You are on page 1of 23

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/283020371

Nonlinear aeroelastic analysis of sugar truss-braced wing wind tunnel model


under in-plane loads

Conference Paper · January 2015

CITATIONS READS

7 5,406

6 authors, including:

Wei Zhao Rakesh Kapania


Oklahoma State University - Stillwater Virginia Tech (Virginia Polytechnic Institute and State University)
52 PUBLICATIONS 503 CITATIONS 558 PUBLICATIONS 8,175 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Wei Zhao on 04 November 2015.

The user has requested enhancement of the downloaded file.


AIAA 2015-1173
AIAA SciTech
5-9 January 2015, Kissimmee, Florida
56th AIAA/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference

Nonlinear Aeroelastic Analysis of SUGAR


Truss-Braced Wing Wind Tunnel Model under
In-plane Loads

Wei Zhao∗, Rakesh K. Kapania†, Joseph A. Schetz‡ and John M. Coggin§


Virginia Polytechnic Institute and State University, Blacksburg, VA, 24060, USA

Timothy J. Allen¶ and Bradley W. Sextonk


The Boeing Company
Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

russ-braced wing (TBW) aircraft consist of two stiffening members, a strut and a jury
Tadded to each wing, which carry nearly all the compression or all the tension loads
depending on load factors. The axial loads in the strut are transferred to the main wing
and cause additional in-plane tensile or compressive loads in the inner wing. The jury con-
necting the strut and the main wing also restrains the wing deflection under aerodynamic
forces. The in-plane loads, deflection restraints and geometric nonlinearities associated
with the large deflection significantly affect the wing structural behavior and hence influ-
ence the wing aeroelastic behavior. We present a nonlinear aeroelastic analysis scheme
where the analysis is performed with respect to the deformed structure considering the
structural nonlinearities. The Doublet Lattice Method (DLM), a linear unsteady aero-
dynamics model, is employed for the unsteady aerodynamic forces. This linear unsteady
aerodynamics model was corrected by utilizing Aerodynamic Correction Factors (ACF)
obtained from CFD to capture the transonic effects. An iterative method is developed to
obtain the flutter speed of the pre-stressed TBW under in-plane loads. The amount and
the distribution of the in-plane loads are influenced by several parameters including the
Angle of Attack (AoA), ACF, and the ballast masses (lumped mass), which have substan-
tial effect on the aeroelastic behavior of the TBW. The influence of these parameters on
flutter speed is investigated for the case of the SUGAR TBW wind tunnel model designed
for aeroelastic testing in the NASA Transonic Dynamics Tunnel. Flutter boundaries of the
updated FEM version of the SUGAR TBW WTM, FEM v20, are computed with both the
Boeing original and updated steady ACF using both linear and nonlinear aeroelastic anal-
ysis. Comparisons between theoretical predictions and experimental results are presented.
The same phenomena is observed that the flutter speed is sensitive to AoA regardless of
Mach number in both the wind tunnel test and theoretical predictions. It is also found
that the steady ACF may be not appropriate to be used as ACF for unsteady aerodynamic
forces in flutter analysis, and further studies are needed.

I. Introduction
The Subsonic Ultra-Green Aircraft Research (SUGAR) project led by the Boeing Company has resulted in
the NASA N+3 initiative of investigating the TBW aircraft design, as shown in Fig. 1, as a next generation
aircraft concept for reduced fuel consumption and emissions.1 Recent studies2, 3, 4 at Virginia Tech on
multidisciplinary design optimization of TBW aircraft design showed that these airplanes have great potential
∗ Graduate Research Assistant, Department of Aerospace and Ocean Engineering, AIAA Student Member.
† Mitchell Professor, Department of Aerospace and Ocean Engineering, Lift time Associate Fellow, AIAA.
‡ Fred D. Durham Chair, Department of Aerospace and Ocean Engineering, Life Time AIAA Fellow.
§ Adjunct Faculty, Department of Aerospace and Ocean Engineering, AIAA Member.
¶ Principal Engineer, Boeing Research&Technology(BR&T), Huntington Beach, CA.
k Principal Engineer, Boeing Research&Technology(BR&T), St. Louis, MO.

