You are on page 1of 15

48th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference<br>15th AIAA 2007-1734

23 - 26 April 2007, Honolulu, Hawaii

Aeroelastic Modeling, Analysis and Testing of a Morphing


Wing Structure

Gerald R. Andersen1 and David L. Cowan2


NextGen Aeronautics, Inc., Torrance, California, 90505

David J. Piatak 3
NASA Langley Research Center, Hampton, Virginia, 23681

NextGen Aeronautics, Inc., of Torrance, CA, has been involved in all aspects of the
design, development and testing of morphing aircraft structures, much of this in conjunction
with NASA Langley Research Center. The single most substantial effort in this area has
been the DARPA Next-Generation Morphing Aircraft Structures project, in which NextGen
has been one of two prime contractors tasked with advancing the state of technology of
morphing aircraft. NextGen developed an in-plane morphing geometry concept with
mechanized four-bar linkages, and motion imparted by computer-controlled linear
actuators. Flexible elastomeric skins with out-of-plane stiffeners accommodated the wing
motion while transmitting air pressure loads to the wing substructure. The development
process involved wind tunnel testing of a full-scale wing for a 2400-lb vehicle and flight
testing of a subscale unmanned air vehicle. Among the more critical challenges has been that
of addressing wing design for aeroelasticity, given the unique features of morphing wings.
Specific relevant issues include: the need to address multiple geometries and flight envelopes
to account for morphing shape changes; in-plane flexibility of the wing resulting from its
mechanisms and soft (relative to typical fixed wing structure) restraint by linear actuators;
nonlinearity and equivalent loss of stiffness in the wing’s out-of-plane direction as a result of
multiple revolute joints in the structure. These unfamiliar features required, in some cases,
the development of new analytical or testing approaches. In addition, due to a somewhat
untried approach to mounting the half-span wind tunnel model of its size at the NASA
Transonic Dynamics Tunnel, particular considerations of support structure dynamics had to
be given particular attention prior to wind-on testing there. Ultimately, all of these
challenges were overcome, and successful wind tunnel and flight demonstrations were
completed.

1
Senior Engineer, Design and Analysis Group, 2780 Skypark Drive, Ste. 490, Torrance, CA, AIAA Member
2
Principal Engineer, Design and Analysis Group, 2780 Skypark Drive, Ste. 400, Torrance, CA, AIAA Member
3
Senior Engineer, Design and Analysis Group, 2780 Skypark Drive, Ste. 490, Torrance, CA, AIAA Member

1
American Institute of Aeronautics and Astronautics

Copyright © 2007 by 07 by NextGen Aeronautics, Inc. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
I. Introduction

D URING the past four years, NextGen Aeronautics, Inc. has been one of the prime contractors in the DARPA-
sponsored Morphing Aircraft Structures (MAS) program. This project seeks to enhance and extend the mission
capability of aircraft through the application of reconfigurable “morphing” structures that enable an aircraft to alter
its configuration in flight to most effectively operate in a given flight regime. NextGen’s morphing concept utilized
two morphing degrees of freedom, a variable wing sweep and variable shear angle of an internal 4-bar linkage
assembly, to accommodate wing geometry changes that met the DARPA MAS programs requirements while
addressing NextGen’s targeted mission profiles. Specifically, the variable geometry allowed the wing to achieve and
maintain, in flight, five distinct configurations – one for each of five identified mission segments: high lift, climb,
loiter, cruise/dash, and high-speed maneuver.
The utilization of a reconfigurable wing structure inherently led to unique issues in the aerodynamic and
structural analysis tasks leading to the November 2005 wind tunnel test at the NASA Langley Transonic Dynamics
Tunnel (TDT). In particular, the analytic flutter clearance procedures were complicated by the large number of
configurations to be analyzed. Additionally, the structural design implementation used to achieve the morphing
capabilities led to unique aeroelastic interactions which needed to be considered during the flutter clearance effort.
The combination of a large number of physical configurations and a flexible structure with unique dynamic
characteristics resulted in a complex and unique flutter clearance investigation.
The unsteady aeroelastic analyses in support of the TDT test evolved through two major efforts. The first was the
analytic prediction of the flutter boundaries and margins for each of the configuration, the results of which were
presented at the September 2005 model system review at NASA Langley. The second major effort took place during
the test activities when updated ground vibration test data of the TDT/model interface structure became available. At
this point, real time flutter clearance runs were made, with an update flutter model, on a daily basis.
This paper describes the details of these efforts, including a summary of NextGen’s wing structural design,
identification of the unique aeroelastic phenomenon, analytic aeroelastic model generation, analysis methods, and
the flutter analysis and clearance efforts that resulted in the first successful wind tunnel test and the first ever in-
flight demonstration of a morphing aircraft wing.

