded airfoil designed for a
local CL of 0.6 and a local Mach number of 0.74. The VC airfoils design point was chosen at a lower
local CL of 0.48 in order to enable the performance increase by deflection for higher CL. The VC airfoil
already had a 4% lower minimum drag compared to the basic airfoil. Without varying the flap deflec-
tion, the VC airfoil showed higher drag for higher CL than the basic airfoil. This clarifies that the VC
specific airfoil has to be operated with deflection for its optimum performance. With flap deflections,
the variable camber airfoil showed superior drag characteristic. Taking into account the increased trim
drag, because of the increase in rear loading by deflection, an L/D improvement of 4% was determined
112 CHAPTER 3 THE DEVELOPMENT OF MORPHING WING RESEARCH
AND TECHNOLOGY

for the family common wing and for long-range aircraft. As control, a simple table lookup function was
proposed, providing the flap deflection at a specific speed, weight, and atmospheric data. Further ben-
efits discussed but not taken into the final assessment were the possible reduction of RBM with a pos-
sible manufacturing weight empty reduction of 1.5%–2%, drag decrease by a  1 degrees reduced wing
root setting affecting fuselage drag, and the possible reduction of horizontal and vertical stabilizer area,
which is connected to a decrease in wing area. Subsequently, a variable camber benefit of 3%–6%
block fuel reduction is stated as expectation for the use in a medium- and long-range family
application [20].
Greff did state in his 1990s publication that although a spanwise differential variable camber might
be advantageous with respect to load alleviation and offer opportunities to adjust to impacts of engine
integration influences, the increase in complexity did not justify an application at that time.
Spillmann concentrated on the application of a spanwise differential camber in a publication in
1992. He noted that birds are able to cope with their very different flight attitudes not only by varying
the angle of incidence but also by varying camber across the span. Most of the mission of a civil trans-
port aircraft takes place in level flight, which determines the performance with respect to aerodynamic
efficiency. However, the structure of the wing is determined by very brief periods of maneuvers and
gusts. In these situations, the aerodynamic optimum lift and load distribution along the span is unfa-
vorable from a structural point of view. If load distribution could be altered, e.g., load shifted toward the
wing root, the structural weight of the wing could be reduced. Studies at Cranfield University incor-
porating this approach resulted in a 10% wing weight reduction for a 150-passenger aircraft, including
weight increases of 25% for the flap structures and 10% for the flight control system [22].
Rossow presents a quite dramatic geometry change in his publication from 1991 [23]. He proposes a
nose flap and a trailing edge flap that deflect from the upper surface from the airfoil in order to establish
and trap a vortex (see Fig. 11). The resulting flow field is stated to achieve CL values up to 6.

Kutta
Equilibrium point
condition satisfied yAF/c
at tip of flap for vortex and sink
1

Rear
Nose flap stagnation
Airfoil point of
separation
U∞
bubble
xAF/c
−2 −1 0 1 2

Forward stagnation Kutta


point on airfoil condition satisfied
at trailing edge
of airfoil
FIG. 11
Lift increase by establishing a trapped vortex [24].
3 AUTOMATED METHODS AS ENABLER FOR LARGE SCALE STUDIES 113

Studies were conducted on a 2D model in a water tunnel and showed that a vortex could be estab-
lished when fluid was removed through a sink at the center of the vortex. A former publication dis-
cussed the use of only one single flap at the front of the airfoil [24]. The aft flap was found to be
beneficial to limit the extent of the vortex. Furthermore, with the aft flap, the sink was necessary only
to establish the vortex.

