ArticlePDF Available

Composite corrugated structures for morphing wing skin applications

Authors:

Abstract and Figures

Composite corrugated structures are known for their anisotropic properties. They exhibit relatively high stiffness parallel (longitudinal) to the corrugation direction and are relatively compliant in the direction perpendicular (transverse) to the corrugation. Thus, they offer a potential solution for morphing skin panels (MSPs) in the trailing edge region of a wing as a morphing control surface. In this paper, an overview of the work carried out by the present authors over the last few years on corrugated structures for morphing skin applications is first given. The second part of the paper presents recent work on the application of corrugated sandwich structures. Panels made from multiple unit cells of corrugated sandwich structures are used as MSPs in the trailing edge region of a scaled morphing aerofoil section. The aerofoil section features an internal actuation mechanism that allows chordwise length and camber change of the trailing edge region (aft 35% chord). Wind tunnel testing was carried out to demonstrate the MSP concept but also to explore its limitations. Suggestions for improvements arising from this study were deduced, one of which includes an investigation of a segmented skin. The overall results of this study show that the MSP concept exploiting corrugated sandwich structures offers a potential solution for local morphing wing skins for low speed and small air vehicles.
Content may be subject to copyright.
This content has been downloaded from IOPscience. Please scroll down to see the full text.
Download details:
IP Address: 137.222.181.79
This content was downloaded on 01/06/2015 at 14:07
Please note that terms and conditions apply.
Composite corrugated structures for morphing wing skin applications
View the table of contents for this issue, or go to the journal homepage for more
2010 Smart Mater. Struct. 19 124009
(http://iopscience.iop.org/0964-1726/19/12/124009)
Home Search Collections Journals About Contact us My IOPscience
IOP PUBLISHING SMART MATERIALS AND STRUCTURES
Smart Mater. Struct. 19 (2010) 124009 (10pp) doi:10.1088/0964-1726/19/12/124009
Composite corrugated structures for
morphing wing skin applications
C Thill1,JAEtches
2,IPBond
2,3, K D Potter2and P M Weaver2
1Messier-Dowty, elizy, France
2Advanced Composites Centre for Innovation and Science (ACCIS), University of Bristol, UK
E-mail: i.p.bond@bristol.ac.uk
Received 22 June 2010, in final form 28 September 2010
Published 11 November 2010
Online at stacks.iop.org/SMS/19/124009
Abstract
Composite corrugated structures are known for their anisotropic properties. They exhibit
relatively high stiffness parallel (longitudinal) to the corrugation direction and are relatively
compliant in the direction perpendicular (transverse) to the corrugation. Thus, they offer a
potential solution for morphing skin panels (MSPs) in the trailing edge region of a wing as a
morphing control surface. In this paper, an overview of the work carried out by the present
authors over the last few years on corrugated structures for morphing skin applications is first
given. The second part of the paper presents recent work on the application of corrugated
sandwich structures. Panels made from multiple unit cells of corrugated sandwich structures are
used as MSPs in the trailing edge region of a scaled morphing aerofoil section. The aerofoil
section features an internal actuation mechanism that allows chordwise length and camber
change of the trailing edge region (aft 35% chord). Wind tunnel testing was carried out to
demonstrate the MSP concept but also to explore its limitations. Suggestions for improvements
arising from this study were deduced, one of which includes an investigation of a segmented
skin. The overall results of this study show that the MSP concept exploiting corrugated
sandwich structures offers a potential solution for local morphing wing skins for low speed and
small air vehicles.
(Some figures in this article are in colour only in the electronic version)
Nomenclature
caerofoil chord, m
Cddrag coefficient, non-dimensional
Cd0 zero-lift drag coefficient,
non-dimensional
Cllift coefficient, non-dimensional
Clmax maximum lift coefficient,
non-dimensional
Cppressure coefficient, non-dimensional
dthickness of MSP, m
dCl/dαlift curve slope, radians1
(Exx)lam
equ longitudinal equivalent tensile elastic
modulus of corrugated laminate, Pa
(Exx)san
equ longitudinal equivalent tensile elastic
modulus of MSP, Pa
3Address for correspondence: Department of Aerospace Engineering,
University Walk, Queen’s Building, Bristol BS8 1TR, UK.
(Eyy)lam
equ transverse equivalent tensile elastic mod-
ulus of corrugated laminate, Pa
(Eyy)san
equ transverse equivalent tensile elastic mod-
ulus of MSP, Pa
Lmax maximum lift force that the MSP
demonstrator sees during testing, N
lMSP length of MSP, m
Paxial force applied to MSP, N
Pactuation force, N
pzaerodynamic pressure applied
transversely to each MSP, Pa
saerofoil span, m
tmax aerofoil maximum thickness, m
Uair velocity, m s1
(uzz)MSP,central central deflection of MSP due to aerody-
namic loads, m
wMSP width of MSP, m
x/cnon-dimensional chord, non-dimensional
0964-1726/10/124009+10$30.00 ©2010 IOP Publishing Ltd Printed in the UK & the USA1
Smart Mater. Struct. 19 (2010) 124009 C Thill et al
αangle of attack, deg
Cd0 change in zero-lift drag coefficient, non-
dimensional
caerofoil chord length change, m
cmax maximum aerofoil chord length change,
m
δtrailing edge control surface deflection
angle, deg
xx)lam
equ longitudinal structural equivalent tensile
elastic strain of corrugated laminate, non-
dimensional
xx)san
equ longitudinal structural equivalent tensile
elastic strain of MSP, non-dimensional
yy)lam
equ transverse structural equivalent tensile
elastic strain of corrugated laminate, non-
dimensional
yy)san
equ transverse structural equivalent tensile
elastic strain of MSP, non-dimensional
θaerofoil trailing edge section angle
relative to the horizontal, deg
1. Introduction
Morphing technology applied to air vehicles allows for
aerodynamic efficiency improvement, expansion of flight
envelope and increased mission capabilities (Thill et al 2008b).
Over the last two decades research projects such as NASA’s
Morphing Project (Florance et al 2003, Bartley-Cho et al 2004,
Kudva 2004), DARPA’s Smart Wing programme (Florance
et al 2003, Bartley-Cho et al 2004, Kudva 2004) and Morphing
Aircraft Structures programme (Love et al 2004,Bowmanet al
2007, Bye and McClure 2007, Love et al 2007) developed and
applied morphing concepts in technology demonstrators for
wind tunnel and flight tests of small unmanned air vehicles.
A major challenge was the design of a skin material that
combines the two inherently different mechanical properties:
stiffness and compliance. A morphing skin needs to allow
large shape change while also being able to resist deformation
under aerodynamic loading. These requirements can be
treated separately when the compliant and stiff directions
differ. Such an application could be a morphing trailing edge
control surface of a wing. In such an application the skin
needs to allow camber and/or chord length change in the
chordwise direction and can be stiff in the spanwise direction
to resist the applied loading. Thus, a morphing skin material
or structure needs to be developed that exhibits orthotropic
stiffness properties.
Corrugated structures exhibit such properties: they are
stiff longitudinal to the corrugation direction and compliant
in the transverse direction (Thill et al 2007). Corrugated
structures in general are not new to the engineering or
natural world and have been used over many decades in civil,
naval, automotive and aerospace applications due to their
efficient performance, i.e. low mass to stiffness (longitudinal to
corrugations) ratio, buckling resistance and energy absorption
(Shanley 1947, Perel and Libove 1978, Wiggenraad et al 2001,
Sun and Spencer 2005). In Nature, numerous insects, such
as the damselfly, desert locust and dragonfly, have corrugated
wings for deployable mechanisms, stiffness requirements
and/or aerodynamic effects (Wootton 1981,1990,1992,
Herbert et al 2000, Wootton et al 2000). Yokozeki et al
(2006) were among the first to suggest corrugated laminates
made from carbon/epoxy composites for morphing wing skin
applications.