II. NextGen MAS Wing Structural Model


The NextGen MAS structural model consisted of three primary components: a fairing, the wing, and a 4-bar
linkage/slider assembly, portions of which are shown in Figure 1. The fairing enclosed the 4-bar/slider assembly and
provided the wind tunnel attachment via the main weldment. A slot in the outboard fairing rib allowed the 4-bar
assembly to move as required. Note that in the figure, only the forward most portion of the fairing weldment and rib
are shown.

2
American Institute of Aeronautics and Astronautics
Fairing Slotted 1 Wing Main Pivot
Rib
Leading Edge Closeout
Fairing Main
Weldment
Spars
Spanwise-running
4-Bar/Slider
Assembly Actuators skin strip
3
Ribs

Underlying support
“ribbons”

Sliders
2 Revolute
Joints
4-Bar Assembly

Figure 1. Morphing Wing Structure Figure 2. Elastomeric Skin and Support Ribbons

The wing morphing degrees of freedom required the use of a unique internal wing structure. Wing sweep was
enabled through a large revolute joint (item 1 in the figure), located at the root of the forward-most wing spar. Wing
area change was enabled by an assembly of 4-bar linkages (item 2) with smaller revolute joints, joining ribs to spars,
allowing shear motion. To accommodate the wing root motion resulting from sweep and 4-bar shear angle changes,
a combination of an additional 4-bar linkage assembly and slider mechanism was employed (item 3).
The internal structure was not the only distinctive aspect of the wing. The external skin, having to accommodate
an exceedingly large shear strain, was made of an elastomeric silicon material. To withstand the airloads, the skin
was supported with an underlying ribbon structure, as shown in Figure 2. This provided out-of-plane stiffness while
allowing the required shearing motion.

III. Aeroelastic Model Development


Comprehensive aeroelastic models of the five configurations of interest were developed. All portions of the
model, including the finite-element model of the structure, the aerodynamic panel model, and the spline connecting
these models, were created in MSC/PATRANTM environment for use in the MSC/NASTRANTM analysis code.
Details of the model generation for each of the components follow.

A. Structural Finite Element Model Development


A detailed finite element model of the wing, main 4-bar linkage, and fairing was created. For each model, node
locations for all wing components and the main 4-bar linkage assembly were unique and had to be determined for
the respective configuration. Figure 3 shows the finite element models of the five configurations.

3
American Institute of Aeronautics and Astronautics
Figure 3. MAS Finite Element Models

1. Wing Finite Element Development


A planform view of the wing structure is illustrated in Figure 4. The aluminum wing spars and ribs were
generated with tapered beam (CBEAM) elements. Aluminum stringers, providing intermediate attachment points
between the spars for the skin panels, were modeled with constant cross section bar (CBAR) elements. The carbon
fiber leading edge and fiberglass trailing edge closeouts utilized quadrilateral (CQUAD4) elements referring to
property and material entries that accurately represent the laminate buildup. Actuators were also modeled, using
CBAR elements, in regards to location and stiffness. Actuator stiffness was estimated based upon the bulk modulus
of the hydraulic fluid in the actuators.
It is worth noting that the wing finite element model is not symmetric about the plane of the wing. The structural
elements forming the wing are fully enclosed by the outer mold line; thus the wing structure follows the camber of
the wing. Attention was paid to this detail in case any predominately in-plane dynamics resulted in out of plane
motion through this asymmetry – a result that might influence the unsteady aeroelastic behavior.