3 AUTOMATED METHODS AS ENABLER FOR LARGE SCALE STUDIES


The publications listed thus far were based on methods used for one specific design and presumably a
large share of nonautomated work. A very large effort would be required for a study of different ge-
ometries. In addition, the studies concentrated on airfoil changes by mostly conventional control sur-
faces such as Fowler flaps. The availability of inexpensive computation power coupled with several
freely available low- and higher-level aerodynamic computational methods led to an increase in aero-
dynamic research in many fields at the beginning of the 1990s. For morphing, two-dimensional var-
iable camber studies included automatic airfoil geometry optimization. In a study published in 1997,
Martins presented a 3D vortex lattice analysis that is coupled with a multivariable function optimizer
algorithm. The basis was a swept tapered wing and its airfoil was deformed from the leading edge to the
20% chord and from the 70% chord to the trailing edge. The deformation was restricted by dividing the
wing surface with a linear finite element mesh and limiting the deformation of these elements. With
constraints considering high-lift performance, an induced drag reduction of 1.8%–2.3% was achieved
[25]. Martins did further research, and a publication in 1998 focused on 2D airfoil sections, which were
allowed to deform by translation and rotation of a leading and trailing edge segment, and a fixed wing
box composed of an unchanged chord section of the airfoil (27.6%–64.5% chord). The kinematics en-
abling these deformations were not part of the research. The aerodynamic analysis was conducted with
a combination of tools, including methods incorporating integral boundary layer, transition, and sep-
aration. Compressibility effects were not included, and therefore all studies were conducted at Mach
numbers (Ma) below the critical Ma. The flight conditions for the analysis were set to 25,000 ft and Ma
0.575. The evaluated target parameter is the range factor, which is the result of the integrated range
equation. For these boundary conditions, an automatic optimization study was conducted. For the case
of a sole deformation of the trailing edge, an increase of about 7.03% was realized, and for simulta-
neous deformation of the leading edge, 24.5% could be achieved. The calculated optimum leading edge
contour showed a very strong downward curvature. It was stated that this geometry most likely would
be not sensible if wave drag were analyzed. The importance of transonic effects on variable camber was
extensively covered in the previously mentioned publications. On its own, the vortex lattice method is
not able to incorporate compressibility or viscous effects. This has to be kept in mind for results for
transonic flight regimes, which are produced with a pure vortex lattice application [26,27].
The application of 2D and 3D Navier Stokes calculations in an extensive research project allowed
the assessment of an innovative trailing edge device concept by Airbus and the German Aerospace
Center. In this study, an extended outboard flap replaced the aileron. Additionally, the trailing edges
of the flap featured small simple hinge flaps with 10% wing chord or, alternatively, 2% chord split
flaps. These devices could be deployed differentially along the span. The stated results of the project
are a 7% increase in L/D and a 7% increase in maximum lift. This was achieved by improving the lift
distribution and reducing the wave drag for higher Ma with a decambering of wing sections.
114 CHAPTER 3 THE DEVELOPMENT OF MORPHING WING RESEARCH
AND TECHNOLOGY

Furthermore, the wing structure weight was estimated to be reduced by 2% through the application of
the trailing edge devices for active load alleviation [28].