1 of 22

American Institute of Aeronautics and Astronautics


Copyright © 2015 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
as a more fuel efficient alternative than conventional commercial transport aircraft using cantilever wings.
The truss system consists of two stiffening members (strut and jury) which brace the wing against bending
loads, thereby enabling the aerodynamic efficiency of a high aspect ratio wing but without the weight penalty
of a cantilevered wing design with the same span. The advantages of the TBW configurations result in larger
aspect ratio wings, lower chords and thinner wings. Unlike the cantilevered wings whose vibration modes
are independent of the load factors, the in-plane loads in the inner wing of the TBW can have significant
influence on the structural dynamic behaviors of TBW aircraft designs.5, 6 The added stiffening members,
the jury and the strut, make the load paths different from those of conventional cantilevered wing designs.
The axial loads generated in the strut will be transferred to the inner wing, and cause additional in-plane
normal tensile or compressive stresses, which can increase or decrease the transverse structural stiffness.
The jury, a vertical stiffening member connecting the main wing and the strut, restrains the deformation
of both the strut and the main wing, which changes the wing loading conditions and hence influences the
wing structural behavior. The generated compression and tension loads in the inner wing will decrease or
increase the flutter speed depending on load factors. The large deflection typically associated with the large
aspect ratio wings and the in-plane loads in the TBW lead to a need for performing a comprehensive and
complicate aeroelastic analysis of the TBW under various aerodynamic loads. The linear aeroelastic analysis,
linear structural model and linear unsteady model, can’t predict the flutter boundary accurately for such
Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

configurations. So, it is necessary to develop a nonlinear aeroelastic analysis capability by considering both
the in-plane loads and the geometric nonlinearity in the structural analysis. For the TBW in transonic flight,
aerodynamics nonlinearity such as compressibility and shock waves should also be considered.

Figure 1: The Subsonic Ultra Green Aircraft Research (SUGAR) initiated by NASA has resulted in the TBW concept shown here

Nonlinear aeroelasticity has received a great deal of attention in recent years motivated by high-altitude
long-endurance (HALE) wing aircraft configurations and unconventional aircraft designs such as the truss-
braced wings and joined wings.5, 6, 7, 8, 9, 10 Proper analysis of nonlinear aeroelastic phenomena can result
in either increased or decreased design margins when compared to the traditional linear analysis or small
perturbation methods. The work presented here is concerned with aeroelastic analysis including static
structural nonlinearities arising from the geometric stiffness properties of a TBW aircraft (see Fig. 2). To
provide context for our work, we first provide a brief review of nonlinear aeroelasticity in general, and then
summarize the recent literature concerning closely related work on geometric nonlinearities and aeroelastic
analysis of nonconventional wing structures.
Dowell et al.8 provided an overview of specific nonlinear aeroelastic phenomena and distinguish between
aerodynamic nonlinearities such as shock and stall behaviors, and structural nonlinearities such as geometric
stiffness, damping, and freeplay. They further classified the nonlinear aeroelastic phenomena into either static
or dynamic nonlinearities. Static nonlinearity in structures implies that the static deformations are large
enough to no longer be proportional to the applied forces. Similarly, static nonlinearity in the aerodynamics
implies that the aerodynamic pressures are no longer proportional to the changes in downwash or fluid
velocity due to the motion of the body within the flow. Dynamic nonlinearity implies that the structural
or aerodynamic perturbations about the statically deformed state are nonlinear, an example of which is
the time dependent growth of the amplitude of vibration until it reaches the limit cycle oscillation. Livne

2 of 22

American Institute of Aeronautics and Astronautics


Figure 2: The Truss Braced Wing Aircraft (top) includes a vertical truss member, jury, not found on the simple Strut Braced Wing
Aircraft (bottom)
Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