4
American Institute of Aeronautics and Astronautics
Spars
Ribs
Actuators
Leading edge closeout

Stringers Trailing edge closeout

Figure 4. Wing Structure at High Lift Configuration

2. Fuselage/Fairing Finite Element Development


The fuselage and fairing model (see Figure 5) evolved through two stages. The first model represented the main
weldment and the main wing pylon with plate and beam elements, respectively. In the first iteration, the fairing skin
stiffness was represented by rigid multi-point constraints between chordwise locations on the main weldment and
the location of the fairing center of gravity. This model was used to determine the flutter margins as presented at the
MSR in September 2007.

Leading Edge
Configuration 1
Secondary Rib

Main Weldment MainWeldment


Weldment
Main

Bulkheads
Pylon to wing
main pivot
Pylon to wing
main pivot
Slot to accommodate
4-bar linkage and
wing spar roots

Configuration 2
Main Rib

Secondary Rib

Figure 5. Fairing Structure with Skin Removed

As the TDT test approached and GVT measurements were obtained, it became apparent that a more realistic
representation of the modal behavior of the model would be obtained by creating a fully built-up model, containing
the main and internal weldments, bulkheads, and significant leading edge solid structure of the fairing substructure,

5
American Institute of Aeronautics and Astronautics
as shown in case 2. Note a highly detailed model of the fairing skin was also developed for the built-up model; it is
not however, shown for clarity. The skin model may be seen in Figure 3.

At this point it is necessary to explain the treatment of the fairing stiffness for each of these fairing models.
Configuration 1 did not employ composite fairing skin models. Rather, all fuselage structural stiffness was assumed
to be included in the main weldment, its attachment to the wing main pivot, and the contribution of the 4-bar linkage
since loads were transferred to the main weldment through the 4-bar links. Fuselage masses and moments of inertia
were lumped and added at the appropriate location to give the proper center of gravity (CG) location and system
mass and moments of inertia. This lumped mass was then rigidly connected to the weldment through multi-point
constraints. To bracket the stiffness contribution of the fairing skin, the length of rigid attachment between the CG
mass and the weldment was varied. In one case, the attachment ran from the leading edge of the fuselage to
approximately 60% of the fuselage length. A stiffer connection, attaching the CG lumped mass to the entire length
of the weldment was also employed. These two versions, referred to as case 1 and case 2, of configuration 1 led to
the results shown in Figure 9-13.
The stiffness bracketing employed for configuration 1 was not necessary for configuration 2 since the fairing
skin was accurately represented by a laminate composite built up structure. As with case 1, additional mass and
moments of inertia were added such that the total fuselage and fairing masses and moments of inertia were the
values and at the location as predicted by SolidWorksTM. This high-fidelity configuration was used in the day-to-day
determination of flutter margins during the actual testing in the TDT.

3. 4-Bar Linkage Finite Element Development


The 4-bar fuselage/wing interface structure is shown in relation to the fuselage structure in Figure 6. This
relatively simple structure was modeled with constant cross section bar elements. MPCs with z-rotation free formed
the 4-bar joints, while MPCs with x-translation free represented the guide block sliders. The connection between the
outboard channel of the 4-bar linkage and the wing spar roots was an additional complication. Because the spars
roots intersected the channel at different locations depending on the configuration, care was taken to enforce proper
node positioning to maintain the correct component geometries.