4 REINTRODUCTION OF FLEXIBLE MATERIALS


Because of the progress in research of smart materials (e.g., advanced composites and shape memory
alloys), flexible wing parts are maturing and becoming more attractive for application in the aviation
industry, especially outside the wing box. These structures, also called compliant structures, are often
cited to achieve lower weight and lower maintenance cost [29]. Monner published the concept of the
adaptive wing project in 1998 [30]. The concept integrates a bump for weakening the shock and a flex-
ible variable camber trailing edge. Weakening the shock reduces wave drag, which is relevant for tran-
sonic flight regimes. Because of the curvature at the shock position, the shock is divided into several
weaker shocks. One approach is to influence the boundary layer in order to create an artificial bump.
This can be realized by a passive system, consisting of holes connected by a ventilation chamber. An
active solution with controlled flow through the holes is also conceivable. Monner, however, states that
in both cases this approach is prone to congestion of the perforation and therefore is not practical from a
technical point of view. The bump for shock control also can be realized by physically deforming the
airfoil surface as with the adaptive wing concept. The flexible Fowler flap is realized by connecting
chord-wise separated flap segments with revolute and prismatic joints, thereby allowing a smooth de-
formation with one actuator per span-wise distributed kinematic station. A demonstrator with a span of
3.2 m and a chord of 0.9 m was built and proved the functionality of the bump and the flexible flap.
A more detailed description of the flexible Fowler flap is given in Ref. [31].
Different approaches for actuation of a variable camber trailing edge also can be found [32]. Most of
the concepts feature finger-like elements with multiple hinges building a framework (e.g., the X-29’s
discrete variable camber flaperon). One concept, which differs considerably, is the horn concept intro-
duced by Mueller [33]. Here the actuation is realized by rotating rigid elements inside the flap, with a
horn or tusk-like shape. When the curved end of these elements is rotated downward, the flexible skin
of the flap is deflected. The main advantage is that the bearings carry the largest part of the aerodynamic
forces and only a small portion has to be countered when the rigid elements are rotated. The actuation
was successfully integrated in a 2.4 m span demonstrator with the outer shape of the outboard flap of an
Airbus A340 and three rotational elements. The kinematics were completely integrated into the flap
body, which offered the same stiffness as the original flap. The additional weight of the demonstrator
was less than 1% compared to the span portion of the original A340 flap. The demonstrator was
successfully used to emulate the usage as a roll-control device with deformation cycles equivalent
to a flight time of 60,000 h and loads imitating the aerodynamic forces.
Another concept for actuation involves the specific control of the stiffness of structural components
of the airfoil. Thereby actuation can either be reduced to very small loads or completely replaced by use
of aerodynamic forces. Although the basic idea already was presented in the 1890s by Parker, modern
materials can change stiffness properties, e.g., by applying electrical voltage, thereby opening new
prospects. An example for this is a study published by Raither in 2015 about an airfoil with variable
stiffness skin realized by electrobonded laminate [34].
The introduction of variable camber in small existing aircraft is a sensible step before designing a
new wing, although the full potential will not be used. Pfeiffer et al. published a report in 2005 testing
4 REINTRODUCTION OF FLEXIBLE MATERIALS 115

FIG. 12
Beechcraft Premier I Model at the UWAL [35].

different active trailing edge devices on a wind tunnel model of the Beechcraft Premier I [35]. The
Beechcraft Premier I is a light business jet that had its first flight in 1998 with a maximum cruise
Ma of 0.5. An existing wind tunnel model (see Fig. 12) was tested at the University of Washington
Aeronautical Laboratory (UWAL).
The four concepts tested were trailing edge wedges, distributed lower surface wedges, a slotted flap,
and a cruise tab. Different trailing edge geometries were attached to the wing for each concept. The
wind tunnel tests were accompanied by transonic CFD-calculations. The cruise tab was the most ben-
eficial wing modification. The results showed that the optimum deflection of the tab is strongly depen-
dent on Ma. A fuel burn reduction of 1.5% was estimated (see Fig. 13). Furthermore, the cruise speed
could be increased by 2–4 knots, leading to additional operational benefits.
The kinematic actuation also was investigated. An electrical actuation was judged to be beneficial,
because of the existing electrical flap actuation of the Beechcraft Premier I and because of easier re-
alization of locking the position after reaching a target deployment. The integration study into the Pre-
mier I is a conservative example because of the small space available in the flap. Further challenges
included the tapered outboard flap and the connections of the tab actuation in the flap panel.
A kinematic actuation with simple pushrods for the inboard flap and bell cranks on the tapered outboard
flap was devised (see Fig. 14). A single interconnected rod could drive the actuators in each flap.
To increase of the credibility of expected benefits, flight tests are an absolute necessity as the un-
conventional technology of morphing matures. In 2006, the FlexSys Company began flight tests of a
morphing airfoil carried by White Knight Aircraft [36]. The goal was to demonstrate the capability of
the L/D increase from the improved camber setting as well as increase of laminar flow region for higher
CL. For this reason, a natural laminar flow airfoil was selected, one that can reach laminar flow up to
65% chord on the upper airfoil side and 90% chord on the lower. Since the application being considered
was on a high-altitude, long-endurance reconnaissance aircraft, Ma were relatively low, with a
116 CHAPTER 3 THE DEVELOPMENT OF MORPHING WING RESEARCH
AND TECHNOLOGY