and Weisshaar11 provided a thorough review of aeroelasticity of nonconventional aircraft configurations and
described various nonlinear phenomena of interest.
The work presented here describes an aeroelastic analysis which includes the structural static nonlin-
earities for large displacements and geometric stiffness changes, but with an aerodynamics model that is
statically linear. The structural and aerodynamic models are both dynamically linear with respect to the
statically deformed equilibrium position. Much of the past work in the literature which included geometric
stiffness nonlinearities had been performed for conventional cantilever wings, primarily emphasizing high
aspect ratio platforms that have inherently large displacements. van Schoor and von Flotow12 described
aeroelastic characteristics of MIT’s highly flexible human-powered airplane, the Michelob Light Eagle. Patil,
Hodges, and Cesnik,13, 14 and Tang and Dowell15 both presented works describing the nonlinear aeroelastic
behavior of a generic HALE aircraft including geometric structural nonlinearities and aerodynamic nonlin-
earities such as dynamic stall. Demasi, Cavallaro, and Razn10 performed postbuckling analysis of a Prandtl
Wing (joined wing) aircraft. While the constrained nature of the joined wing shares similarities to the TBW,
the work did not consider the flutter behavior.
Demasi and Livne9 developed two formulations for the aeroelastic coupling of geometrically nonlinear
structures and linear unsteady aerodynamics. The geometrically nonlinear structural model for thin-plate
aerospace structures used was created using flat triangular elements based on the Discrete Kirchhoff Theory.
The doublet lattice method was used as a representative linear unsteady aerodynamic method, which was
expressed in the time domain using the Fourier transform. To capture the aerodynamic nonlinearity due
to transonic effects, Bartels et al.16 coupled linearly stressed TBW WTM with the unstructured FUN3D
Navier-Stokes computational fluid dynamics(CFD) code. In their work, the geometric nonlinearity was
ignored.
Finally in a relevant work, Sulaeman17 considered the flutter analysis of a Strut-Braced Wing (SBW)
aircraft (see Fig. 2) where the compressive force in the inner wing due to 1g flightloads was accounted for.
In that work, Sulaeman examined the influence of aircraft/strut geometry on the flutter speed. He utilized a
linear elastic solution to calculate the compressive force and showed that the predicted flutter speed decreases
relative to a baseline flutter analysis in which the compressive force was not included.
The nonlinear aeroelastic analysis presented here is performed using beam element structural FE models
and a corrected doublet lattice aerodynamics model. In contrast to the work by Sulaeman, we utilize both
linear elastic and nonlinear elastic large displacement formulations to calculate the structural in-plane loads
of the vehicle subject to self-weight and aerodynamic loads. The present work also examines a range of
AoA, ACF, and ballast mass for their influence on flutter speed. The results presented here for linear and
nonlinear pre-stressed aeroelasticity are compared against both the experimental results and a traditional
flutter analysis not including pre-stress. This traditional flutter analysis is referred to hereafter as the
”unloaded model”.

3 of 22

American Institute of Aeronautics and Astronautics


II. SUGAR Truss-braced Wing Wind-tunnel Model
As part of the NASA SUGAR research program, a wind tunnel model (WTM) as shown in Fig. 3 has
been designed, fabricated and tested in the NASA TDT during late 2013 and early 2014. The flutter analysis
presented here is developed to predict the aeroelastic behavior of the WTM under a variety of flight conditions
for the wind tunnel tests. The specific finite element model utilized for the present study was developed by
the Boeing Company with the purpose of simulating the TBW WTM. The structural characteristics of the
FEM were updated to reflect the WTM stiffness distribution as constructed.
Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

Figure 3: NASA TDT SUGAR Truss-braced Wing Wind-tunnel Model

The detailed structural full-scale FEM, shown in Fig. 4, was used as the starting point for developing the
SUGAR N+3 wind tunnel model. This full-scale model has a 2039.3 inches span with an aspect ratio 19.55.
The detailed FEM was developed for calculating structural loads and performing structural optimization,
and it was based on the Outer Mold Lines (OML) defined by the high fidelity multidisciplinary disciplinary
optimization(MDO).2, 3 The FEM mesh was defined sufficient to capture potential local buckling behavior
and provided good stress results. Detailed control surface models were incorporated.

Figure 4: The detailed full-scale structural FEM was the starting point for development of the beam-rod FEM used for the present
work. The strut connects the fore spar and the gear pylon.

The Boeing Company was responsible for determining the proper dynamic scale factors, and it was
decided that a semispan wind tunnel model would be best to test. Static analysis of the detailed full-scale
FEM was conducted to determine the equivalent beam stiffness for all relevant structural components (main
wing, strut, and jury). The equivalent beam stiffness and mass data were scaled and provided to NextGen
Aeronautics (model builder) for model design. The scaled stiffness and mass data are incorporated in the
scaled analytical model which is used in all aeroelastic analyses of the TBW WTM in this paper. The
analytical model was periodically updated as the NextGen Aeronautics Company finalized the model design,
and it was updated to accurately simulate the WTM as constructed. The latest equivalent beam-rod FEM
of the TBW WTM is version 20 as shown in Fig. 5.

4 of 22

American Institute of Aeronautics and Astronautics


Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

Figure 5: The latest beam-rod FEM version 20, developed by The Boeing Company.