The present authors have investigated the corrugated
concept in more detail. An overview of this work is given in
the next section. The second part of this paper presents work
on the application of these corrugated structures when used
as morphing skin panels (MSPs) in the trailing edge region
of a scaled morphing aerofoil section. The aerofoil section
features an internal actuation mechanism that allows chordwise
length and camber change of the trailing edge region (aft 35%
of the chordwise section). Wind tunnel testing was carried
out to demonstrate the MSP concept but also to explore its
limitations. Suggestions for improvements arising from this
study were deduced, one of which, involving a secondary skin
on top of the MSP, was investigated.
2. Composite corrugated structures for morphing
skins: previous work
In a preliminary study (Thill et al 2007,2010b), tensile and
flexural experimental testing on various corrugated laminate
designs, both longitudinal and transverse to the corrugations,
was undertaken. The corrugated specimens were manufactured
using a variety of woven fibre/epoxy prepreg materials (aramid,
glass and carbon) and variables such as number of plies and
corrugation pitch were investigated. The aim was to investigate
the effects of constituent material and corrugation geometry on
the overall mechanical properties of the structure. The results
highlight the extreme anisotropic nature of these structures
and the primary dependence of the transverse tensile elastic
modulus on the laminate thickness. It was also found that the
use of pre-impregnated tape lay-up is impractical for consistent
corrugated laminate manufacturing quality. Resin transfer
moulding (RTM) processes were suggested as an alternative to
address some of the challenges and improve laminate quality
and reproducibility in a production environment. However,
developing such methods was considered beyond the scope of
this project and a manual manufacturing process was pursued.
A fundamental drawback of corrugated laminates is their
very low out-of-plane stiffness which is likely to be unable
to resist the aerodynamic loading at typical speeds required
for flight. Hence, a corrugated sandwich structure was
suggested wherein two corrugated laminates are bonded to a
foam core. In a second study, Thill et al (2008c)presented
experimental and analytical results for a sandwich structure
with trapezoidal corrugated skins. Furthermore, a parametric
study analysed the effects of corrugation geometry and form
(re-entrant, rectangular, round, trapezoidal, sinusoidal and
triangular) on the structural elastic strain and equivalent elastic
tensile modulus transverse to the corrugation direction. It
showed that for a rectangular corrugation, structural strains
of up to 50% can be realistically achieved if the stiffness is
not of primary concern. Conversely, corrugated laminates
with a transverse equivalent elastic modulus of about 4 GPa
2
Smart Mater. Struct. 19 (2010) 124009 C Thill et al
Initial design
dimensions & materials
Analysis
aerodynamic, structural
& kinematic
MSP design
unit cell dimensions
Revised design
dimensions, materials &
manufacturing
Figure 1. Design iteration routine.
can be designed. It has been shown that for fixed projected
dimensions, re-entrant corrugations give the highest strain
while triangular corrugations provide the highest modulus. The
results give a good indication of the wide envelope within
which corrugated laminates can be designed to meet the
requirements for morphing skin applications.
While corrugated structures offer a potential structural
solution for morphing wing skin applications, it seems to be
an aerodynamic drawback to have non-smooth wing surfaces.
Hence, aerofoils with corrugated trailing edges have been
studied experimentally and computationally using various
corrugation geometries and forms (sinusoidal, trapezoidal and
triangular) at different Reynolds numbers (Thill et al 2010a).
The investigation showed that the aerodynamic performance
is highly dependent on corrugation amplitude, wavelength,
gradient (combination of amplitude and wavelength) and
Reynolds number. Evidence is given highlighting that
penalties for having a non-smooth trailing edge surface can be
eliminated for the lift curve slope and minimized for the zero-
lift drag coefficient by adequately designing the corrugation
geometry and choosing the operating Reynolds number range.
One such consideration is to recess the corrugations below the
aerofoil profile line. A segmented skin on top of the corrugated
profile was suggested for the suction surface to reduce the drag
penalties still further.
The results of these earlier studies led to an overall
conclusion that composite corrugated sandwich structures
could offer a potential solution for morphing wing skin panels
for slow speed, lightweight and small air vehicles. One
particular application could be as a continuous multi-purpose
Figure 2. Final design of MSP demonstrator front aerofoil section
without skin.
trailing edge control surface panel. This concept is to be
proven in this study through the design, manufacturing and
wind tunnel testing of an MSP demonstrator.
3. Methodology
3.1. Design and analysis of MSP demonstrator
Several design iterations of the demonstrator layout and
material selection (figure 1) were carried out before the final
design was determined (figures 2and 3). A NACA 0024
aerofoil profile was selected for the following reasons.
A simple and relatively cheap actuation system needs to be
fitted into the aerofoil section at about 65% chord hence at
that position adequate aerofoil thickness of about 100 mm
is required.
The larger the corrugations of the morphing skin panel
are the easier it is to make using in-house hand lay-
up processes. This means about 5–10 mm in height;
furthermore, the height of one corrugated cell should
be around 1% chord or less for acceptable aerodynamic
performance and realistic scale (Thill et al 2008a). Hence,
thechordneedstobelessthan1m.
The demonstrator has to fit in a wind tunnel working
section of 1.52 m ×1.52 m ×2.13 m (5 ft ×7ft×7ft)
by keeping blockage effects to a minimum, hence keeping
the overall size of the demonstrator as small as possible.
Since the chord is limited to less than 1 m the span cannot
be more than 1.52 m (5 ft) to avoid excessive wind tunnel
blockage.
The above meant that a relatively thick aerofoil profile had
to be selected for which aerodynamic data were available. The
NACA 0024 profile fits the requirements. The final overall
dimensions of the demonstrator wing section are a span, s=
1.520 m, a chord, c=0.800 m, and a maximum thickness at
30% chord, tmax =0.192 m.
The angle of attack of the aerofoil section, the morphing
trailing edge control surface deflection angle (camber) and
3
Smart Mater. Struct. 19 (2010) 124009 C Thill et al
Figure 3. Final design of MSP demonstrator front aerofoil section with skin (left) and actuation system in between MSPs (right).
chord length can all be altered. The actuation system (figure 3)
of the trailing edge control surface consists of two cross-bar
elements that can be extended or retracted by turning a lead
screw manually using a torque wrench. This allows the chord
to be changed, c, to any length within the elastic limit of the
MSP. The camber can be changed by removing a locator pin
and adjusting the trailing edge deflection angle δaround the
65% chord position. Seven fixed positions are possible (0,
±6,±12,±18). Furthermore, 30 static pressure tappings
on one surface ahead of the MSP are incorporated in the
demonstrator to assess the flow behaviour ahead of the MSP
i.e. from 0% to 65% chord. The pressure tappings consist
of brass tubes with an outer diameter of 1.59 mm and a wall
thickness of 0.36 mm.
In order to find the maximum aerodynamic forces applied
to the MSP demonstrator, XFOIL (open source code), a
two-dimensional computational fluid dynamics (CFD) panel
method with viscous effects (Drela 1989), is used to carry out
an aerodynamic analysis at a Reynolds number 2×106and
Mach number 0.1 over a range of angles of attack for the
conventional NACA 0024. XFOIL predicts that about 2% of
the total lift is generated aft of 65% chord i.e. over the trailing
edge region. For the following analysis it is assumed that
5% of the total lift acts aft of 65% chord. This assumption
is conservative and simplified; it is used in the analysis to
dimension the structure and actuation mechanism.