4-Bar Actuator

Cutouts for Pylon


and Sweep Actuators

4-Bar Linkage Assembly

Figure 6. 4-Bar Linkage within Fairing

4. TDT Support Structure


The final major structural component modeled consisted of the TDT interface structure – that which connected
the fuselage main weldment to the wind tunnel ground mass (see Figure 7). This portion of model underwent a
number of changes as updated stiffness data was received as a result of continuing GVT and static load tests
performed on the interface structure. Initially, based upon data provided by NASA Langley, the interface structure
was modeled by matching representing the EI and GJ distributions of the actual balance, inner and outer spindles,
and collar with appropriate boundary conditions between the collar and outer spindle (essentially roller bearings). As
stiffness data was received, the model was updated to include springs representing the stiffness of the electronic
turntable, the oscillating turntable, and the connection between the electronic turntable and the ground mass.
Because the final FEM of the interface structure is a one-dimensional beam model, it is not illustrated here.

5. Mass and Stiffness Distribution

6
American Institute of Aeronautics and Astronautics
It was assumed that the structural stiffness of the NMAS model, to include the wing, fuselage, and 4-bar
substructures, was entirely contained in relatively rigid structure – spars, ribs, stringers, composite structures,
actuators, etc. Flexible skin contributions to the wing stiffness were not included. This was considered to be
conservative as any additional stiffness from the skin would tend to raise modal frequencies contributing to flutter
mechanisms.
The mass distribution of the NMAS model was based upon mass values obtained from the components as
modeled in SolidWorksTM. For each major structural component, the mass was adjusted such that it matched the
SolidWorksTM value. Components included in this approach were the spars, ribs, leading and trailing edge closeouts,
fairing weldment, ribs, and bulkheads, and 4-bar linkage structures. Running masses such as the flexible skin and
associated fastening hardware, stringers, hydraulic lines, and instrumentation lines were lumped as appropriate by
area or length and added at each of the spar/rib joints in the wing. Concentrated masses not structurally modeled
such as lugs at revolute joints and 4-bar guide blocks, were also lumped according to their distance from a spar/rib
joint and thereby accounted for. Similarly, all mass in the fairing structure not accounted for in the structural FEM
was positioned to give the correct center of gravity location and value for the fuselage structure. Additionally, in the
case of the fuselage, moments of inertia were added as well at the center of gravity to account for those inertias not
preserved by the finite element model.

B. Aerodynamic Model Development


Aerodynamically, the wing was modeled as a flat plate lifting surface, consistent with MSC/NASTRANTM
aeroelastic analysis with four distinct aerodynamic macroelements, while a fifth macroelement comprised the
fuselage fairing. Even though the wing planform experiences relatively large changes between configurations,
because the changes are accommodated entirely by shearing motion, aerodynamic panel aspect ratios were not
greatly affected. A check of the panel aspect ratios indicated that the same paneling scheme could be used for all
five configurations, which are shown in Figure 8.

Loiter
Climb

HSM

High-Lift

Cruise/Dash

Figure 8. Aerodynamic Models for all Configurations

C. Aerodynamic/Structure Connections
Standard NASTRAN aeroelastic spline techniques were used to connect the aerodynamic model with the
structural model. In this case, the finite plate spline (FPS) was used the three inboard lifting surface macroelements
rather than the infinite plate spline (IPS). This is because the structural nodes used in the spline did not extend to the

7
American Institute of Aeronautics and Astronautics
edge of the aerodynamic model (only nodes at the intersections of spars and ribs were used, thereby omitting nodes
on the leading and trailing edge composite structures). In these situations it is generally recognized that the FPS
outperforms the IPS. The outboard most wing macroelement was connected to the wing substructure with a beam
spline since the only available structural nodes lay along a singe axis. In the case of the fairing aerodynamic model,
the aerodynamic panels were also connected to the main weldment with a beam spline. In the case of the surface
splines, spanwise smoothness of the separate splines on the wing was improved by using the convention of selecting
structural nodes slightly outside the aerodynamic macroelement used in the spline.
It should be noted that the structural nodes used in the spline generally changed among configurations since the
spanwise and chordwise motion of the panels did not match the shearing motion of the underlying structural nodes.
Thus for each configuration, the splines were checked and manually updated as required to maintain the correct
relationship between the aerodynamic macroelements and the underlying structure.