3
% Reduction in fuel burn (baseline - cruise tab)

Notes:
(1) Comparison is between baseline and cruise tab
configurations
(2) All missions performed using max takeoff weight
(3) All missions direct climb to 41,000 feet.
(4) ISA conditions throughout mission
2 (5) NBAA IFR 100 nm alternate reserves

Note: The cruise tab configuration performs at Long range cruise mission
2–4 knots faster than the baseline mission in High speed cruise mission
both HSC and LRC missions.

0
0 200 400 600 800 1000 1200 1400
Mission length - nautical miles
FIG. 13
Beechcraft Premier I cruise tab modification results [35].

FIG. 14
Beechcraft Premier I cruise tab actuation concept [35].
4 REINTRODUCTION OF FLEXIBLE MATERIALS 117

FIG. 15
FlexSys adaptive compliant trailing edge on NASA Gulfstream GIII Aircraft [37].

maximum of 0.55. The variable camber function was realized by a deformable trailing edge extending
from 70% to 100% chord, driven by two electrical servomotors inside the flap. The internal structure
deforms as a whole, and differential flap setting and actuation speeds of 30 degrees/s are possible. This
would enable later use for load alleviation for RBM reduction and gust load alleviation. No more details
of the kinematics causing the deformation were given. The tested wing section had a span of 50 in. and
a constant 30-in. chord. A wake probe collected the airfoil momentum deficit and hot film sensors de-
termined boundary layer transition location. The results showed that the region of laminar flow could
be increased by 10% compared to a fixed trailing edge. FlexSys continued with the morphing trailing
concept. In cooperation with NASA, a flight test campaign was planned in order to validate structural
effectiveness and airworthiness of an adaptive compliant trailing edge (ACTE). The ACTE was
designed to replace the conventional Fowler flap on NASA’s subsonic research aircraft testbed, which
is based on a Gulfstream G-III (see Fig. 15).
The ACTE consisted of five main elements: the main flap section, a connection spar, the actuation
system, and two transition sections connecting the inboard and outboard edges of the flap section to the
main wing without a gap. The dimensions of ACTE are given as a 19 ft span and 2 ft chord. After
the creation of finite element methods (FEMs) of the ACTE, ground vibration tests were conducted.
For the aeroelastic analysis, the joined ACTE and aircraft FEM were coupled with a linear panel
method. Flight tests were undertaken from Nov. 2014 to Apr. 2015. The ACTE were not actuated
in flight, but on the ground between flights. In order to excite the structure in the relevant frequencies,
multiple maneuvers were conducted including roll, pitch, yaw. The collected data showed that the
ACTE was airworthy from a structural point of view [37].
The effort of developing morphing trailing edge devices is mainly motivated by the long-term goal
of integration in large civil aircraft. Apart from practical experiments, which approach the topic with
increasing size of flight tests, the theoretical research focuses on the optimization of concepts and the
refinement of the applied methods. One example is NASAs variable camber continuous trailing edge
flap (VCCTEF) concept [38] (Fig. 16).
118 CHAPTER 3 THE DEVELOPMENT OF MORPHING WING RESEARCH
AND TECHNOLOGY

FIG. 16
Variable camber continuous trailing edge flap [39].

Early predictions for the VCCTEF state benefits up to 9.8% for an aircraft comparable in size to a
Boeing B757 and aerodynamic analysis with vortex lattice methods [40]. Another example is the re-
search of the University of Milano. Gaspari published several papers presenting the combination of
airfoil design optimization methods, including the internal compliant structures to achieve the desired
deformation, and aeroelastic analysis [41–43].