The structural analysis is carried out only for some critical
parts of two load cases. It uses linear-elastic stress–strain
relationships; bending, torsion and shear analysisof closed thin
sections; bending and buckling analysis of thin plates; bending
and buckling analysis of beams; lug and thread analysis.
The relevant dimensions from the initial design are input
into a spreadsheet from where they are analysed by Matlab
(version 7.2.0.232; R2006a) to provide outputs of minimum
reserve factors for each analysed part and load case in a new
spreadsheet. A safety factor of 1.5 was used throughout. The
kinematic analysis of the actuation system yielded the forces
required to deform the MSP as a function of the cross-bar angle
and the resulting in-plane deflections.
During the design of the MSP the aim was to optimize
the geometry and form of the corrugated sandwich structure in
Figure 4. Schematic of MSP location relative to horizontal (x-axis).
terms of stiffnesses in the chord- and spanwise directions. The
maximum structural strain expected from the MSP is given by
yy)san
equ cmax
cos θwMSP
.(1)
The angle θcomes from the fact that the MSP is at an
angle relative to the y-direction which is dependent on the
aerofoil profile (figure 4; for NACA 0024 θ=16.5) while
cmax is equal to the elastic limit of the MSP. The maximum
force Papplied to each MSP depends on the actuation force
P(figure 4) and hence determines the maximum allowed
equivalent transverse tensile elastic modulus of the MSP:
(Eyy)san
equ P
dlMSP
1
yy)san
equ
.(2)
A two-dimensional beam bending analytical analysis and
finite element analysis (FEA) using ABAQUS/CAE (version
6.6-1) are carried out to determine the maximum central
deflection. The MSP is assumed to withstand the aerodynamic
pressure forces (distributed transverse load: lift force over
aerofoil section lifting area). Note that in reality the upper
panels take more load than the lower panels, however, for
simplicity both upper and lower panels are treated similarly.
The pressure is calculated to be
pz=0.05Lmax
cs .(3)
4
Smart Mater. Struct. 19 (2010) 124009 C Thill et al
Figure 5. Schematic of MSP corrugated unit cell.
Table 1. MSP geometry.
s
(mm)
t
(mm)
b
(mm)
h
(mm)
(deg)
b1
(mm)
b3
(mm)
12.0 0.3 6.0 4.9 0.0 2.0 4.0
e
(mm)
ncor
(—)
d
(mm)
LUC
(mm)
nUC
(—)
wMSP
(mm)
lMSP
(mm)
1.5 2 10.0 15.0 14 210 760
The maximum central deflection of the MSP based on simple
beam bending theory (clamped–clamped boundary conditions)
is (Young 1989)
(uzz)MSP,central =12 pzw4
MSP
384(Eyy)san
equd3.(4)
Since the equivalent elastic strain and modulus have been
fixed, the detailed dimensions of the MSP unit cell can be
back-calculated. The MSP is a corrugated sandwich panel
as described in Thill et al (2008c). Its equivalent moduli
in the transverse (yy) and longitudinal (xx) directions are
calculated using the analytical models for the corrugated
sandwich structure (Thill et al 2008c). The corrugation form
and geometry, and sandwich geometry must be manipulated
until the transverse equivalent structural elastic strain and
modulus of the MSP equal the allowed values from (1)and(2),
respectively. This could notbe matched exactly due to practical
limitations of the dimensions of the MSP. Note that from Thill
et al (2008c) it is best to have a re-entrant corrugation form for
a low modulus/high strain MSP, however, these are difficult to
manufacture hence the next best option is chosen: rectangular
corrugations (figure 5). Due to the in-house manufacturing
limitations, two panels were used for each surface (top and
bottom) i.e. four panels in total. The final dimensions and
properties of each MSP are given tables 1and 2, respectively.
3.2. Manufacturing
The corrugated laminates were manufactured from pre-
impregnated E-glass/977-2 epoxy tape (Cytec, UK) that was
Figure 6. Picture of laying-up process of prepreg onto tool.
Figure 7. Picture of rectangular corrugated sandwich structure.
Table 2. Mechanical properties of MSP.
(Exx)lam
equ
(GPa)
xx)lam
equ
(%)
(Exx)san
equ
(GPa)
xx)san
equ
(%)
61.36 1 3.83 1
(Eyy)lam
equ
(MPa)
yy)lam
equ
(%)
(Eyy)san
equ
(MPa)
yy)san
equ
(%)
13.08 21 0.82 17
hand-laid onto an aluminium tool (figure 6). The prepreg,
sandwiched between layers of release film, was formed
corrugation by corrugation with the help of a heat gun to ease
forming of the prepreg onto the tool which has rectangular
grooves. Rectangular aluminium bars were located on top
and weights were used to hold the complete assembly in place
until the vacuum bagging process. After applying vacuum to
the entire tool, the assembly was cured at the recommended
prepreg cure cycle (180 C for 3 h at 689.5 kPa) in the autoclave
(Quicklock Thermoclave, Leeds & Bradford Boiler Company
Limited). The properties of the prepreg material used are given
in table 3.
Two types of corrugated sandwich panels were manufac-
tured: one without and one with an external discontinuous
segmented skin.
Two laminates manufactured from one ply of pre-
impregnated tape E-glass/977-2 epoxy (table 3) were made per
corrugated sandwich panel. The laminates were then bonded
to a foam core (Rohacell®31) using a two part epoxy paste
adhesive (Redux 810, Hexcel Composites). To accelerate the
cure of the adhesive the assembled sandwich structure was
put in an oven for 1 h at 70C. Figure 7shows a finished
rectangular corrugated sandwich structure.
5
Smart Mater. Struct. 19 (2010) 124009 C Thill et al
Figure 8. Schematic (left) and picture (right) of segmented skin bonded to corrugated laminate.
Table 3. Properties of prepregs.
Va
fρaE11 =Eb
22 Gc
12 νc
12
Fibre type Weave Resin (%) (kg m3) (GPa) CoV (GPa) (—)
Glass Five harness satin Cytec 977-2 42.7 2080 23.3 2% 5 0.2
aData from manufacturer. bData from experiment. cEstimated.
Figure 9. Picture of morphing skin panel (MSP).
For the corrugated sandwich panel with an external skin,
first the panel was manufactured as described above. Then
20 mm wide strips (segments) made from one ply of pre-
impregnated tape E-glass/977-2 epoxy (table 3) were bonded
on to one surface of the panel using a two part epoxy paste
adhesive (Redux 810, Hexcel Composites). To accelerate the
cure of the adhesive the assembled sandwich structure was put
in an oven for 1 h at 70 C. The strips are bonded as shown in
figure 8to form a segmented outer skin.
The individual components of the demonstrator were
either procured commercially or manufactured in-house. The
MSPs were manufactured as described in the previous two
sections. On each end an aluminium connector was bonded
using a two part epoxy paste adhesive (Redux 810, Hexcel
Composites) to the laminates to allow installation of the MSPs
on the demonstrator. A completed MSP is shown in figure 9.
Note that the corrugations are recessed below the aerofoil
profile line as recommended by Thill et al (2008a). It has
nominal projected dimensions of 760 mm ×258 mm×10 mm.
The MSPs were fastened through their connectors to the
secondary and tertiary spars of the demonstrator using M5
socket head countersunk screws.
