Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

CHAPTER

MORPHING SKIN: FOAMS

Oliver Schorsch, Christof Nagel, Andreas L€uhring


7
Fraunhofer IFAM, Bremen, Germany

CHAPTER OUTLINE
1 Introduction ................................................................................................................................... 208
2 Design Principles ........................................................................................................................... 208
3 Low Temperature Elastomers .......................................................................................................... 210
4 Material Properties of HYPERFLEX ................................................................................................... 215
5 Properties of Bonded Joints ............................................................................................................ 218
6 Properties of Morphing Skin ........................................................................................................... 222
7 Skin Manufacturing ........................................................................................................................ 225
8 Summary and Conclusions .............................................................................................................. 228
References ........................................................................................................................................ 229

NOMENCLATURE
 nearly
ATED Adaptive wing Trailing Edge Device
D drag
Di temperature-dependent material parameter
J volume ratio
k material-dependent parameter
K0 initial bulk modulus
L lift
n Poisson’s ratio
N material parameter
Nf number of cycles to failure
PDMDPS polydimethyldiphenylsiloxane
PDMS polydimethylsiloxane
R displacement ratio
r.h. relative humidity
TE trailing edge
U strain energy density
UV ultra-violet (radiations)
αi temperature-dependent material parameter

Morphing Wing Technologies. https://doi.org/10.1016/B978-0-08-100964-2.00007-1


# 2018 Elsevier Ltd. All rights reserved.
207
208 CHAPTER 7 MORPHING SKIN: FOAMS

εa strain amplitude
εa0 material-dependent parameter
λi deviatoric principal stretch
μi temperature-dependent material parameter

1 INTRODUCTION
From the beginning of aircraft industry, aerodynamic morphing has been used to optimize aerodynamic
performances along the flight by changing surface area or camber. Today, entirely mechanical actu-
ation systems are established, where changes of angle/shape or surface area are obtained by moving
solid elements. Examples are the movable engine of Osprey helicopters, thrust vectoring of the Harrier,
swing wing of fighter aircrafts, as well as slats, flaps and spoilers of conventional aircraft [1].
Although widely used, moving solid elements have many disadvantages. Some solutions are too
heavy, complex, or a gapless approach is demanded for reduction of noise and fuel consumption [2].
In contrast to moving solid elements, seamless or gapless morphing is still a research topic. Gapless
morphing combines an appropriate actuation system and a skin layer which is able to deform and to
carry loads simultaneously. In the past, many actuation systems have been investigated, but up to now
there is no single material known yet that solves the mechanical paradox of morphing skins for aviation
applications satisfactorily [3].
As shown in many former morphing wing projects, the implementation of large morphing areas
leads to heavy and expensive solutions. If small shape or camber changes are requested, one possible
solution is the integration of elastomer covered adaptive elements.
The advantages of elastomer based skins are the possibility of a completely seamless approach in-
cluding outstanding compression and tension properties. On the other hand, elastomers are known for
stiffening at low temperature, low abrasion resistance, and fatigue caused by high strain.
A new morphing approach was investigated in the SARISTU project, funded by the European
Union’s Seventh Framework Programme [4–6]. The work focused on the integration of smaller poly-
mer based morphing elements to established structures like the wing trailing edge and winglet.
The SARISTU Adaptive wing Trailing Edge Device (ATED) was designed with reference to the
outer wing of a CS-25 category aircraft. Fraunhofer IFAM developed a polymer based morphing skin to
seamlessly cover a multifinger rib architecture enabling conformal and differential airfoil camber
morphing. Wing shape is controlled during flight (cruise condition) to compensate weight reduction
following fuel burning, by allowing the trimmed configuration to remain optimal in terms of efficiency
(L/D ratio) or minimal drag (D). Trailing edge adaptations were investigated to achieve significant ben-
efits in aircraft fuel consumption whose reduction may range from 3% to 5% due to the improved aero-
dynamic efficiency.

2 DESIGN PRINCIPLES
In this section it is investigated how seamless morphing can be achieved in an active trailing edge de-
vice (ATED). The basic idea here is to keep the continuity of the aerodynamic surfaces by the elastic
deformation of a solid material, which will be referred to as “morphing skin” hereafter. The morphing
2 DESIGN PRINCIPLES 209

FIG. 1
Primary structure of the ATED [7,8].