5 THE FINAL STEP TO INDUSTRIAL APPLICATION


More research activities are bringing morphing to an industrial level. Integrating a deforming skin to an
aircraft component, while maintaining the necessary functions (e.g., ice and lightning strike protection)
has been part of the EU research program SARISTU (Smart Intelligent Aircraft Structures) that con-
cluded in 2015 (leading edge [44], trailing edge [45], and morphing winglet [46]). This integration was
successfully tested in a wind tunnel. The benefit of the combination of these technologies has realized a
block fuel reduction of more than 10% [47]. Research also is focusing on operational aspects for system
failures and mission planning [48, 49].
Almost no morphing features have been implemented into current large civil aircraft and conven-
tional control surfaces have been neglected, likely because aircraft manufacturers believe predicted
potential is too little, they have low confidence in the quality of the predictions, or they believe the
technologies are not adequately mature. Aircraft that already have limited capability for morphing ap-
plications likely will make use of such functions as deforming skin. During the development of the
Boeing 787, several trailing edge devices were considered, including a variable camber concept pat-
ented by Sakurai in 2007 [50]. Subsequently, however, a simple-hinged flap with drooping spoilers was
incorporated in the B787, a design that can only facilitate small flap adjustments in flight [51]. The
Airbus A350 with its adaptive drop hinge flap would be able to deploy its trailing edge device during
cruise without creating a gap [52]. So far, certification activities have focused on load alleviation for
take-off and gusts. In general, this is an application of almost conventional high-lift devices for variable
camber. The next step could be a stronger segmentation of control devices up to the point where the
actuation and kinematics make a further increase uneconomical. This optimum can be found only by
REFERENCES 119

increasing the level of assessment from a sole aerodynamic view to the overall aircraft level. Aspects of
weight implications because of changes in the system architecture (e.g., kinematics, actuation, power
demand) have to be taken into account in order to increase the level of confidence toward the predicted
benefits. This is even more crucial if morphing is used for primary control functions because safety
aspects and subsequent redundancies must be considered. In addition, the topic of controlling these
devices has to be investigated further. Although publications concerning control date back to the
1980s [53], it has to be clarified if additional sensors (e.g., static ports at the trailing edge) are needed.
Since deforming morphing surfaces are a big step away from currently produced aircraft, their intro-
duction is most likely to be expected after the application of conventional control surfaces for morph-
ing. The ever-growing pressure on the aircraft manufacturers to maximize revenues and mitigate risks
in certification and industrial production of new components pose a big challenge to the introduction of
morphing by deformation. Ongoing and advancing research in materials, actuation concepts and flight-
testing for morphing by deformation continuously decrease the gap toward an industrial application
and thereby decrease the risk for the application by aircraft manufacturers.