3.3. Experimental testing
Two types of experimental tests were undertaken: in and out
of the wind tunnel. The out of wind tunnel testing validated
the structural and kinematic integrity of the MSP demonstrator.
This comprised taking the actuation system through its entire
range of movement of trailing edge control surface chord
length and camber change to guarantee kinematic reliability
in the wind tunnel.
The wind tunnel testing was carried out in the University
of Bristol large low speed wind tunnel which has an octagonal
test section with dimensions 2.13 m (7 ft) by 1.52 m (5 ft) and
a maximum speed of 60 m s1. The effects of lift, drag and
pitching moment coefficients on the aerofoil with extended and
deflected morphing trailing edge control surface at different
angles of attack and wind speeds were investigated. The actual
forces and moments of the MSP demonstrator were measured
using an OR6-7 2000 series force platform (AMTI) which uses
strain gauges to measure three force and three moment compo-
nents. It is interfaced via a six-channel strain gauge amplifier
(MSA-6 MINIAMP, AMTI) to a dSpace (DS1103/CLP1103)
unit using six BNC connectors. The air velocity is recorded
via a FCO510 Micromanometer (Furness Controls Limited)
connected to the dSpace unit via a RS232 connector. The
dSpace unit is connected to a PC data logger via ‘ControlDesk’
software with data acquired at a rate of 100 Hz for 10 s.
The static surface pressures on one surface of the aerofoil
section ahead of the MSP were measured electronically using
piezoresistive pressure transducers. The measurement and data
acquisition system (Scanivalve Corporation) uses an electronic
pressure scanner (ZOC22B32PxX2) connected via tubes to the
pressure tappings, an analogue to digital converter (RAD3200)
connected to a PC via USB and a power supply (RPM 1000).
Data are acquired via ‘Radlink’ software at a rate of 100 Hz for
10 s.
The MSP demonstrator is mounted vertically in the wind
tunnel to minimize blockage effects and to make interfacing
with the force platform easier. The force platform is attached
to the demonstrator via an aluminium interface to one end of
the secondary spar. The force platform and the interface are
located underneath the wind tunnel test section floor. In order
6
Smart Mater. Struct. 19 (2010) 124009 C Thill et al
Figure 10. Pictures of out of wind tunnel tests.
to ensure two-dimensional flow over the majority of the MSP
demonstrator, the gap between each of its ends and the side
walls of the wind tunnel was restricted to 2.5 mm (less than
0.2% of the full height of the tunnel). Note that the forces
and moments are usually measured at the quarter chord point,
however, for stability reasons in this case, the measurements
were taken at the secondary spar location (about 65% chord)
which coincided with the centre of gravity. Note that the
calibration factors of the force platform and the amplifier were
taken into account. The wind tunnel data were also corrected
for free stream conditions using three theoretical correction
factors to account for streamline curvature, solid and wake
blockage. These were applied using methods suggested by
Barlow et al (1999).
Five tests (runs) were carried out for the fully retracted
configuration, while a single run was carried out for the other
configurations. Data points given in this article are averaged
over the 10 s data acquisition interval for each run; the
corresponding coefficient of variation is calculated.
4. Results and discussion
The out of wind tunnel testing proved the integrity of the
actuation system and the MSPs under deformation. Figure 10
shows the morphing trailing edge section of the demonstrator
(aft 35% chord) with the upper and lower MSPs as it is
stretched and deflected. In these figures the tip is displaced
up to 4% chord and deflected up to 12. Note that the
lower surface bulges inwards when it is not under tension.
If this lower surface is the pressure surface, no significant
aerodynamic penalty is expected, although more detailed
studies would have to be undertaken for verification.
The repeatability of the five wind tunnel tests for the
fully retracted configuration is shown in figures 11 and 12 for
free stream velocities of 20 and 30 m s1. Generally, good
repeatability is achieved for both cases, however for the case
at the higher velocity the drag data are inconsistent. This
is attributed to interference of the MSP demonstrator with
the wind tunnel walls that contract during operation. Hence,
the aerodynamic data for the higher velocity case need to be
treated with care as absolute values cannot be taken as being
representative but relative changes are comparable.
Since no published experimental data for an aerofoil like
the MSP demonstrator are publicly available it is difficult to
Figure 11. Polar plot of lift and drag coefficients for MSP
demonstrator at U=20 m s1.
Figure 12. Polar plot of lift and drag coefficients for MSP
demonstrator at U=30 m s1.
validate the present experimental results. As a reference the
data for a conventional NACA 0024 aerofoil can be used for a
preliminary validation. These are taken from published data
(Sarpkaya 1973) at a Reynolds number of 0.82 ×106and
from XFOIL including viscous effects at a Reynolds number
of 1.1×106and Mach number of 0.06 over a range of angles
of attack. Figure 13 shows the lift coefficient against angle of
attack curves for the XFOIL data for the conventional NACA
0024 and the experimental data for the MSP demonstrator. The
data for three cases are presented when the morphing trailing
edge control surface is retracted (δ=0;c=0), extended
(δ=0;c=2.5%c), and deflected (δ=12;c=0). For
7
Smart Mater. Struct. 19 (2010) 124009 C Thill et al
Figure 13. Validation of experimental results: lift coefficients against
angle of attack for NACA 0024 from XFOIL and for MSP
demonstrator from experimental results at U=20 m s1.
Figure 14. Effect of morphing trailing edge control surface chord
length on lift coefficients against angle of attack at U=20 m s1.
the same cases table 4compares lift curve slopes, maximum
lift coefficients and zero-lift drag coefficients for the XFOIL,
published and experimental data.
From these results it can be concluded that the lift curve
slope does not change significantly as the control surface
is extended or deflected and the maximum lift coefficient
increases as the surface is deflected. The zero-lift drag
coefficient increases as the surface is deflected as predicted by
the XFOIL results. These results agree well with published
data for conventional trailing edge control surface deflections
(Kermode 1996). In terms of validation of the experimental
data with the XFOIL and published data, the lift curve slope
and maximum lift coefficient agree well, however, the zero-
lift drag coefficients have been measured to a lesser degree
of agreement. The possible reasons are the wind tunnel wall
interference as was discussed above and the corrugated trailing
edge region as was shown in Thill et al (2008a).
Figures 14 and 15 reiterate the same findings by showing
the experimentally determined lift coefficients of the MSP
demonstrator with a morphing trailing edge control surface
with respect to changing chord length and surface deflection.
No changes are observed when changing the chord length
while the maximum lift coefficient increases with control
surface deflection.
From figures 14 and 15, a further observation can be
made: the stall angle decreases when the morphing trailing
Figure 15. Effect of morphing trailing edge control surface
deflection on lift coefficients against angle of attack at
U=20 m s1.
Figure 16. Coefficients of pressure against non-dimensional chord of
suction surface of MSP demonstrator with morphing corrugated
trailing edge control surface fully retracted.
edge control surface is deflected or extended. This agrees with
the pressure plots for the fully retracted and fully deflected and
extended morphing trailing edge control surface (figures 16
and 17). For the fully deflected and extended case (δ=12;
c=2.5%c; figure 17) the pressures on the suction surface
aft of 10% chord at an angle of attack of 15fall below the
values of the pressures at lower angles of attack. This is an
indication of separated flow over the suction surface which
decreases the total lift produced. Furthermore, for the fully
retracted case (δ=0;c=0%c; figure 16) the flow seems
to remain attached up to 65% chord up to 15angle of attack.