skin is to be placed onto a primary substructure, which differs from a conventional structure in that the
ribs contain hinges, such that the camber can be changed, as shown in Fig. 1.
A major challenge is to make a highly deformable, yet stable structure, which can be operated at low
( 55°C) and high (+80°C) temperatures, and which is stable under fatigue loading conditions and en-
vironmental impact. The morphing skin should fill the gap between two stiff parts of the outer aero-
dynamic surface, one of those remaining in its initial position and the other one moving with respect to
it. The morphing skin must be able to follow dimensional changes of the gap evolving between the stiff
parts but must also maintain conformity in the deflected configuration for aerodynamic reasons. The
morphing skin must withstand movements of about 5 degrees of the trailing edge and relevant air pres-
sure in operation. Furthermore, it should allow for maintenance and repair. This means that the morph-
ing skin should have a removable connection to the aircraft structure. Also, the weight of the morphing
skin must be as low as possible. In addition, it is crucial that the actuator capacity needed to deform the
morphing skin remains sufficiently low for the whole system to be economic. To obtain sufficient stiff-
ness against air pressure and—at the same time—flexibility for aerodynamic continuity and economic
operation is probably the biggest challenge within elastomer based morphing approaches [1].
In order to solve this problem, the morphing skin was segmented, alternating stiff and soft blocks.
Thus, only small, soft parts are subject to air pressure and the related deformation remains small be-
cause the support length of the soft segments is relatively short. It should, however, not be too short
because otherwise the TE movement would generate strains which are critical to the material. Primary
structure and kinematics were defined on the base of specific functional constraints, so the space left for
the thickness of the morphing skin was limited to values between 4 and 12 mm, depending on the dis-
tance from the trailing edge tip. For this reason, a significant thickness variation, which would have
been an alternative way for stiffness variation, was impossible. Hence, the design process was focused
on the length of the soft segments.
Intermediate results of the design process are shown in Fig. 2. Results of the preliminary design
phase were that out-of-plane displacements due to air pressure loading could be kept below 1 mm
if the distance of the supporting points were below 50 mm and the stiffness would be in the order
of about 150 N/mm per 100 mm span wise edge length and 10 mm skin thickness.
210 CHAPTER 7 MORPHING SKIN: FOAMS

FIG. 2
Design study of morphing skins, representing panels of 200 mm span wise (z direction) and 160 mm chord
wise (x direction) edge length and 10 mm thickness. Alternating soft and hard segments are differently shaded.
The top layer, perpendicular to the y-direction, is shown in dark.Direction of strain load: x, direction of air
pressure load: y.

A joining concept was needed in order to assemble soft and stiff segments and the top layer. Be-
cause of the soft materials involved, and the material mix with stiffness variations over several orders in
magnitude, a laminating process was chosen. This is an effective technology to avoid concentrated
forces, which would immediately cause failure in soft materials. Because it was also required at a later
stage to take off the skin for maintenance of actuators, hinge mechanics, instrumentation, etc., bolts
were foreseen in areas where stiff skin segments go onto the ribs of the primary substructure shown
in Fig. 1.

3 LOW TEMPERATURE ELASTOMERS


The core material of elastomer based morphing skins has to meet several important requirements. Most
importantly, almost constant compression and tension properties are needed as well as constant storage
modulus over the whole temperature range of an aircraft wing surface (55°C to 80°C) and over a very
large number of temperature cycles.
Moreover, resistance to abrasion (sand, small stones on runway), environmental resistances
(weather, UV radiation) as well as chemical resistance (Skydrol, deicing agent) are essential. The strain
within the soft skin segments is of great importance for the material selection as well. It is given by
geometry and deflection angle of the morphing trailing edge. Maximum strain should be kept below
10%, otherwise the elastomer would fail due to fatigue after several loading cycles. At least
100,000 cycles are expected to be sufficient for a full-service time of a common aircraft. Finally, high
adhesion strength between elastomer skin layers, elastomer foam, and aluminum parts are required to
load transmissions between hard and soft segments.
There are several certified elastomers available for aircraft applications. Examples are silicone,
EPDM, and polysulfide. All these elastomers are used as sealants, but not for high strain. The main
reason is the loss of elasticity as soon as environmental temperatures falls below the glass transition
temperature of the elastomer (usually between 45°C and 10°C, Fig. 3).
3 LOW TEMPERATURE ELASTOMERS 211

10,000

Storage modulus (MPa) 1000

100

10

1
–70 –50 –30 –10 10 30 50 70 90
Temperature (°C)
Universal V4.1D TA Instruments
FIG. 3
€ller, circle) and low temperature polyurethane Sikaflex
Dynamic mechanical analysis of EPDM sheeting (Dr. D. Mu
553 (Sika, square) [9].

Silicone elastomers are known to provide excellent elasticity over a wide temperature range.
Among the variety of silicone elastomers, polydimethylsiloxane (PDMS) is the most used for technical
applications. Fluorosilicone is based on polytrifluoropropyl-methylsiloxane and is used for applica-
tions that require fuel or hydrocarbon resistance.
Although glass transition temperatures of these silicones are well below 100°C, common PDMS
and fluorinated silicone lose flexibility below 35°C due to cold crystallization of polymer segments.
Crystallization can be reduced or even inhibited by substitution of functional groups at polymer side
chains. Polydimethyldiphenylsiloxane (PDMDPS) is the most important crystallization hindered sili-
cone. Because of almost constant elasticity between 100°C and 100°C, phenylated silicone is an im-
portant encapsulation material for space applications (Fig. 4).
Several commercial low temperature silicones are available, mainly used for the space industry or
other low temperature applications. Most of low temperature silicones are either peroxided or platinum
catalyzed products. Up to now no commercially available low temperature silicone sheets and foams
meet all the material properties for a morphing skin. Furthermore, silicones are known to be very dif-
ficult to bond to adhesives.
Reactive formulations can be tailored to improve the needed material properties and to maximize
the adhesion strength between all material interfaces of the morphing skin.
Examples for low temperature platinum curing encapsulants are Momentive RTV655 and Dow
Corning 3–6121. After curing the two-part formulations, the resulting elastomers show almost constant
storage modulus between 60°C and 100°C (see Fig. 5).
212 CHAPTER 7 MORPHING SKIN: FOAMS