IV. MAS Aeroelastic Considerations


The nature of the structure utilized in the NextGen MAS design gives rise to a number of conditions causing the
aeroelastic behavior of this wing to differ from one of more conventional construction. The use of numerous rotating
and sliding mechanisms and a very flexible wing skin as well as the location of the wing elastic axis and center of
pressure created the potential of aeroelastic interactions different from those commonly observed.

A. In-plane Wing Dynamics


The most immediately noticeable behavior was relatively low wing in-plane stiffness. The use of revolute joints
and in the wing primary structure and sliding mechanisms in the main 4-bar assembly enabled the two morphing
degrees of freedom, but also allowed two zero frequency rigid body modes to appear in the flutter analyses.
Although these modes were stabilized through the modeling of the actuators; the frequencies of the in-plane modes
were well below those of wing of more traditional design. In fact, the lowest two modes of wing at all configurations
were in-plane, and within the first 10 modes, roughly half the modes for each configuration were in-plane. This
resulted in the possibility of modes with predominantly in-plane dynamics generating out-of-plane motion (through
the x-y asymmetric structural model) of participating in the flutter mechanisms.

B. Low Torsional Rigidity


The use of a highly flexible elastomeric skin generated the possibility that the wing would be relatively flexible
in torsion compared to bending. In fact, early in the design activity it occurred that the torsional frequencies might
be below the corresponding bending frequencies, thereby eliminating the traditional bending/torsional coupling
flutter mechanism. This turned out not to be the case; thus the possibility then arose that the onset of flutter caused
by bending/torsional coupling would occur more quickly because the torsional frequency would begin close to the
bending frequency.

C. Elastic Axis and Center of Pressure Location


The NextGen MAS model was designed to place the wing elastic axis near Spar 1 – approximately 8-12 percent
of the wing chord, depending upon the configuration – further forward than traditionally designed wings. This
elastic axis location was far enough forward to place it ahead of the center of pressure for all configurations.
Therefore, an increase in lift would result in the wing twisting downward, or unloading – a result beneficial for wing
divergence characteristics. Less intuitive, however, were the effects on the flutter behavior.

V. Flutter Analysis Technique


All flutter analyses were performed using MSC/NASTRAN Aeroelasticity I capabilities. This technique provides
a linear frequency analysis to determine the flutter eigenvalues. The non-looping PK flutter analysis was chosen as
the method used in Solution 145. The first five out-of-plane modes were retained in all analyses for all five
configurations. Initially, modal damping was assumed to be two percent, however, for the final clearance runs
during the TDT test, modal damping was assumed to be zero, thereby providing a more conservative calculation of
the flutter margin.
Using the above-mentioned method, PK flutter analyses were performed to obtain the matched-point flutter
boundary for all five configurations in the appropriate gas medium (air or R134A) as required. Flight velocities were
swept over a range of density ratios at a set Mach number to determine the non-matched-point flutter boundary for a
given Mach number. The point at which each of these curves intersected the input Mach number was determined.

8
American Institute of Aeronautics and Astronautics
The point at which the curve formed by the collection of these points intersected the desired Mach number was
noted, and from this the matched-point flutter velocity and dynamic pressure were directly obtained.

VI. Flutter Results


Figures 9 through 13 are the results of flutter analyses performed with the NMAS model using case 1 of the
fairing model. Again, two versions (Case 1 and Case 2 in the following charts) of the skin stiffness contribution are
used, leading to the dual curves shown in each plot below. The variation of flutter dynamic pressure is shown in the
left hand plots, with TDT tunnel performance boundaries for both air and R134A presented. Flutter velocity in knots
equivalent airspeed (keas) is indicated in the right hand side charts. In all charts, the flutter boundary is presented
along with the corresponding test envelope.
As appropriate, the flutter boundary is shown for both air and R134A. Both cruise/dash and HSM configurations
used test points achievable only with R134A. For example, in Error! Reference source not found.e, heavy gas
flutter margins are indicated with an open rectangle symbol. Care must be taken in interpreting these charts. If a test
point lies in the heavy gas region of the TDT performance envelop, then the flutter boundary for R135A must be
consulted. Note the heavy gas flutter margins are not shown in figure 9 because they are well off the chart in the
upward direction.
These flutter boundary results, presented at the Model System Review generated further confidence that the
model would encounter no flutter instabilities in the desired test envelop. It should be noted that these results were
obtained before the dynamics arising from the TDT interface structure were encountered.
As explained below, with updated GVT information on the TDT structure available, it became apparent that
marked differences in flutter velocities could exist between model using the original TDT interface structure models
and the updated model including the jack screw, turntables, and updated connection to ground. Because of this, it
was decided to perform additional flutter analyses using the updated TDT structural interface model and NMAS
model adjusted to match the first five fundamental frequencies.