III. Aerodynamic Correction Factors


To simulate the compressibility and shock behavior lacking in the DLM, CFD-based aerodynamic correc-
tion factors(ACF) were utilized to correct the DLM results in NASTRAN. The ACF constitutes ”panel by
panel” corrections to the lift slope and also to the pressure magnitude. The key behavior captured through
addition of ACF is the shift of the center of pressure due to transonic effects, which can alleviate the flut-
ter stability at critical transonic Mach numbers. In NASTRAN, ACF can be input using the matrix Wkk ,
experimental dynamic pressure can be input using the vector FA2J, and adjustments to the downwash for
effects of camber and initial incidences, can be input using the vector W2GJ.18
In the integral model formulation, the matrix W kk is used to correct the lift slope, which is generally a
diagonal matrix, and f ej is used to correct the dynamic pressure.

f ej
P k = W kk S kj f ej + q̄S kj (1)

The initial downwash from the camber and the effect from twist is included using the W2GJ vector, wjg

f j = A−1
jj w j (2)

wj = D jk uk + D jx ux + wgj (3)
where the subscript j denotes the j-th aerodynamic element in the DLM.
Figure 6 shows the results from previous studies1 from the Boeing Company on ACF used to correct
DLM for aerodynamic forces. Comparisons of sectional lift slope CL and aerodynamic center (AC) along the
span from the CFD code, Overflow, and corrected DLM with ACF for the full-scale TBW model show that
they are in good agreement. The ACF, including the lift slope correction Wkk , and the dynamic pressure
corrections, FA2J, were included in the static aeroelastic analysis to simulate transonic effects in all the
following aeroelastic analyses. The steady ACF were updated based on the wind-tunnel test results. The
Boeing steady ACF can also be used as unsteady ACF in flutter analysis. The following work will also
evaluate the use of steady ACF as unsteady ACF.
The flight conditions employed for the TBW WTM flutter analysis are based on the full-scale aircraft
when subjected to the flutter speed constraint of 1.09Vd in structural optimization. Results for the full-scale
aircraft indicated that the critical Mach number for flutter was 0.82, when ACF are applied for both the
steady and the unsteady aerodynamic forces. The WTM was tested in the NASA TDT, where the heavy
gas, R-134a gas, was used. So the heavy gas density was used instead of the normal gas density at sea level.

5 of 22

American Institute of Aeronautics and Astronautics


Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

(a) Sectional lift curve slope CL comparison

(b) Sectional aerodynamic center (AC) comparison

Figure 6: Comparisons of CL and AC from Overflow CFD and Corrected Doublet Lattice Method (DLM)(Mach=0.70, 0.76, 0.82, 0.86,
0.90 and 0.94)

6 of 22

American Institute of Aeronautics and Astronautics


IV. Closed-form Aeroelastic Analysis
To perform the closed-form aeroelastic analysis of the TBW WTM under in-plane loads, we developed a
solution scheme for this analysis as shown in Fig. 7.5, 6 Several solution sequences available in NASTRAN
for stress and flutter analysis are utilized. The solution scheme involves several steps: 1) obtain the gravity
loads due to self-weight; 2) obtain initial/updated estimate of dynamic pressure (Q) at the flutter velocity
of the unloaded model or previous output flutter speed for static aeroelastic analysis; 3) select the desired
Angle of Attack (AoA); 4) obtain aerodynamic loads from a static aeroelastic analysis; 5) perform linear or
nonlinear static analysis to determine the in-plane loads; 6) perform pre-stressed modal analysis to get the
pre-stressed mode information; and finally 7) perform flutter analysis using the pre-stressed modes. Once
the pre-stressed flutter solution is obtained, the dynamic pressure at this new flutter velocity is used as an
improved estimate as input to the calculation of aerodynamic loads in Step 4. Iteration is performed on
Steps 4, 5, 6, and 7 as the flutter dynamic pressure converges to a preset level,
Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

Figure 7: Flow chart of the solution scheme for closed-from aeroelastic analysis of SUGAR TBW WTM using NASTRAN based solution
sequences.