1000

Storage modulus (MPa)


CH3 CH3 CH3
100
H2C CH Si O Si O Si O Si CH CH2
CH3 CH3 m CH3
n

10

1
–70 –50 –30 –10 0 10 20 30 40 50 60 70 80 90
Temperature (°C) Universal V4.1D TA Instruments

FIG. 4
Dynamic mechanical analysis of common silicone sheeting (MVQ Silicones GmbH, star, n ¼ 0) and low
temperature silicones with different methyl/phenyl ratio (m/n), Sylgard 184 (Dow Corning, square), Elastosil S690
(Wacker, circle) [9].

10

8
Storage modulus (MPa)

0
–60 –40 –20 0 10 20 30 40 50 60 70 80 90 100
Temperature (°C)
Universal V4.1D TA Instruments
FIG. 5
Dynamic mechanical analysis of two commercial low temperature platinum curing encapsulants, showing
almost constant storage modulus between 60°C and 100°C. Dow Corning 3–6121 (square) and Momentive
RTV655 (circle).
3 LOW TEMPERATURE ELASTOMERS 213

30
Dow Corning 3-6121
Momentive 655
25

Tear strength (N/mm) 20

15

10

0
0 5 10 15 20 25 30
Evonik Aerosil R 8200 (%)
FIG. 6
Effect of fumed silica on tear strength of different low temperature silicones [9].

Unfilled silicone elastomer has poor mechanical properties. Tensile strength, elongation at break,
and tear strength can be improved by filler materials. In opposite to pure encapsulant, Momentive 655
Dow Corning 3-6121 already contains an unknown amount of fumed silica.
The effect of fumed silica on tear strength and tensile properties is shown in Figs. 6 and 7. Different
levels of Evonik AEROSIL R 8200 were added to Momentive 655 and Dow Corning 3-6121. All for-
mulations were vacuum dispersed by Planetary Centrifugal Vacuum Mixer and cured 1 h at 100°C.
A tear strength test was carried out with DIN 53515 using Graves angle test piece. A tensile test was
carried out with DIN 53504 to measure tensile strength and elongation at break.
Mechanical properties can already be improved by low filler concentrations. But >15% fumed sil-
ica is needed to increase the resistance to tearing (“locking effect”). On the other hand, additional filler
material always increases formulation viscosity that makes processing more difficult or even impos-
sible. Special surface treated fumed silica is available for silicone rubber to reduce this effect.
Due to crosslinking reaction, the formulation viscosity increases after mixing part A and B. Fig. 8
shows the viscosity changes of different encapsulants. The time to double viscosity is defined as pot
life. Depending on how to process the liquid formulation, the pot life can be increased by the addition of
inhibitors.
Silicone foams are usually manufactured with special foaming equipment. Fig. 9 shows an example
of silicone foam, which was formed in an extrusion process based on peroxide crosslinking.
Liquid formulations of reactive silicone are difficult to foam without special foaming equipment.
One option is the addition of physical or chemical foaming agents. While chemical foaming agents are
thermal reactive components, physical foaming agents consist of encapsulated low-boiling solvents.
Both foaming agents release gas after reaching a critical temperature. Those reactive formulations must
214 CHAPTER 7 MORPHING SKIN: FOAMS

8 Dow Corning 3-6121


Momentive 655
7

Tensile strength (N/mm2)


6

2
0 5 10 15 20 25 30
Evonik Aerosil R 8200 (%)
FIG. 7
Effect of fumed silica on tensile strength and elongation at break of different low temperature silicones.

110
100
90
80
Viscosity (Pa s)

70
60
50
40
30
20
Dow Corning 3-6121
10
Momentive RTV 655 + 20% R8200
0
0 50 100 150 200 250 300
Time (min)
FIG. 8
Rheological measurement of reactive silicone encapsulants (viscosimeter, 4 cm plate 10 1/s @ 23°C, shear time
150–300 s).
4 MATERIAL PROPERTIES OF HYPERFLEX 215

FIG. 9
Example of BIW silicone foam extrusion with silicone sheeting on top.

be tailored specifically, since the curing reaction and foaming reaction occur at the same time. The
speed of crosslinking reactions can be adjusted by adding inhibitors or crosslinking catalysts.