This suggests that the flow is still attached when it reaches the
MSP and thus changing its geometry will change the overall
aerodynamic performance of the aerofoil.
The results presented herein and the previous study by
Thill et al (2008a) highlight a major aerodynamic drawback
between a morphing aerofoil with a corrugated and a smooth
trailing edge region: increased generation of drag. One way of
mitigating this effect is a suggestion to introduce an external
discontinuous layer on the MSP suction surface which has
no effect on structural performance. Table 5compares the
lift curve slopes, maximum lift coefficients and zero-lift drag
coefficients of the previously tested corrugated arrangement
against the configuration with a discontinuous segmented skin
for the cases where the morphing trailing edge control surface
8
Smart Mater. Struct. 19 (2010) 124009 C Thill et al
Table 4. Validation of experimental results: lift curve slope, maximum lift coefficient and zero-lift drag coefficient for NACA 0024 from
XFOIL, from published experimental data (Sarpkaya 1973) and for MSP demonstrator from experimental results (coefficient of variation
given in brackets) at U=20 m s1.
XFOIL data Published experimental data Experimental data
dCl/dα
(rad1)
Clmax
(—)
Cd0
(—)
dCl/dα
(rad1)
Clmax
(—)
Cd0
(—)
dCl/dα
(rad1)
Clmax
(—)
Cd0
(—)
δ=0;c=0 4.6 1.2 0.009 5.3 1.0 0.020 5.1 1.1 (5%) 0.023 (10%)
δ=0;c=2.5%c5.0 1.3 0.009 5.0 1.1 (5%) 0.018 (7%)
δ=12;c=0 4.2 1.6 0.016 5.5 1.4 (4%) 0.024 (6%)
Table 5. Comparison of aerodynamic experimental data of MSP demonstrator without (corrugated) and with (segmented) discontinuous skin
(coefficient of variation given in brackets).
U=20 m s1U=30 m s1
dCl/dα
(rad1)
Clmax
(—)
Cd0
(—)
Cd0
(%)
dCl/dα
(rad1)
Clmax
(—)
Cd0
(—)
Cd0
(%)
δ=0;c=0 Corrugated 5.1 1.1 (5%) 0.0232 (10%) 4.6 1.0 (4%) 0.0241 (10%)
Segmented 4.6 1.1 (5%) 0.0179 (14%) 23 4.3 1.0 (2%) 0.0214 (7%) 11
δ=12;c=2.5%cCorrugated 4.4 1.3 (4%) 0.0220 (8%) 5.0 1.2 (5%) 0.0185 (7%)
Segmented 4.7 1.4 (8%) 0.0155 (7%) 30 5.0 1.3 (2%) 0.0101 (17%) 45
Figure 17. Coefficients of pressure against non-dimensional chord of
suction surface of MSP demonstrator with morphing corrugated
trailing edge control surface fully extended.
is retracted (δ=0;c=0) and fully deflected and extended
(δ=12;c=2.5%c). While no major differences can be
observed for the lift data a significant reduction in the zero-lift
drag coefficients is measured at both air flow speeds. While
at this point it cannot be stated how close these zero-lift drag
coefficients are to a NACA 0024 with a smooth morphing
trailing edge control surface, it is certain that a discontinuous
segmented skin improves the aerodynamic performance of the
MSP concept.
Another way to reduce the drag increase might be to use
a different, thicker, aerofoil. It is to be noted though that
the NACA0024 is a fairly thick aerofoil compared to most
conventional aerofoils. Nevertheless for a thicker aerofoil the
pressure drag should dominate (Anderson 2001) and hence
the skin friction drag created by the corrugations (they can
be considered as an extremely rough surface) becomes less
important. Thus the MSP concept might be an interesting
alternative solution for trailing edge control surfaces of thick
aerofoils. However further work is necessary to confirm this
assumption.
Apart from the increase in drag generated from
the corrugated surface, some additional limitations were
experienced during the wind tunnel testing. When the MSPs
were not under tension significant out-of-plane deflections and
vibrations were observed at wind speeds above 20 m s1.
This suggests that the concept may only be viable for small,
light and slow air vehicles. Should it be applied to larger
air vehicles, corresponding morphing ribs to support the
MSP without restricting their deformation might be necessary.
Furthermore, while an elastic chord length change of 4%
chord with no trailing edge control surface deflection was
achieved during the out of wind tunnel testing this may not
be enough for aircraft flight control. This chord length change
is variable subject to increasing the width of the MSP or the
transverse elastic structural strain capability of the MSP. The
latter can only be achieved at the expense of the out-of-plane
stiffness thus possibly requiring development of corresponding
morphing ribs.
A further point of discussion is the actuation power
required to morph the MSPs. Ideally the power required
to actuate a morphing skin and hold it in a desired shape
should be minimal. The holding power depends mainly on
the actuation system unless the skin shape can be ‘locked’,
i.e. shape memory materials (McKnight and Henry 2005, Reed
et al 2005). In the presented concept the lead screw allowed
for a zero holding power. The actuation power depends mainly
on the stiffness of the skin material. There has to be a
compromise between low skin stiffness (e.g. elastomeric based
concepts) for good morphing characteristics and low actuation
power and high stiffness (e.g. stiffness tailored composites)
for good load transfer. In theory shape memory materials do
not have this issue but their stiffness changes are typically
triggered by changes in temperature (Lendlein and Kelch 2002)
and thus require additional power; furthermore during the
low stiffness stage of the skin, the out-of-plane displacements
limit the operational airspeed, similar to the MSP concept.
Thus overall the MSP concept is believed to compare fairly
9
Smart Mater. Struct. 19 (2010) 124009 C Thill et al
similarly to other morphing skin concepts on an actuation
power requirements comparison.
5. Conclusions
This paper summarizes the main findings of extensive
structural and aerodynamic studies on corrugated composites
for morphing structures undertaken over the last few years.
Arranged in a corrugated sandwich structure their extremely
orthotropic properties can be effectively used in a morphing
trailing edge control surface application. A wide envelope
exists in which these structures can be designed to suit
the precise morphing skin requirements. For acceptable
aerodynamic performance the corrugation geometry and form
need to be designed adequately or an external discontinuous
segmented skin can be used to improve surface smoothness.
This study has combined the design, manufacture and
test of a wind tunnel demonstrator aerofoil with a morphing
trailing edge control surface using MSPs made from corrugated
sandwich structures. Chord length changes of up to 4% chord
and control surface deflections of up to 12, while maintaining
a continuous aerodynamic surface, were demonstrated in out of
wind tunnel testing. Wind tunnel testing showed the concept to
work well under aerodynamic loads at low speeds. The use of
a discontinuous segmented skin showed a significant reduction
in the drag generated by the corrugated trailing edge region
without altering the structural performance. The concept could
be further improved by the development of corresponding
morphing ribs to provide additional skin support.
Acknowledgments
The authors would like to thank the UK EPSRC (Engineering
and Physical Sciences Research Council) for their support and
funding of this work through the University Technology Strate-
gic Partnership—SMARTCOMP (EP/D03423X/1). Special
thanks go to the Faculty of Engineering workshop staff without
whose efforts this project would have been impossible: Lee
Winter, Russ Eyre and Alan Bishop; John Morrison, Michael
Powell, Chris Hunt and John Byles; Ian Chorley and Mike
Jones; Mark Fitzgerald; and to members of the Aerospace
Department academic staff: Drs Giuliano Allegri, Panagiotis
Margaris and Neil Taylor.