Flutter Boundary and Test Envelopes Flutter and Test Envelops


on TDT Operating Envelopes
Test Point Key Cruise/Dash Configuration
Cruise/Dash Configuration Climb σ = 8.0 σ = 4.0 σ = 2.0
10000 Loiter 800

Cruise/Dash
High-Lift
σ = 1.0
HSM 700

Analysis Case
Key 600
Case 1
1000 Case 2
Dynamic Pressure (psf)

Equivalent Airspeed (keas)

500 σ = 0.5

400 σ = 0.3

σ = 0.2
300
100

σ = 0.1
200

100

10 0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2

Mach Mach

Figure 9. Cruise/Dash Flutter Boundary

9
American Institute of Aeronautics and Astronautics
Test Point Key
Flutter Boundary and Test Envelopes Climb Flutter Boundary and Test Envelops
on TDT Operating Envelopes Loiter Hi-Lift Configuration
Hi-Lift Configuration Cruise/Dash
High-Lift σ = 8.0 σ = 4.0 σ = 2.0 σ = 1.0
1000 600
HSM

Analysis Case
Key
Case 1 500 σ = 0.5

Case 2

Equivalent Airspeed (keas)


Dynamic Pressure (psf)

400
σ = 0.3

σ = 0.2
100 300

σ = 0.1
200

σ = 0.05

100 σ = 0.02

10 0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2

Mach Mach

Figure 10. High Lift Flutter Boundary

Test Point Key


Flutter Boundary and Test Envelopes Climb Flutter and Test Envelops
on TDT Operating Envelopes Loiter Configuration
Loiter
Loiter Configuration σ = 8.0 σ = 4.0 σ = 2.0 σ = 1.0
Cruise/Dash
1000 600
High-Lift
HSM

Analysis Case σ = 0.5


500
Key
Case 1
Case 2
400
σ = 0.3
Equivalent Airspeed (keas)
Dynamic Pressure (psf)

σ = 0.2
100 300

σ = 0.1
200

σ = 0.05

100

10 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4
Mach
Mach

Figure 11. Loiter Flutter Boundary

10
American Institute of Aeronautics and Astronautics
Flutter Boundary and Test Envelopes Flutter and Test Envelops
on TDT Operating Envelopes Test Point Key Climb Configuration
Climb Configuration σ = 8.0 σ = 4.0 σ = 2.0
10000 Climb 800
Loiter
Cruise/Dash
High-Lift 700 σ = 1.0

HSM

Analysis Case 600


Key
1000
Case 1

Equivalent Airspeed (keas)


Case 2
Dynamic Pressure (psf)

500 σ = 0.5

400 σ = 0.3

σ = 0.2
300

100

σ = 0.1
200

σ = 0.05

100

10 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4

Mach Mach

Figure 12. Climb Flutter Boundary

Test Point Keyσ = 1.0


Flutter Boundary and Test Envelopes Flutter and Test Envelops
Climb
on TDT Operating Envelopes HSM Configuration
HSM Configuration Loiter σ = 4.0 σ = 2.0
σ = 8.0
10000 Cruise/Dash 1000

High-Lift
HSM 900

Analysis Case
Key 800
Case 1
Case 2 700 σ = 1.0
1000 σ = 3.25
Equivalent Airspeed (keas)
Dynamic Pressure (psf)

600

500 σ = 0.5

400 σ = 0.3
100
σ = 0.2
300

σ = 0.1
200

100

10
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2
Mach
Mach

Figure 13. High Speed Maneuver Flutter Boundary

11
American Institute of Aeronautics and Astronautics
This procedure began with obtaining frequency response information from a suite of accelerometers, shown
Figure 14, for each of the five configurations as mounted in the TDT. The frequency response function (FRF) for
each configuration was obtained from a NASA Langley data acquisition system. As an example, the FRF for the
high lift configuration is shown in Figure 15.