4 MATERIAL PROPERTIES OF HYPERFLEX


A flexible HYPERFLEX-02 foamed elastomer was formulated specifically for application in morphing
devices. Quasistatic tensile and compression tests were performed at 55°C, 23°C, and +80°C at a
nominal strain rate of approximately 0.1/s (ISO 527-2 and ISO 604, [10,11]). A total of six specimens
were tested under each condition. The material parameters derived from the tests are summarized in
Table 1 and the stress-strain data are shown in Fig. 16. Also, the material properties of HYPERFLEX-
03 homogeneous, flexible elastomer from [9] were used. Results are summarized in Table 2. It can be
seen that both materials behave like soft elastomers over the whole temperature range. No significant
stiffening effect is visible. The observed increase in strain at fracture with decreasing temperature is
useful because most of the operation cycles will occur in the flight phase, which is at low temperature.
Most important results from compression tests were low and temperature insensitive initial elastic
moduli and the absence of failure until 70% or 80% strain.
Poisson’s ratio (ν) of HYPERFLEX-02, as determined by video analysis, was found to be close to
0.2, which is significantly lower as compared to HYPERFLEX-03, but nonzero. This means that there
is a transverse contraction for HYPERFLEX-02 parts if loaded in uniaxial tension, but loading of a
typical soft segment is biaxial and transverse strain will consequently be even lower. The value mea-
sured for HYPERFLEX-03 at +80°C is given in parentheses since it is larger than 0.5. This effect is due
to the relatively high value of ν for this material and the sensitivity of the method. A physically mean-
ingful value of 0.5 is obtained if two standard deviations are subtracted from the mean value. The tear
strength was measured in accordance with ISO 34 [12] using angle test pieces at a test speed of
216
CHAPTER 7 MORPHING SKIN: FOAMS
Table 1 Results of Quasistatic Tensile Tests on HYPERFLEX-02 at a Strain Rate of 0.1/s and Indicated Temperatures
Initial Elastic Modulus Strain at Tensile Strength Tear Strength
Condition (°C) (MPa) Fracture (%) (N/mm2) (N/mm) Poisson’s Ratio

55 3.63  0.28 94  9 1.53  0.07 8.74  0.94 0.14  0.04


23 2.17  0.04 81  7 1.17  0.08 3.91  0.26 0.22  0.00
80 3.60  0.55 59  6 0.67  0.10 3.00  0.16 0.22  0.02
4 MATERIAL PROPERTIES OF HYPERFLEX
Table 2 Results of Quasistatic Tensile Tests on HYPERFLEX-03 at a Strain Rate of 0.1/s and Indicated Temperatures
Initial Elastic Strain at Tensile Strength Tear Strength (N/
Condition (°C) Modulus (MPa) Fracture (%) (N/mm2) mm) Poisson’s Ratio

55 2.33  0.09 309  66 10.08  1.78 28.94  4.71 0.46  0.03
23 1.96  0.05 173  11 5.49  0.38 12.90  2.70 0.47  0.03
80 1.90  0.06 140  10 4.88  0.39 4.93  2.00 (0.59  0.05)

217
218 CHAPTER 7 MORPHING SKIN: FOAMS

500 mm/min. Results given in Table 2 show that there is significant tear strength but in comparison
to typical sealing elastomers, the measured values tend to be low.

5 PROPERTIES OF BONDED JOINTS


T-joints consisting of aluminum parts bonded with HYPERFLEX02 and coated with HYPERFLEX03
were tested under quasistatic conditions at 55°C, 23°C, and +80°C. The T-joint represents a piece of
the finalized skin module (see Fig. 17). It can be loaded in-plane in tension or compression, which
represents stretching or compression of the skin module induced by trailing edge movements. Shape
and dimensions of the T-joint are shown in Fig. 10. The purpose of the current tests was to confirm
stability of adhesion among elastomer and aluminum parts and to provide experimental data for finite
element calculations and sizing.
A crosshead speed of 0.03 mm/s was chosen, which corresponds to a nominal strain rate of 0.001/s.
All samples were tested until visual damage, which was coincident with a sudden load drop. The
recorded load-displacement data are shown until the maximum load in Fig. 11. From these data,
the nominal strength, the displacement at maximum load, and the initial stiffness were calculated.
The nominal strength was calculated as the maximum load divided by the bonding area. The initial
stiffness was defined as the slope of the load-displacement curve in the range from zero to 10% nominal
strain. The nominal strain was defined as the measured displacement divided by the initial length.
Test results are summarized in Table 3, showing similar variations as uniaxial test data but lower
nominal values. Reasons for the low nominal values are the constraint effect imposed onto the foam
layer by the relatively stiff aluminum parts, and locally weak adhesion as indicated by the fracture pat-
tern, Fig. 12.
T-joints were tested under cyclic loading at 23°C using harmonic load-time functions and displace-
ment control. A displacement ratio of R ¼  1 was chosen to simulate upward and downward TE
movements. T-joints were clamped such that a clearance of 1 mm was obtained between grip and short
limb of the aluminum L-profile on either side. A value of 50% stiffness change was used as a failure
condition because displacement control hardly produces total fracture.
Results are shown in Fig. 13 using nominal strain amplitudes, which were calculated as the mea-
sured deflection divided by the initial length of the sample (Fig. 10). Statistical data analysis was

30
10
2
1
35

Foam
80 Aluminium profile
Top layer

FIG. 10
T-joint.
5 PROPERTIES OF BONDED JOINTS 219

500

400

Load (N) 300

200

23°C
100 –55°C
80°C

0
0 2 4 6 8 10 12 14
Displacement (mm)
FIG. 11
Load-displacement curves from quasistatic tests of T-joints, tested under quasistatic loading with a loading rate of
0.03 mm/s. Test temperatures are indicated.