References
Anderson J D Jr 2001 Fundamentals of aerodynamics (New York:
McGraw-Hill)
Barlow J B et al 1999 Low-Speed Wind Tunnel Testing (New York:
Wiley)
Bartley-Cho J D et al 2004 Development of high-rate, adaptive
trailing edge control surface for the smart wing phase 2 wind
tunnel model J. Intell. Mater. Syst. Struct. 15 279–91
Bowman J et al 2007 Development of next generation morphing
aircraft structures 48th AIAA/ASME/ASCE/AHS/ASC Structures,
Structural Dynamics, and Materials Conf. (Honolulu, Hawaii)
Bye D R and McClure P D 2007 Design of a morphing vehicle 48th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics,
and Materials Conf. (Honolulu, Hawaii)
Drela M 1989 Xfoil: an analysis and design system for low reynolds
number airfoils Low Reynolds Number Airfoil Aerodynamics
(Notre Dame, IN)
Florance J P et al 2003 Contributions of the NASA Langley Research
Center to the DARPA/AFRL/NASA/Northrop Grumman Smart
Wing Program (Norfolk, VA) (Reston, VA: American Inst.
Aeronautics and Astronautics Inc.)
Herbert R C et al 2000 The hind wing of the desert locust
(Schistocerca gregaria Forskal) III. A finite element analysis of
a deployable structure J. Exp. Biol. 203 2945–55
Kermode A C 1996 Mechanics of Flight (Essex: Longman Group
Limited)
Kudva J N 2004 Overview of the DARPA smart wing project
J. Intell. Mater. Syst. Struct. 15 261–7
Lendlein A and Kelch S 2002 Shape-memorypolymers Angew.
Chem. Int. Edn 41 2035–57
Love M H et al 2004 Impact of Actuation Concepts on Morphing
Aircraft Structures (Palm Springs, CA) (Reston, VA: American
Inst. Aeronautics and Astronautics Inc.)
Love M H et al 2007 Demonstration of morphing technology
through ground and wind tunnel tests 48th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics,
and Materials Conf. (Honolulu, HI)
McKnight G and Henry C 2005 Variable Stiffness Materials for
Reconfigurable Surface Applications (San Diego, CA)
(Bellingham, WA: International Society for Optical
Engineering)
Perel D and Libove C 1978 Elastic buckling of infinitely long
trapezoidally corrugated plates in shear J. Appl. Mech. 45
579–82
Reed J L Jr et al 2005 Adaptive wing structures Smart Structures and
Materials and Nondestructive Evaluation for Health Monitoring
and Diagnostics (San Diego, CA) (Bellingham, WA: The
International Society for Optical Engineering)
Sarpkaya T 1973 Lift and drag measurements on a hydrofoil in dilute
polyox solutions Nature 241 114–5
Shanley F R 1947 Cardboard-box wing structures J. Aeronaut. Sci.
14 713–5
Sun H-H and Spencer J 2005 Buckling strength assessment of
corrugated panels in offshore structures Marine Struct.
18 548–65
Thill C et al 2007 Corrugated composite structures for aircraft
morphing skin applications ICAST: 18th Int. Conf. of Adaptive
Structures and Technologies (Ottawa, ON)
Thill C et al 2008a Aerodynamic analysis of corrugated trailing
edges for morphing wing applications The Aerodynamics of
Novel Configurations: Capabilities and Future Requirements
(London: Royal Aeronautical Society)
Thill C et al 2008b Morphing skins Aeronaut. J. 112 117–38
Thill C et al 2008c Experimental and parametric analysis of
corrugated composite structures for morphing skin applications
ICAST: 19th Int. Conf. on Adaptive Structures and Technologies
(Ascona)
Thill C et al 2010a Aerodynamic study of corrugated skins for
morphing wing applications Aeronaut. J. 114 237–44
Thill C et al 2010b Investigation of trapezoidal corrugated
aramid/epoxy laminates under large tensile displacements
transverse to the corrugation direction Composites A41 168–76
Wiggenraad J F M et al 2001 Development of a crashworthy
composite fuselage concept for a commuter aircraft National
Aerospace Laboratory NLR
Wootton R J 1981 Support and deformability in insect wings
J. Zool. 193 447–68
Wootton R J 1990 The mechanical design of insect wings Sci. Am.
263 114–20
Wootton R J 1992 Functional-morphology of insect wings Annu. Rev.
Entomol. 37 113–40
Wootton R J et al 2000 The hind wing of the desert locust
(Schistocerca gregaria Forskal) I. Functional morphology and
mode of operation J. Exp. Biol. 203 2921–31
Yokozeki T et al 2006 Mechanical properties of corrugated
composites for candidate materials of flexible wing structures
Composites 37 1578–86
Young W C 1989 Roark’s Formulas for Stress and Strain (New York:
McGraw-Hill)
10
... It is important to note that in any closed system relying solely on the deformation of a relatively stiff material, without an area where the morphing skin can contract significantly (such as with a sliding panel [85] or corrugated/flexible skin section [13,[86][87][88][89], which will be discussed later), the overall deformation will always be minimal. This emphasises the need, even for bending-dominated skins, to have a section capable of significant strain or to provide alternative means to compensate for the change in shape. ...
... Thill et al. [13] tested a corrugated composite skin with sliding covering panels (see Figure 2.6) on a wing section at 20 m s −1 and observed a 23% to 30% decrease in drag compared to an uncovered corrugation. Similar sliding coverings for corrugated skins were also considered by Previtali [87] and Jenett et al. [97]. ...
... Segmented sliding skin bonded to a corrugated sub-structure[13] ...
Thesis
Full-text available
Traditional fixed-wing aircraft have relied on rigid trailing edge control surfaces to modify the camber distribution which changes the pressure distribution over the wing. These trailing edge flaps, while proven to be effective in adjusting lift, produce a significant induced drag when stowed and deployed due to sharp edges in the segmented rigid elements and small gaps, resulting in an increased fuel burn. One promising solution are morphing wings. Unlike traditional flaps, morphing wings can change their shape continuously offering the adaptability to alter the profile for optimal aerodynamic performance throughout the entire flight mission. One of the main reasons why morphing wings are not widely adopted year is the competing design requirements that form the aerodynamic surface. It requires high out-of-plane stiffness to resist the aerodynamic pressures while also needing low in-plane stiffness to keep the actuation forces low. This implies that this challenge can only be addressed using purposely tailored structures with anisotropic properties. This study introduces such a solution in the form of the Geometrically Anisotropic ThermOplastic Rubber (GATOR) design strategy. At its core, the GATOR strategy is based on three foundational principles. Firstly, it leverages the potential of thermoplastic elastomers in combination with Fused Filament fabrication (FFF) 3D~printing, allowing complex geometries tailored for specific aerodynamic needs to be manufactured. Secondly, it employs multi-material 3D printing. This technique integrates materials with different stiffness properties into one single component. The third principle seeks to exploit the geometrical anisotropy and structural scaling laws to help decouple the competing design constraints of morphing skin. To evaluate the efficacy and advantages of the GATOR approach, comprehensive research was conducted, employing both experimental and numerical methods, on a zero Poisson's ratio MorphCore. This core was printed directly onto the skin sheet using stiff and flexible formulations of Thermoplastic Polyurethane (TPU) respectively. Initial printing trials were performed and were tested for their mechanical response in the primary loading directions. Additionally, a novel differential pressure experiment was developed to assess the local skin sheet's out-of-plane performance. A parameterised GATOR panel was examined using an FEA framework created for this analysis, based on the CAD and meshing software GMSH and the FEA solver Abaqus. This framework was employed to undertake a design space analysis and a sensitivity study. The focus was on understanding how design parameters might be adjusted to achieve desirable high out-of-plane and low in-plane stiffness characteristics, as well as under differential pressure loading for both small and large in- and out-of-plane deformations up to stretch ratios of 1.6. Results from this research underscored the inherent benefits of GATOR morphing skins. In particular, the study emphasised the transformative potential of multi-material 3D printing. Allowing the creation of composite materials in a single print cycle obviates the need for secondary processes, which is commonly proposed in the literature. This results not only in a simplified manufacturing process but also produces a very strong bond between the skin sheet and the core structure. The FEA study has shown that a parameterised GATOR skin can be manipulated to partially overcome the competing design requirements, where the core thickness has the largest contribution to increased out-of-plane stiffness. The geometrically coupled term β/L encompassing three design parameters has been shown to exhibit the optimal balance between axial and flexural rigidity when β/L ~ 0.1. In summary, it has been shown in this work that FFF 3D printed GATOR skins significantly reduce the complexity of extension dominated morphing skins and provide a broad design space open for further improvements. Most importantly, the GATOR morphing skin concept has been demonstrated to offer a promising solution for tailoring highly non-linear flexible skins, strategically utilising the geometry to produce a structure capable of meeting the competing design requirements.