REFERENCES
[1] J. Valasek (Ed.), Morphing Aerospace Vehicles and Structures, American Institute of Aeronautics and
Astronautics, Chichester, West Sussex/Reston, VA, 2012.
[2] F. Culick, Wright brothers: first aeronautical engineers and test pilots, in: Proc. 45th Annual Symposium The
Society of Experimental Test Pilots, Los Angeles, CA, 26–29 September, 2001.
[3] H.F. Parker, The Parker variable camber wing, NACA Technical Report 77, 1920.
[4] S. Spooner, The Makhonine Way, Flight—The Aircraft Engineer and Airship, The Royal Aero Club of the
United Kingdom, vol. 20(1220), 1932, pp. 417–418.
[5] D.R. Jenkins, T. Landis, J. Miller, American X-Vehicles: An Inventory—X-1 to X-50, Monographs in
Aerospace History No. 31, June 2003.
[6] Y. Gibbs, NASA Armstrong Fact Sheet: X-5 Research Aircraft, https://www.nasa.gov/centers/armstrong/
news/FactSheets/FS-081-DFRC.html (retrieved September 14, 2016).
[7] W.D. Painter, L.J. Caw, Design and Physical Characteristics of the Transonic Aircraft Technology (TACT)
Research Aircraft, NASA-TM-56048, January 1979, http://www.nasa.gov/centers/dryden/pdf/87897main_
H-976.pdf (retrieved March 4, 2016).
[8] R. Hardy, AFTI/F-111 mission adaptive wing technology demonstration program, in: Aircraft Prototype and
Technology Demonstrator Symposium, 1983, pp. 121–126.
[9] W.W. Gilbert, Mission adaptive wing system for tactical aircraft, J. Aircr. 18 (7) (1981) 597–602, https://doi.
org/10.2514/3.57533.
[10] S. Goecke Powers, L.D. Webb, E.L. Friend, W.A. Lokos, Flight Test Results From a Supercritical Mission
Adaptive Wing With Smooth Variable Camber: NASA Technical Memorandum 4415, NASA, Washington,
DC, USA, 1992.
[11] R.W. DeCamp, R. Hardy, D.K. Gould, in: Mission-adaptive wing, International Pacific Air and Space Tech-
nology Conference and Exposition, vol. 6, SAE International, Warrendale, PA, 1987, pp. 1795–1801.
[12] D. Schmitt, Die aerodynamische Auslegung des Airbus A300-600, in: DGLR (Ed.), DGLR Jahrbuch, 83-123,
1983.
[13] J. Mantel, Leistungssteigernde Modifikationen an transsonischen Tragfl€ ugeln von Verkehrsflugzeugen, in:
Jahrestagung, DGLR-NR, 82-031, 1982.
[14] A.O. Awani, The Investigation of a Variable Camber Blade Lift Control for Helicopter Rotor Systems,
National Aeronautics and Space Administration, NASA Contractor Report, 1982.
120 CHAPTER 3 THE DEVELOPMENT OF MORPHING WING RESEARCH
AND TECHNOLOGY