Table 3 Results of Quasistatic Tests of T-Joints Consisting of Aluminum Parts and


HYPERFLEX, Tested Under Quasistatic Loading With a Nominal Strain Rate of 0.001/s
Condition Nominal Strength Displacement at Initial Stiffness
(°C) Fracture Load (N) (MPa) Fracture (mm) (N/mm)

55 358.6  23.9 1.09  0.10 12.23  0.97 70.77  6.43


23 219.1  16.4 0.64  0.04 10.64  0.94 69.47  5.56
80 138.5  11.1 0.40  0.02 7.98  1.12 52.76  5.27
Test temperatures are indicated.

performed in order to estimate the number of cycles to failure at given strain amplitude with statistical
significance. A power law has been used to represent the data in Fig. 13 in the form

εa ¼ εa0  Nf 1=k (1)

where εa denotes the strain amplitude, Nf denotes the number of cycles to failure, and εa0 and k are
material-dependent parameters. Since εa acts as controlled variable and Nf as dependent variable,
Eq. (1) was resolved for Nf and subsequently least-squares fitted to measured Nf data as function of
applied εa0, assuming a lognormal distribution of measured Nf.
Strain-life data measured at 55°C and +80°C tend to fall below the results at 23°C, specifically at
higher amplitudes. Only 55°C and 23°C test data were considered in the fit because no significant
contribution to damage was to be expected at 80°C, as effectively, no fatigue loading occurred at that
220 CHAPTER 7 MORPHING SKIN: FOAMS

FIG. 12
Fracture surfaces of T-joints consisting of aluminum parts, HYPERFLEX-02, and HYPERFLEX-03, tested under
quasistatic loading with a nominal strain rate of 0.06/s.

100
Strain amplitude (%)

10
23°C
80°C
–55°C
E-N curve
Prediction limit

1
102 103 104 105 106 107
Cycles to failure
FIG. 13
Strain-life data from a displacement controlled, fully reversed fatigue test of T-joints, consisting of aluminum parts,
HYPERFLEX-02 foam and HYPERFLEX03 top layer, at given temperatures and a test frequency of 5 Hz. Dashed
and full lines represent the limits for a probability of 50% survival and 90% survival, respectively.

range. The fit curve is shown in Fig. 13 as a dashed line, giving a reasonable representation of the test
data. The values of the corresponding fit parameters are εa0 ¼ 102.160.13% and k ¼ (5.32  0.57).
Based on the standard deviation of the regression and a t-distribution for log(Nf), the lower prediction
limit of the E-N curve was calculated for a chosen survival probability of P ¼ 90% and n  2 degrees of
5 PROPERTIES OF BONDED JOINTS 221

freedom using a one-sided tolerance limit following the procedure outlined in ISO 12107 [13]. The
lower limit is shown as a full line in Fig. 10. It can be seen that the fatigue life for a chosen probability
of survival of 90% is reduced by about one order in magnitude as compared to the dashed E-N curve,
which corresponds to a probability of survival of 50%.
Visual inspection of the fracture surfaces after completion of the fatigue test indicated that the joints
failed predominantly by cohesive fracture within the elastomer material, similarly to Fig. 12. This
means that the adhesion between elastomer and aluminum surface is stable against fatigue loads at
55°C, 23°C, and +80°C.
T-joints as described in the preceding paragraphs were exposed to cyclic laboratory aging condi-
tions followed by quasistatic testing. One set of samples was exposed according to simulated “exterior,
land, and air” environment (method C in ASTM D 1183-03 [14]), one cycle consisting of 48 h at
(71  3)°C and <10% relative humidity (r.h.), 48 h at (23  1)°C immersed in water, 8 h at
(57  3)°C and 100% r.h., 64 h at (38.5  2)°C and 100% r.h. A furnace and climate chambers
equipped with calibrated temperature and humidity sensors were used. A second set of samples was
prepared by immersion in SKYDROL 500-B-4 at room temperature for a certain number of cycles,
one cycle equal to 168 h. After completion of one cycle, two samples were taken and tested immedi-
ately while the other samples were exposed to the next cycle. This was repeated until the last two sam-
ples had passed the last cycle. Test specifications and data evaluation were exactly as described in the
section on quasistatic testing.
An overview on the joint properties after exposure to laboratory aging conditions to method
C (“exterior, land, and air”) in ASTM D 1183-03 is given in Fig. 14. Error bars refer to one standard
deviation and full lines represent the mean value over all cycles. Neither strength, nor displacement at
maximum load, nor the initial stiffness show systematic variation with the exposure time, indicating
that the material was not affected by the environmental attack. Visual inspection of the fracture surfaces
indicated that from cycles 1 to 4, joints failed predominantly by mixed cohesive fracture within the
elastomer and adhesive failure at the foam-aluminum interface. From cycles 5 to 9, cohesive fracture
within the elastomer was the predominant mechanism.