An iterative method was developed to perform aeroelastic analysis of the TBW WTM as shown in
Fig. 8. Flutter speed of the unloaded TBW WTM was used as the initial flow velocity V 0 to calculate the
dynamic pressure q 0 in the static aeroelastic analysis. Following the analysis steps, we have the output flutter
velocity V 1 . Comparing the difference of |V 1 − V 0 | is less than the convergence criteria  (a user-defined
value) or not to determine whether the program is terminated or not. If not converged, the velocity V 1 is
considered as input flow velocity for dynamic pressure q 1 in the next iteration until convergence is achieved,
i.e. |V i+1 − V i | ≤ .
For some cases, there is no output flutter speed V i+1 following the analysis steps shown above under the
input velocity V i . Then, bisection for the input velocity was utilized by averaging previous velocity values
V i +V i−1
2 as the input velocity V i+1 in the next iteration. When there is no flutter at the first initial guess
velocity, a user-defined perturbation value δ was used to adjust the input velocity value V 0 .
An alternative method is the discrete frequency method, shown in Fig. 9. In this method, a number
of separate calculations are performed to obtain the flutter structural damping corresponding to a series of
input velocities/dynamic pressures in the static aeroelastic analysis, followed by linearly interpolating the
data in the V − g plot to find the flutter dynamic pressure or flutter velocity where the damping is equal
to zero or a user-defined structural damping value. However, due to a possible change of the modes whose
coupling causing flutter under various flight conditions, it is not easy to recognize a priori which critical
mode in the V − g plot passes through zero damping (or user-defined damping value). Also, there is no
flutter in some cases. However, the flutter speed could be more efficiently obtained by using the discrete
frequency method once the critical mode in the V − g plot is fixed at Mode 3 of the SUGAR TBW WTM
studied here, for all AoAs and Mach numbers, by using parallel computing. This method is mainly to be
used as a validation tool to examine the results obtained from the iterative method. In this paper, the
iterative method is used to find the open-loop flutter boundary, and the discrete frequency method is used
for validation and as an alternative method for cases where convergence could not be obtained when using

7 of 22

American Institute of Aeronautics and Astronautics


Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

Figure 8: Iteration framework used to find the flutter boundary in MATLAB@ environment using NASTRAN based solution sequences
for flutter analysis

the iterative method.

Figure 9: Flow chart of the discrete frequency method utilized for computing flutter dynamic pressure. Parallel computing was executed
for this method.

A. Calculation of Loads
There are two types of loads that are used in our calculations, self-weight and aerodynamic loads. In contrast
to an aircraft in free flight, which can experience arbitrary g-loadings associated with maneuvers, the TBW
WTM is fixed to the wind tunnel wall through a mounting system. Therefore it is only subject to a 1g
inertial loading, i.e. self-weight. This self-weight can be obtained by including the gravity force in the linear
static analysis (NASTRAN Sol 101) or static aeroelastic analysis (NASTRAN Sol 144). The aerodynamic
loads due to the given AoAs are obtained from static aeroelastic analysis in the range of -3 degrees to +5
degrees AoA. The vehicle deformations for the case of -3 degrees AoA and 5 degrees AoA are shown below
in Fig. 10a and Fig. 10b, respectively. From these deformation plots, we can see the wing bending downward
and upward under negative AoA and positive AoA, respectively.

B. In-plane Loads
Both linear static analysis (NASTRAN solution 101) and geometrically nonlinear static analysis (NASTRAN
solution 106) were performed to calculate the in-plane loads in the structural members. The structural

8 of 22

American Institute of Aeronautics and Astronautics


Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

(a) AoA=-3 degree

(b) AoA=5 degree

Figure 10: NASTRAN Solution 144 static aeroelastic analysis used to calculate the aerodynamic loads

9 of 22

American Institute of Aeronautics and Astronautics


stiffness was updated during the nonlinear static analysis to generate the in-plane loads considering the
geometric nonlinearity. Stiffness was not updated in the linear static analysis for the in-plane loads. In either
case, the gravity and aerodynamic loads calculated previously were applied to the FEM, and deformations,
stresses, and the geometric stiffness matrix are obtained. The geometric stiffness matrix was saved for use
in the subsequent pre-stressed mode and flutter analysis. Although the geometrically nonlinear analysis
will provide a better estimate of the deformations and stresses, it is interesting to examine the linear static
solution for its potential to provide useful estimates at a reduced computational expense.

C. Pre-stressed Modal Analysis


Once the vehicle deformations are obtained using either linear or nonlinear static analysis, a normal mode
analysis (NASTRAN solution 103) is performed using the pre-stressed or differential stiffness matrix asso-
ciated with these deformations. The in-plane loads cause a shift in the normal frequencies, just as a guitar
string changes pitch when the tension in it is altered. In the case of a simple beam, the bending modal fre-
quencies increase with beam in-plane tension and decrease with in-plane compression. The TBW structure
consisting of a wing, strut, and jury is more complex than a simple beam, and the global mode shapes and
the corresponding modal frequencies may be influenced by the in-plane loads developed in the inner wing as
Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

well as the strut and the jury.