[15] L.U. Dadone, Design and Analytical Study of a Rotor Airfoil, National Aeronautics and Space Administra-
tion, NASA Contractor Report, 1978.
[16] L.R. Centolanza, E.C. Smith, B. Munsky, Induced-shear piezoelectric actuators for rotor blade trailing edge
flaps, Smart Mater. Struct. 11 (1) (2002) 24. http://stacks.iop.org/0964-1726/11/i¼1/a¼303.
[17] G.A. Trippensee, D.P. Lux, X-29A Forward-Swept-Wing Flight Research Program Status, National
Aeronautics and Space Administration, NASA Technical Memorandum, November 1987.
[18] F.A. Johnsen, Sweeping Forward. Developing and flight testing the Grumman X-29A forward Swept Wing
Research Aircraft, Library of Congress Cataloging-in-Publication Data, 2013.
[19] J. Szodruch, R. Hilbig, Variable wing camber for transport aircraft. Prog. Aerosp. Sci. 25 (3) (1988) 297–328,
https://doi.org/10.1016/0376-0421(88)90003-6.
[20] E. Greff, The development and design integration of a variable camber wing for long/medium range aircraft,
Aeronaut. J. 94 (939) (1990) 301–312.
[21] H. Hilbig, H. Wagner, Messerschmitt-B€olkow-Blohm GmbH/Vereinigte Flugtechnische Werke GmbH, Var-
iable wing camber control for civil transport aircraft, in: 14th Congress of the International Council of the
Aeronautical Sciences, 1984, pp. 243–248.
[22] J.J. Spillman, The use of variable camber to reduce drag, weight and costs of transport aircraft, Aeronaut. J.
96 (951) (1992) 1–9.
[23] V.J. Rossow, Lift Enhancement by Trapped Vortex, National Aeronautics and Space Administration, NASA
Technical Report, 1992.
[24] V.J. Rossow, Lift enhancement by an externally trapped vortex, J. Aircr. 15 (9) (1978) 618–625, https://doi.
org/10.2514/3.58416.
[25] A.L. Martins, F.M. Catalano, Aerodynamic optimization study of a mission adaptive wing for transport air-
craft, in: 15th Applied Aerodynamics Conference, 1997.
[26] A.L. Martins, R.S. Ribeiro, Aircraft induced drag minimization using a constrained optimization method, in:
20th Congress of the International Council of the Aeronautical Sciences, 1996, pp. 2549–2557.
[27] A.L. Martins, F.M. Catalano, Drag optimization for transport aircraft mission adaptive wing, in: 38th Aero-
space Sciences Meeting and Exhibit, 2000.
[28] G. Dargel, H. Hansen, J. Wild, T. Streit, H. Rosemann, K. Richter, Aerodynamische Fl€ ugelauslegung mit
multifunktionalen Steuerfl€achen, in: DGLR (Ed.), DGLR Jahrbuch, vol. 1, DGLR, 2002, p. 1605.
[29] R.M. Ajaj, C.S. Beaverstock, M.I. Friswell, Morphing aircraft: the need for a new design philosophy, Aerosp.
Sci. Technol. 49 (2013) 154–166, https://doi.org/10.1016/j.ast.2015.11.039.
[30] H.P. Monner, T. Bein, H. Hanselka, E. Breitbach, Design aspects of the adaptive wing—the elastic trailing
edge and the local spoiler bump, in: RAS—Multidisciplinary Design and Optimisation Conference, 1998, pp.
15-1–15-9.
[31] H.P. Monner, H. Hanselka, E.J. Breitbach, Development and design of flexible Fowler flaps for an adaptive
wing, in: 5th Annual International Symposium on Smart Structures and Materials, SPIE, 1998, pp. 60–70.
[32] I. Dimino, G. Diodati, A. Concilio, A. Volovick, L. Zivan, Distributed electromechanical actuation system
design for a morphing trailing edge wing, SPIE Smart Structures/NDE, Las Vegas, Nevada (USA), March
2016. in: Proc. SPIE 9801, Industrial and Commercial Applications of Smart Structures Technologies 2016,
980108, April 16, 2016. https://doi.org/10.1117/12.2219223.
[33] D. M€uller, Das Hornkonzept. Realisierung eines formvariablen Tragfl€ ugelprofils zur aerodynamischen Leis-
tungsoptimierung zuk€unftiger Verkehrsflugzeuge, Universit€at Stuttgart, Stuttgart, Germany, 2000.
[34] W. Raither, E. Furger, M. Zundel, A. Bergamini, P. Ermanni, Variable-stiffness skin concept for camber-
morphing airfoils, J. Intell. Mater. Syst. Struct. 26 (13) (2015) 1609–1621, https://doi.org/10.1177/
1045389X14546780.
[35] N.J. Pfeiffer, J.G. Christians, Active Tailoring of Lift Distribution to Enhance Cruise Performance, NASA
Langley Research Center, Contract Report, April 1, 2005.
[36] J. Hetrick, R. Osborn, S. Kota, P. Flick, D. Paul, Flight testing of mission adaptive compliant wing, in: 48th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 2007.
REFERENCES 121