2.50 30.00 100.00


Displacement at max. load (mm)

90.00
25.00
Nominal strength (MPa)

2.00 80.00
Initial stiffness (N/mm)

70.00
20.00
1.50 60.00
15.00 50.00
1.00 40.00
10.00
30.00
0.50 20.00
5.00
10.00
0.00 0.00 0.00
0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10
Number of cycles Number of cycles Number of cycles

–55°C 23°C 80°C –55°C 23°C 80°C –55°C 23°C 80°C

FIG. 14
Mechanical properties of T-joints tested under quasistatic conditions as function of the number of exposure cycles
to ASTM D 1183–03C, “exterior, land, and air”. Temperatures refer to the test conditions.
222 CHAPTER 7 MORPHING SKIN: FOAMS

1.40 20.00 90.0

Displacement at max. load (mm)


18.00 80.0
1.20
Nominal strength (MPa)

Initial stiffness (N/mm)


16.00 70.0
1.00 14.00
60.0
0.80 12.00
50.0
10.00
0.60 40.0
8.00
30.0
0.40 6.00
4.00 20.0
0.20
2.00 10.0
0.00 0.00 0.0
0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10
Number of cycles Number of cycles Number of cycles

–55°C 23°C 80°C –55°C 23°C 80°C –55°C 23°C 80°C

FIG. 15
Mechanical properties of T-joints tested under quasistatic conditions as function of the number of exposure cycles
in SKYDROL at room temperature. Temperatures refer to the test conditions.

T-joint test results after immersion in SKYDROL 500-B-4 at room temperature are shown in
Fig. 15. The error bars refer to one standard deviation and full lines represent the mean value over
all cycles. It can be seen that after an initial drop, strength, displacement, and initial stiffness show
no changes with increasing cycle number. This means that the joint properties are initially rapidly chan-
ged by immersion in SKYDROL, but stabilize after a couple of cycles. Visual inspection of the fracture
surfaces indicated that the joints failed predominantly by cohesive fracture within the elastomer.
A small adhesive failure fraction at the foam-aluminum interface was also visible, but it was signif-
icantly smaller than the adhesive failure fraction of the T-joints exposed to simulated “exterior, land,
and air” conditions.

6 PROPERTIES OF MORPHING SKIN


It should be noted at this point that the stiffness of elastomers decreases with increasing number of load-
displacement cycles. It is therefore not adequate for such materials to derive a material model based on
the first loading cycle of a quasistatic tensile test. Load-displacement cycles of a T-joint are shown in
Fig. 16, where the displacement was increased in each loading cycle. It can be noticed from Fig. 16 that
the initial stiffness decreases with increasing number of load cycles, but the load-displacement curve in
the current cycle approaches that of the preceding cycle as soon as the maximum imposed to the ma-
terial in the preceding cycle is exceeded. This is commonly known as Mullins effect [15,16].
An Ogden type material model [17] was used in the form
X
N
2μi  αi αi αi  X N
1  el 2i
U¼ λ1 + λ2 + λ3  3 + J 1 (2)
i¼1
αi 2 D
i¼1 i

with
λi ¼ J 1=3 λi (3)
6 PROPERTIES OF MORPHING SKIN 223

300 1.6 30

Initial shear modulus µ (MPa)


1.4
200 m 25

Stretch exponent a
a
1.2
100 20
Load (N)

1.0
0 15
0.8

–100 10
0.6

–200 0.4 5
–20 –15 –10 –5 0 5 10 15 20 0.00 0.05 0.10 0.15 0.20
Displacement (mm) Nominal strain
FIG. 16
Full hysteresis of T-joints, measured with increasing strain amplitudes (left), parameters of hyperelastic Ogden
potential as function of nominal strain for HYPERFLEX-02 (right).

to describe the material stiffness as a function of the prestrain. Herein, λi refers to the deviatoric prin-
cipal stretches, λi refers to the principal stretches, and J refers to the volume ratio. In Eq. (3), N is a
material parameter, and αi, μi, and Di are temperature-dependent material parameters. The initial shear
modulus and bulk modulus are given by μ0 ¼ Σ μi and K0 ¼ 2/D1. The initial bulk modulus was esti-
mated from the measured Poisson’s ratio. For the current material, it was sufficient to use a single term
approach of Eq. (2). The values obtained for α and μ are shown in Fig. 16.
The initial shear modulus and the stretch exponent are depicted in Fig. 16 as a function of the nom-
inal strain. A two-parameter power law

y ¼ a  εb (4)

was fitted to the relation between initial shear modulus and nominal strain, yielding a ¼ 0.24 0.01
and b ¼  0.55  0.02. The relation between exponent alpha and nominal strain was also fitted by
Eq. (4), leading to a ¼ 4.02  0.41 and b ¼  0.58 0.04. For nominal strains between 2% and 7%
appearing in the soft segments of the morphing skin, this means that adequate values of μ and α are
between 1 and 2 MPa and between 19 and 39, respectively.
A model of the finalized cross section of the upper and lower skins is shown in Fig. 17. The soft
segments are on top and below the hinge lines, which are the regions where the highest deformations
are generated. Between the hinge lines, there is almost no deformation generated in the skin due to the
relatively high stiffness of the rib segments. Consequently, there was no flexibility needed and stiff
segments were placed in those regions. Boundary conditions of the final design were a specified
amount of rotation (typically two degrees around hinge lines) and a linearized pressure distribution,
pressure values typically ranging from 10,000 to 6000 Pa at the upper skin and from 8000 to
6000 Pa at the lower skin (chord-wise).
224 CHAPTER 7 MORPHING SKIN: FOAMS

FIG. 17
Finalized geometry of upper (top) and lower (bottom) morphing skins with actuation mechanism (middle).
Morphing skins consist of soft segments (dark shaded), stiff segments (bright shaded), and top layer (medium
shaded).