Figure 14. Accelerometer Locations for TDT GVT

Figure 15. Frequency Response Function for High Lift

From the obtained FRFs, it was possible to identify the first five fundamental modal frequencies to within
approximately 0.25 Hz. Table 1 shows the frequency results for the high-lift configuration. Each morphed
configuration model was then adjusted, as discussed previously, to match these frequencies. Table 2 shows the GVT
frequencies with the frequencies obtained after model tuning and the corresponding percent differences. A further
qualitative check of the modal response of the structure was obtained by comparing the modeshapes obtained from
the GVT with those obtained analytically. Figures 16 and 17 show such a comparison for the high-lift configuration
for the first wing bending and wing first torsion modes.

12
American Institute of Aeronautics and Astronautics
Table 1: GVT Frequencies for High Lift Configuration

Configuration Frequency (Hz) Damping (%) Amplitude Phase (RAD) MCF


05.780 0.400 3.17E+01 1.571 0.999
10.265 0.909 3.70E+01 1.571 1.000
High Light 10.740 0.304 2.53E+01 1.571 0.996
15.300 0.755 3.21E+02 1.571 0.999
26.120 0.682 5.10E+02 1.571 1.000

Table 2: Analytically Obtained Frequencies from Tuned Model

Mode Test Freq (Hz) Analytic Freq (Hz) Difference (%)


1 05.78 05.58 -03.46
2 10.27 09.87 -03.89
3 10.74 11.45 06.61
4 15.30 17.88 16.86
5 26.12 25.71 -01.57

Test Analytic

Figure 16: Mode Shape Correlation – Wing First Bending – High Lift

Figure 17: Mode Shape Correlation – Wing First Torsion – High Lift

13
American Institute of Aeronautics and Astronautics
VII. Opportunities for Further Research and Development
NextGen’s experience during the design, analysis, and wind tunnel testing of the MAS model indicated several
areas in which advances in modeling and analysis techniques would enable more efficient and effective prediction of
steady and unsteady aeroelastic phenomenon. The opportunities for research and development work, both in terms
of general modeling techniques and flutter specific analyses are presented below.

A. Aeroelastic Modeling
A great deal of time was spent developing the aeroelastic model for each of five distinct configurations. Each of
the structural, aerodynamic, and aero/structure connection models was unique for each configuration, resulting in a
very time consuming process to generate all of the required models. In some cases, a certain amount of automation
was possible. For instance, it was possible to automate a fair portion of the structural FEM development. Knowing
the geometry of each configuration, it was possible to use the PATRAN Command LanguageTM (PCL) to correctly
position each wing structural node based up the known wing sweep angle and internal angle of the wing 4-bar
mechanism. Other modeling details, particularly the connections between the wing spar roots and outer channel of
the main 4-bar mechanism had to be changed by hand for each configuration.
Similar issues arose concerning the aerodynamic model development. The placement of each macroelement had
to be changed for each configuration. Fortunately, a check of the panel aspect ratios revealed that the same panel
model within each macroelement could be used for each configuration. This was because the morphing mechanism
was through shear only, resulting only a moderate amount of aspect ratio change. However, in the more general case
of morphing where shear and telescoping mechanisms are employed aspect could change considerably, thereby
requiring unique paneling for each configuration.
Finally, the spline model providing the aero/structural connection had to be inspected for each configuration.
Generally, the shear motion of structural model did not correspond with the spanwise change of each aerodynamic
macroelement. This resulted in the fact that a structural node that lay under a particular macroelement in one
configuration lay under a different panel in another. This required the splines to be reworked for each configuration.
These challenges could be addressed through the implementation of a kinematics solver that, knowing the
appropriate morphing degrees of freedom, could appropriately locate structural nodes and aerodynamic
macroelement corner locations. Beyond this, finite element and aerodynamic mesh generators would have to be
employed to fully automate the aeroelastic model development process.