... The Hexcel® SGP370-8H/8552 woven-fabric carbon/epoxy composite specimen (Hexcel specimen), is used in aerospace applications in primary and secondary aircraft structures (Hexcel, 2016). Most commonly, woven fabrics are used to fabricate aircraft skin structures (Thill, 2010;Botelho, Silva, Pardini, & Rezende, 2006). Pristine cross-ply [0/90/90/0] Hexcel specimens were used for fire testing and char removal experiments. ...
Technical Report
Full-text available
Fire can dramatically alter the exposed surfaces of composite aircraft structures in ways that inhibit forensic analysis. The char, soot, and several other fire by-products often mask relevant aspects of the structural damage morphology, impeding identification of the underlying failure mechanisms of composite materials. In this report, the effects of small-flame direct fire exposure and efficacy of char removal techniques on mechanically failed Cytec T40-800/Cycom® 5215 graphite/epoxy and pristine Hexcel® SGP370-8H/8552 carbon/epoxy specimens were examined. Small-flame vertical and horizontal Bunsen burner fire tests were performed on mechanically failed unnotched compression, short beam strength, and in-plane shear graphite/epoxy specimens. Vertical burn tests were conducted on pristine carbon/epoxy specimens for initial char removal assessment. Both visual inspection and scanning electron microscopy (SEM) were used to characterize the fracture surface morphology of the mechanically failed specimens, investigate the induced thermal damage due to fire exposure, and assess the effectiveness of the char removal techniques. The fire damage consisted of matrix decomposition, melt dripping, char and soot deposition, matrix cracking, delamination, and residual thickness increases due to explosive outgassing. The composite thermal degradation due to heat conduction, combustion, and/or thermal deformation was significantly affected by the specimen layup, ply orientation relative to the heat source, and the fracture surface morphology. Char removal approaches, including ultrasonication, chemical solvent soaking, liquid nitrogen dipping, and a combination of thermal cycling and ultrasonication, were explored, resulting in partial char removal. This research is intended to be related work to the Composite Failure Analysis Handbook, DOT/FAA/AR-91/23 | WL-TI-91-4032, published through a joint research effort by the U.S. Air Force at Wright-Patterson AFB and the FAA. It should be noted these tests were conducted using a small methane flame rather than a large pool fire. Therefore, conclusions pertaining to aircraft accidents involving fire may not be directly applicable. The next phase of this project will address the fuel type and flame intensity to more accurately represent realistic accident conditions.
... In addition, flexible skin is required to withstand sufficiently large global strains (on the order of 50-100%) for large deformation applications such as span change, chord change, and planform area change, which makes skin design more challenging (Olympio and Gandhi, 2010). To date, with the advancement of manufacturing technology, numerous skins based on different concepts have emerged, such as elastomeric silicone flexible skins (Kudva, 2004;Andersen and Cowan, 2007;Ahmad et al., 2022), corrugated composite flexible skins (Yokozeki et al., 2006;Thill et al., 2010;Dayyani et al., 2015), shape memory polymer flexible skins Bye and McClure, 2007;Sun et al., 2014), and sandwich core structure flexible skins consisting of honeycomb core layers and flexible elastic panels (Olympio and Gandhi, 2009, 2012. ...
... Composite corrugated plates are widely used in many branches of engineering practice, such as bridge decks, cardboard boxes and horticultural structures [1][2][3] due to their beneficial properties of high strength-to-mass ratio [4]. Recently, composite corrugated plates have been applied to the shape morphing wing skins because it is stiff enough to bear aerodynamic loads with the spanwise direction and flexible enough in the corrugation direction to realize large deformation [5,6]. The composite corrugated plate with one side exposing to supersonic flow may suffer from panel flutter caused by self-excited oscillation due to its potential application to morphing skins [7]. ...
Article
Full-text available
Composite corrugated plates have a great potential in the application to morphing wings. However, it takes high computational cost to conduct flutter analysis with detailed 3D finite element models due to its structural complexity. In this study, an analytical method is proposed for flutter stability analysis of composite corrugated plates in supersonic flow. The trapezoidal and sinusoidal composite corrugated plate is homogenized as an equivalent anisotropic plate based on an energy approach. The flutter model for the composite corrugated plate is derived based on Kirchhoff plate theory and the equivalent stiffness properties. The unsteady aerodynamic pressure is evaluated by using the supersonic piston theory in which the corrugated section is taken into account. Hamilton's principle with the assumed mode method is applied to formulate the aeroelastic equation of the composite corrugated plate. The eigenvalue criterion is utilized to reveal the flutter mechanism and evaluate the stability of composite corrugated plates in supersonic flow. The accuracy and reliability of the present method are verified by comparing aeroelastic responses with those obtained from commercial software. Parametric studies concerning different parametric variables are also conducted. It is shown that the proposed method has sufficient accuracy and requires less computational effort, providing a theoretical basis for the utilization of trapezoidal and sinusoidal composite corrugated plates in morphing wings.
... The replacement of conventional actuators such as electric or hydraulic motors and mechanisms in aerodynamic applications in the automotive industry or in aeronautics with such shape memory alloy hybrid composites (SMAHC) is vastly advantageous, because of the reduced complexity and savings in weight and installation space [1]. Typical applications are aerodynamic components such as wing profiles or rear spoilers [2]- [4]. ...
Conference Paper
Full-text available
In this work, we present first results on a new approach for the full integration of shape memory alloy (SMA) wires into fibre-plastic composites on material level. By combining textile technology and additive manufacturing methods we took advantage of the stereo lithography printing process (SLA) which enabled the formation of structures with graded stiffness and the usage of UV curing material at low processing temperatures, to integrate a hybrid glass fibre-SMA textile preform into a morphing composite structure, a so called shape memory alloy hybrid composites (SMAHC). This work also includes a first characterization of the shape deflection and actuation capability of the component as well as an assessment of the heat distribution inside the material resulting from electric activation.
... A common strategy for adapting to changes in the environment is morphing. For example, the morphing functionality, which raptors possess [36], has already been applied to industrial products (e.g., airplanes) [37][38][39][40]. In the field of tribology, several adaptive surfaces and systems have also been developed. ...