The thickness of the soft segments was between 4 and 12 mm and the chordwise length was between
25 and 30 mm, depending on thickness. The variations are due to different amounts of space locally
available on the primary structure and due to the variation of the pressure as a function of the chordwise
position. The average value of the maximum principal strain at maximum trailing edge displacement
and applied pressure load was typically 2% and did not exceed 7%. A typical strain distribution in a soft
segment is shown in Fig. 18. The maximum out-of-plane displacements were below 0.5 mm.
The number of actuation cycles at full TE deflection was estimated as N ¼ 20,000. Based on the
maximum value of the maximum principal strain of 7% and the lower limit of the E-N curve in Fig. 13
one obtains cycle numbers of N ¼ {1E6, 3E5, 6E4} if the probability of failure Pfail is set to {10%, 1%,
0.1%}. If the average maximum principal strain of 2% is considered, which is reasonable because the
T-joint contains, due geometric similarity, the same local strain peaks like the morphing skin, one ob-
tains N ¼ 7E4 even if Pfail is set to a value as low as 1E-6. It can hence be seen that the probability of
skin failure due to TE actuation is very low.
The force needed to deform upper and lower skins of the model in the position of maximum dis-
placement at applied pressure was typically 2 kN per 1000 mm span width. This means that only a
small part of the actuator load capacity would be necessary to deform the morphing skins in operation.

FIG. 18
Strain distribution in a soft segment of the morphing skin.
7 SKIN MANUFACTURING 225

It is often discussed that morphing materials should have a Poisson’s ratio of zero. In the current
design, the Poisson’s ratio of the foamed Hyperflex-02 material is close to 0.2 which in fact means that
the transversal strain is about 0.4% at the average chord-wise strain level of 2%. Because the skin is in a
state of planar strain, transversal strain can only be effective in the direction perpendicular to the skin,
and the related out-of-plane deformation can be estimated as 50 μm in regions where the skin thickness
is 12 mm, which is in fact very low. Out-of-plane deformations found in the model were larger but this
effect was related to bending as well as to suction and compression due to air pressure. This indicates
that a Poisson’s ratio of zero is not so important for the current design.

7 SKIN MANUFACTURING
The design of the SARISTU ATED morphing skin consists of the three main components. The space
between the aluminum profiles is filled with low temperature foam. The foam region as well as the
aluminum profiles are covered with a 1 mm thin protective layer of a low temperature silicone elas-
tomer (Fig. 19).
The soft segments are located above and under the hinges of the rib architecture and are responsible
for skin movement. The lower and upper skin each have three foam segments (Fig. 20).
A new manufacturing process was developed due to the unavailability of appropriate adhesives to
bond these materials for low temperature purposes. A core feature of this process is to perform curing,
foaming, and adhesive bonding of different reactive silicone formulations and aluminum profiles at the
same time. Chemical crosslinking during the curing reaction lead to a strong connection between the
foam and the elastomer top layer. To increase the adhesion strength between the silicone and metal
interface, the metal profiles were degreased and primered with an adhesion promoter for platinum cat-
alyzed silicones.
In the first step, the liquid reactive rubber formulation is poured into a cast to obtain a 1 mm thin
protective layer after curing (Fig. 21).

FIG. 19
Aluminum profiles, low temperature silicone foam and a protective layer of low temperature silicone elastomer are
the main skin components of the SARISTU ATED.
226 CHAPTER 7 MORPHING SKIN: FOAMS

FIG. 20
The morphing skin consists of hard and soft segments. Elastomer foam is used for soft segments which are
located above and under the rib hinges. Hard segments are aluminum profiles (reversed U profiles, top and
bottom boundaries). Hard and soft segments are covered with a thin elastomer layer. The rib kinematic is shown
simplified (the two hinged rods, at the middle of the picture).

FIG. 21
A liquid rubber formulation is poured into a heatable cast. After curing the rubber leads to a 1 mm thin
protective layer.

In the next step, the aluminum profiles are positioned on the still liquid rubber layer. A vacuum
process is used to prevent bubbles between the protective rubber layer and the aluminum profiles
(Fig. 22).
A reactive rubber formulation containing a foaming agent was filled between the aluminum profiles
leading to the soft segments of the morphing skin after curing (Fig. 23).
Fig. 24 shows a sketch of a heating device to manufacture morphing skins up to 2.3 m length. The
heating device consists of several heating pads which are separately controlled to obtain a homogenous
temperature distribution and can reach temperatures up to 150°C.
After the heating process, the foam segments are cut with an oscillating knife to obtain the required
foam thickness (Fig. 25).
FIG. 22
Vacuum bag to prevent bubbles between the protective rubber layer and the aluminum profiles.