B. Flutter Analyses
As mentioned previously, there are unique aeroelastic mechanisms presented by the novel nature of the MAS
wing structure. Of particular interest was the in plane motion as modal frequencies much lower than are traditionally
encountered. Even when in plane modes are encountered, they are typically assumed to contribute no out of plane
motion. Since the classical flutter techniques within NASTRAN only consider pitch and plunge motion, in plane
dynamics do not affect the flutter solution. However, the potential for in-plane dynamics to affect the flutter
behavior does exist, particularly when the in-plane modal frequencies are of the same magnitude as the out-of-plane
modal frequencies, as was the case with the MAS model. In addition to any out-of-plane motion that may be induced
through structural asymmetries, the fore-aft motion itself introduces a changing local velocity that would change the
local lift coefficient periodically with at the fore-aft modal frequencies. This would result in an out-of-plane
component induced by purely in-plane motion – a phenomenon not captured within the current frequency domain
NASTRANTM flutter solutions. Additionally, both steady and unsteady aeroelastic effects stemming from the motion
in the morphing degrees of freedom were not considered, and in fact were treated as quasi-static. For rapid motion in
the morphing degrees of freedom, however, the potential exists to generate transient dynamics and aerodynamics
that may be important in an aeroelastic sense. As morphing aircraft structures are pursued in the future, the
analytical means of handling a broader class of motion extending beyond the usual pitch and plunge motions is
desirable.

VIII. Conclusion
An evolutionary process was utilized to analyze and characterize the dynamic aeroelastic characteristics of the
NMAS TDT model. Increasingly accurate structural models were used to further confidence that the model would
encounter no aeroelastic instabilities within the test envelop of interest for each configuration. A purely analytic
process was used to present positive flutter margin results at the Model Systems Review. Beyond the MSR, GVT
and other structural stiffness data from both the NMAS wing model and the TDT interface structure became

14
American Institute of Aeronautics and Astronautics
available. Upon examination, it became clear that the stiffness as originally modeled in the TDT structure was
significantly higher than in reality.
Because of this situation, it became necessary to perform additional flutter margin checks as the actual TDT test
was being conducted. By this time a high-fidelity fuselage structural model had been added to the wing model.
Knowing the fundamental wind-off modal frequencies from GVT data, it was possible to tune each configuration
model to match, with reasonable accuracy, the test frequencies. Qualitative examination of the mode shapes
generated further confidence that the aeroelastic models were sufficient for flutter margin prediction. Positive flutter
margins over all the desired test points were predicted for all configurations, allowing an essentially unrestricted test
matrix throughout the NMAS TDT test.
NextGen’s experience gained during the flutter prediction efforts indicated significant opportunities for
aeroelastic modeling and analysis methods development. Aeroelastic characterization of morphing aircraft structures
would benefit from both advances in automated structural, aerodynamic, and spline modeling. Furthermore, the
ability to handle aeroelastic phenomena unique to morphing structures such as low frequency in-plane motion and
dynamics associated with the morphing degrees of freedom would be advantageous.

Acknowledgement
The research and development effort documented herein was performed under Contract No. F33615-02-C-3257,
“Next-Generation Morphing Aircraft Structures (N-MAS).” This contract was sponsored and funded by DARPA
with Dr. Terrence A. Weisshaar as the program manager. AFRL-Dayton, OH, served as the contracting agency with
Dr. Brian Sanders and Dr. Bryan Cannon as the technical monitors. HyPerComp was the prime contractor for the
program with NextGen Aeronautics as the major subcontractor; NextGen also served as the technical lead
organization.

15
American Institute of Aeronautics and Astronautics

You might also like