Article
Full-text available
With the increasing demand for active friction control, we newly proposed to use dielectrophoresis to change the flow of PG-droplet-containing PAO4 to reduce the friction coefficient. The friction result with a 1-mm roller shows 20% reduction in friction coefficient (from 0.065 to 0.052) at AC 100 V, and in situ observation exhibits that PG tracks are formed over the contact area. On the other hand, at a high bias of 1000 V, the friction coefficient increases to 0.065. In this situation, in situ observation exhibits that PG forms a horseshoe-shaped track covering only the roller edges. Controlled friction tests and FEM analysis using 5-mm rollers revealed a unique behavior; a balanced bias effectively attracts the PG to the roller surface, and surface forces can resist mild dielectrophoretic forces to spread the PG across the roller surface. The present study strongly suggests the importance that the bias strength should be controlled to achieve a balance between surface force and dielectrophoretic force in order to obtain excellent lubrication conditions.
... In particular, the packaging industry has widely employed corrugated boards in box structures [1,2], while civil structures have benefited from corrugated roofing sheets, walls and web beams [3,4]. In addition, the aerospace industry has explored the potential of corrugated structures in morphing wing structures [5][6][7][8][9][10]. Moreover, corrugated structures have been employed in numerous lightweight designs [11][12][13][14][15][16][17]. ...
Chapter
Morphing wing applications have been used to change various dimensional properties of aircraft. Various 2D and 3D parameters can be changed on the aircraft’s wings, tail surfaces, or fuselage with these applications. Two primary schools are becoming widespread today in these application areas: mechanisms school and smart surfaces that employ shape-memory materials and smart actuators. The fishbone active camber approach is a research field that focuses on controlling the deflection on the wing’s trailing edge. In this approach, a tendon-like structure is designed that can be brought into desired shapes by creating tension on the wing structure. In this study, previous studies on fishbone active camber were revisited, the present situation was evaluated, and a roadmap was put forward for future studies from the authors’ perspective.
Article
Full-text available
Reconfigurable and morphing structures can potentially provide a range of new functionalities including system optimization over broad operational conditions and multi-mission capability. Previous efforts in morphing surfaces have generally focused on small deformation of high stiffness structural materials (e.g. aluminum, CFRP) or large deformation of low stiffness non-structural materials (e.g. elastomers). This paper introduces a new approach to achieving large strains in materials with high elastic moduli (5 to 30+ GPa). The work centers on creating variable stiffness composite materials which exhibit a controllable change in elastic modulus (bending or axial) and large reversible strains (5-15%). Several prototype materials were prepared using a commercial shape memory polymer, and measurements on these materials indicate a controllable change in stiffness as a function of temperature along with large reversible strain accommodation. We have fabricated and tested several design variations of laminar morphing materials which exhibit structural stiffness values of 8-12 GPa, changes in modulus of 15-77x, and large reversible bending strain and recovery of 2% area change in specific sample types. Results indicate that significant controllable changes in stiffness are possible.
Article
Medical devices such as implants, surgical instruments, extracorporal devices, and wound covers, as well as controlled drug delivery systems (CDDS) require a specific combination of material properties and functions including, for example, mechanical stability, biocompatibility, and biofunctionality. Polymeric biomaterials are of high relevance for such applications, as properties and functions can be tuned in a wide range by only small defined variations of their chemical or morphological structure. The rapid progress in surgical techniques, especially in minimally invasive surgery, requires smart materials, which are capable of an active on-demand movement and which do not need to be removed in a second surgery. These challenges can be addressed by shape-memory polymers (SMPs) described in this chapter. SMPs are of high technological significance for biomedical applications as they enable on-demand predefined changes in the shape of a device upon exposure to a suitable stimulus. Multifunctional materials are obtained when the shape-memory effect is combined with an additional function such as hydrolytic degradability, biofunctionality, and controlled drug release. Selected biomaterials with shape-memory capability are presented, including data on their biocompatibility. The potential of SMPs as a platform technology for biomedical applications is sketched by an overview on SMP-based medical devices being developed and the potential use of SMPs as matrix for CDDS.
Article
Corrugated structures offer a potential solution for morphing wing skin applications due to their anisotropic behaviour that allows chordwise camber and length changes. Aerofoils with corrugated skins in the aft 1/3 of the chordwise section have been studied experimentally and computationally using various corrugation shapes and forms (sinusoidal, trapezoidal and triangular) at different Reynolds numbers. The study showed that the aerodynamic performance is highly dependent on corrugation amplitude, wavelength, gradient (combination of amplitude and wavelength) and Reynolds number. Evidence is given highlighting that penalties for having a nonsmooth surface in the aft 1/3 of the chordwise section of an aerofoil can be eliminated for the lift curve slope and minimised for the zero lift drag coefficient.
Conference Paper
An overview of the contributions of the NASA Langley Research Center (LaRC) to the DARPA/AFRL/NASA/ Northrop Grumman Corporation (NGC) Smart Wing program is presented. The overall objective of the Smart Wing program was to develop smart** technologies and demonstrate near-flight-scale actuation systems to improve the aerodynamic performance of military aircraft. NASA LaRC's roles were to provide technical guidance, wind-tunnel testing time and support, and Computational Fluid Dynamics (CFD) analyses. The program was divided into two phases, with each phase having two wind-tunnel entries in the Langley Transonic Dynamics Tunnel (TDT). This paper focuses on the fourth and final wind-tunnel test: Phase 2, Test 2. During this test, a model based on the NGC Unmanned Combat Air Vehicle (UCAV) concept was tested at Mach numbers up to 0.8 and dynamic pressures up to 150 psf to determine the aerodynamic performance benefits that could be achieved using hingeless, smoothlycontoured control surfaces actuated with smart materials technologies. The UCAV-based model was a 30% geometric scale, full-span, sting-mounted model with the smart control surfaces on the starboard wing and conventional, hinged control surfaces on the port wing. Two LaRC-developed instrumentation systems were used during the test to externally measure the shapes of the smart control surface and quantify the effects of aerodynamic loading on the deflections: Videogrammetric Model Deformation (VMD) and Projection Moiré Interferometry (PMI). VMD is an optical technique that uses single-camera photogrammetric tracking of discrete targets to determine deflections at specific points. PMI provides spatially continuous measurements of model deformation by computationally analyzing images of a grid projected onto the model surface. Both the VMD and PMI measurements served well to validate the use of on-board (internal) rotary potentiometers to measure the smart control surface deflection angles. Prior to the final entry, NASA LaRC also performed three-dimensional unstructured Navier Stokes CFD analyses in an attempt to predict the potential aerodynamic impact of the smart control surface on overall model forces and moments. Eight different control surface shapes were selected for study at Mach = 0.6, Reynolds number = 3.25×106, and + 2°, 3°, 8°, and 10° model angles-of-attack. For the baseline, undeflected control surface geometry, the CFD predictions and wind-tunnel results matched well. The agreement was not as good for the more complex aero-loaded control surface shapes, though, because of the inability to accurately predict those shapes. Despite these results, the NASA CFD study served as an important step in studying advanced control effectors.
Article
Buckling loads are calculated for a class of infinitely long trapezoidally corrugated plates under uniform in-plane edge shear. This class consists of corrugated plates in which all interior plate elements are of equal width, with the end elements clamped. The energy method is used as the basis of analysis, which considers the corrugated plate as a nonplanar assemblage of flat plates. Buckling loads calculated using this method are compared to those obtained by using orthotropic plate theory, which represents the corrugated plate by an orthotropic flat plate. The comparison shows significant differences in predicted buckling loads.