FIG. 23
The liquid reactive rubber formulation contains a foaming agent leading to the soft segment between the
aluminum profiles after curing.

FIG. 24
2.8 m large heating device to process curing, foaming, and adhesive bonding of morphing skin components at the
same time [9].
228 CHAPTER 7 MORPHING SKIN: FOAMS

FIG. 25
Cutting the rubber foam segments with an oscillating knife.

FIG. 26
Upper and lower skin sides are connected to stringers located between the ribs.

The obtained upper and lower skin side was attached to stringers located between the active ribs
(Fig. 26) to transmit load in a spanwise direction and to assure attachment/detachment for ease of main-
tenance and repair.
During the SARISTU project several large skin panels were manufactured and successfully assem-
bled into a true-scale wind tunnel demonstrator (see Fig. 27).

8 SUMMARY AND CONCLUSIONS


In this chapter, a morphing skin concept was investigated and manufactured for an active trailing edge
device of an adaptive wing demonstrator. In the current design, geometric gap changes are converted
into mechanical strain to maintain a smooth transition between connected parts. The proposed solution
was developed up to the physical subcomponent level. The manufacturing technology was demon-
strated and adaptive parts were delivered for demonstrator assembly, ground tests, and wind
tunnel tests.
REFERENCES 229

FIG. 27
Lower side morphing skins for the SARISTU adaptive trailing edge.

The stiffness measured in cyclic tests on bonded joints was accurately predicted by a hyperelastic
material model. It was shown that the elastomeric skin materials have excellent fatigue and aging re-
sistance as well as stable adhesion on aluminum surfaces within aircraft operational temperature range.
The presented results suggest the elastic deformation approach based on elastomeric multimaterial de-
sign is a potential solution for the morphing paradox to deform and carry loads simultaneously.
The research leading to these results has gratefully received funding from the European Union’s
Seventh Framework Programme for research, technological development, and demonstration under
grant agreement no. 284562.

REFERENCES
[1] C. Thill, J. Etches, I. Bond, K. Potter, P. Weaver, Morphing skins, Aeronaut. J. (March) (2008). Paper no.
3216.
[2] P.K.C. Rudolph, High-Lift Systems on Commercial Subsonic Airliners, NASA Ames Research Center,
Moffet Field, CA, 1996.
[3] R. Pecora, F. Amoroso, M. Magnifico, Toward the bi-modal camber morphing of large aircraft wing flaps: the
CleanSky experience, in: Proc. SPIE—The International Society for Optical Engineering, Volume 9801,
2016, Article No. 980106.
[4] 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 Engineer-
ing, Volume 9801, 2016, Article No. 980107.
[5] 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, CA, USA, Jan. 4–8, 2016.
[6] I. Dimino, A. Concilio, M. Schueller, A. Gratias, An adaptive control system for wing TE shape control,
in: Proc. SPIE International Conference on Smart Structures 2013, San Diego, CA, USA, Mar. 10–14, 2013.
[7] Pecora, R., Concilio, A.,Dimino, I., Amoroso, F., Ciminello, M., “Structural design of an adaptive wing trail-
ing edge for enhanced cruise performance”, 24th AIAA/AHS Adaptive Structures Conference, 2016; San
Diego; United States; 4 January 2016-8 January 2016.
230 CHAPTER 7 MORPHING SKIN: FOAMS

[8] 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, Article No. 980108, Apr. 16, 2016. https://doi.org/10.1117/12.2219223.
[9] C. Nagel, A. Fiedler, O. Schorsch, A. L€uhring, in: Design, manufacture, and testing of a seamless morphing
concept for a smart aircraft wingtip, Proceedings of the 7th ECCOMAS Thematic Conference on Smart
Structures and Materials (SMART 2015), Ponta Delgada, Azores, 2015.
[10] ISO 527, Plastics—Determination of Tensile Properties, International Organizationfor Standardization, ISO,
Geneva, 2012.
[11] ISO 604, Plastics—Determination of Compressive Properties, International Organization for Standardiza-
tion, ISO, Geneva, 2002.
[12] ISO 34, Rubber, Vulcanized or Thermoplastic—Determination of Tear Strength—Part 1: Trouser, Angle and
Crescent Test Pieces, International Organization for Standardization, ISO, Geneva, 2015.
[13] ISO 12107, Metallic Materials—Fatigue Testing— Statistical Planning and Analysis of Data, International
Organization for Standardization, ISO, Geneva, 2012.
[14] ASTM D1183–03, Standard Practices for Resistance of Adhesives to Cyclic Laboratory Aging Conditions,
ASTM International, West Conshohocken, PA, 2011.
[15] L. Mullins, Softening of rubber by deformation, Rubber Chem. Technol. 42 (1969) 339.
[16] R.W. Ogden, D.G. Roxburgh, A pseudo–elastic model for the Mullins effect in filled rubber, Proc. R. Soc.
Lond. Series A: Math. Phys. Eng. Sci. 455 (1999) 2861–2877.
[17] R.W. Ogden, Recent advances in the phenomenological theory of rubber elasticity, Rubber Chem. Technol.
59 (3) (1986) 361–383.

You might also like