[37] C.Y. Herrera, N.D. Spivey, G. Ervin, P. Flick, Aeroelastic airworthiness assessment of the adaptive compliant
trailing edge flaps, in: 46th SFTE Annual International Symposium, 2015.
[38] J. Urnes, N. Nguyen, A mission adaptive variable camber flap control system to optimize high lift and cruise
lift to drag ratios of future N + 3 transport aircraft, in: 51st AIAA Aerospace Sciences Meeting including the
New Horizons Forum and Aerospace Exposition, 2013.
[39] N. Nguyen, A framework for adaptive aeroelastic wing shaping control, Presentation at the Advanced
Modeling & Simulation (AMS) Seminar Series, April 22, 2014.
[40] C.A. Ippolito, N.T. Nguyen, J. Totah, K. Trinh, E.B. Ting, Initial assessment of a variable-camber continuous
trailing-edge flap system for drag-reduction of non-flexible aircraft in steady-state cruise condition, in: AIAA
Infotech@Aerospace (I@A) Conference, 2013.
[41] A. de Gaspari, S. Ricci, A two-level approach for the optimal design of morphing wings based on compliant struc-
tures, J. Intell. Mater. Syst. Struct. 22 (10) (2011) 1091–1111, https://doi.org/10.1177/1045389X11409081.
[42] A. de Gaspari, S. Ricci, L. Travaglini, L. Cavagna, A. Antunes, F. Odaguil, G. Lima, Active camber morphing
wings based on compliant structures: an aeroelastic assessment, in: 53rd AIAA Aerospace Sciences Meeting,
2015.
[43] A. de Gaspari, S. Ricci, L. Riccobene, Design, manufacturing and wind tunnel test of a morphing wing based
on compliant structures, in: 24th AIAA/AHS Adaptive Structures Conference, 2016.
[44] M. Kintscher, J. Kirn, S. Storm, F. Peter, Assessment of the SARISTU Enhanced Adaptive Droop Nose, in: P.
C. W€olcken, M. Papadopoulos (Eds.), Smart Intelligent Aircraft Structures (SARISTU), first ed., Springer
International Publishing, Cham, Switzerland, 2016, pp. 113–140.
[45] R. Pecora, F. Amoroso, M. Magnifico, I. Dimino, A. Concilio, KRISTINA: kinematic rib-based structural
system for innovative adaptive trailing edge, in: Proc. SPIE—The International Society for Optical
EngineeringVolume 9801, (2016), Article No. 980107.
[46] A. Wildschek, S. Storm, M. Herring, D. Drezga, V. Korian, O. Roock, Design, optimization, testing, veri-
fication, and validation of the wingtip active trailing edge, in: P.C. W€ olcken, M. Papadopoulos (Eds.), Smart
Intelligent Aircraft Structures (SARISTU), first ed., Springer International Publishing, Cham, Switzerland,
2016, pp. 219–255.
[47] F. Peter, E. Stumpf, G.M. Carossa, M. Kintscher, I. Dimino, A. Concilio, R. Pecora, A. Wildschek, Morphing
value assessment on overall aircraft level, in: P.C. W€ olcken, M. Papadopoulos (Eds.), Smart Intelligent Air-
craft Structures (SARISTU), first ed., Springer International Publishing, Cham, Switzerland, 2016,
pp. 859–871.
[48] F.N. Peter, K. Risse, F. Schueltke, E. Stumpf, Variable camber impact on aircraft mission planning, in: 53rd
AIAA Aerospace Sciences Meeting, 2015.
[49] I. Dimino, A. Concilio, R. Pecora, Safety and reliability aspects of an adaptive trailing edge device (ATED),
in: Proc. 24th AIAA/AHS Adaptive Structures Conference, San Diego, United States, January 4–8 2016.
[50] S. Sakurai, S.J. Fox, K.W. Beyer, D.S. Lacy, P.L. Johnson, S.L. Wells, J.S. Noble, P.T. Meredith, N.
V. Huynh, R.R. Christianson, Multi-Function Trailing Edge Devices and Associated Methods, US Patent
No. US 7,243,881 B2, 2007.
[51] M. Goldhammer, The Next Decade in Commercial Aircraft Aerodynamics Aircraft Aerodynamics. A Boeing
Perspective, Aerodays 2011, European Commission, Madrid, Spain, 2011.
[52] H. Str€uber, The aerodynamic design of the A350 XWB-900 high lift system, in: 29th Congress of the
International Council of the Aeronautical Sciences, 2014.
[53] J.H. Renken, Mission-adaptive wing camber control systems for transport aircraft, in: 3rd Applied
Aerodynamics Conference, 1985.

You might also like