D. Flutter Analysis
The flutter analysis (NASTRAN solution 145) performed in Steps 2 and 7 above utilizes a non-looping
matched-point flutter method, also known as the PKNL method. The flutter search for the Mach 0.82 case
includes density ratios and velocities spanning the test flight envelope. Reduced frequencies are chosen from
0.001 to 10.0 in order to cover the frequency range of interest. The modal damping in the aeroelastic analysis
is assumed as ξ = 0.02. In Step 2 above, a traditional flutter analysis of the unloaded model is performed in
order to obtain an approximate dynamic pressure for use in the next step to generate the aerodynamic loads.
The unloaded (no pre-stress) normal modes are utilized in Step 2. In contrast, the flutter solution in Step
7 utilizes the pre-stressed normal modes associated with the deformed structure. As the modal damping is
included in flutter analysis, flutter speed is obtained by interpolating the data in the V − g plot at structural
damping g = 0.

V. Results and Discussions


The above FEM and analysis procedures were utilized to generate results across a range of AoAs. Para-
metric studies were performed to examine the influence of ACF, AoA and ballast mass on flutter boundaries.
As a first step, analysis of the unloaded TBW WTM was performed to provide a baseline for comparison
and to provide the initial dynamic pressure in static aeroelastic analysis for in-plane loads. The ACF were
always applied to the steady aerodynamic forces in the static aeroelastic analysis in the following work to
generate aerodynamic loads regardless of the analysis type. The steady ACF could also be used to correct
the unsteady aerodynamic forces. Flutter dynamic pressure of the cases with ACF used for the unsteady
aerodynamic forces was examined to study the use of the steady ACF for unsteady aerodynamic forces. The
case w/ ACF or w/o ACF in the subsequent work means that there are ACF or no ACF included for the
unsteady aerodynamic forces in the flutter analysis.

A. Traditional Flutter Analysis (Unloaded Model)


Traditional flutter analysis of the FEM does not include pre-stress and is referred to here as the ”unloaded
analysis”. This unloaded analysis is used as a baseline for comparison to the pre-stressed solutions. Analysis
of the unloaded model is a single step in NASTRAN solution 145, since the influence of aerodynamic loads
and gravity loads are not considered. The flutter analysis parameters and PKNL method are consistent with
those described in Section IV.D above. Results for the unloaded model presented in Fig. 11 are for ballast
masses of 2.92 lbs., and include the ACF for both the steady and unsteady aerodynamic forces.
Figure 11 illustrates the flutter results for the unloaded model. Figure 11a provides the velocity vs.
damping response of the first four normal modes, and it can be seen that mode 3 crosses the zero damping
line and produces flutter instability at approximately 120 KEAS. Inspection of Figure 11b indicates that

10 of 22

American Institute of Aeronautics and Astronautics


Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

(a) V-g plot

(b) V-f plot

Figure 11: V-g plot and V-f plot for the unloaded SUGAR TBW WTM at Mach=0.82 with each ballast mass 2.92lbs.

11 of 22

American Institute of Aeronautics and Astronautics


mode 3 is coupling with mode 4 to produce the flutter instability. It was found that these same modes 3
and 4 are responsible for the flutter instabilities for most of the cases reported in the subsequent work we
present for the pre-stressed aeroelastic solutions.

B. Influence of ACF on Flutter Velocity


The doublet-lattice method (DLM) aerodynamics model utilized in NASTRAN can’t capture the transonic
phenomena such as the aft shift in center of pressure (Cp) due to shock formation. To improve the prediction,
the aerodynamic forces calculated based on DLM were modified by using ACF. These ACF was obtained
from CFD analysis in this paper. In the following figures, we illustrate the impact that the ACF can have
on the flutter speed (Vf) prediction. The results were generated for the case of ballast masses of 2.92 lbs.,
and for AoA ranging from -1 to 5 degrees. Results are provided for both linear and nonlinear in-plane loads,
with and without ACF applied to the unsteady aerodynamic forces.
Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

Figure 12: Flow velocity vs. damping for the case of 1 degree AoA. Linear and nonlinear analysis provided w/ and w/o ACF (23 initial
guess of velocities).

Figure 12 shows the velocity vs. damping response of the critical mode 3 for the case of 1 degree
AoA, which is obtained from the discrete frequency method. It can be seen that the linear analysis w/o
ACF provides the lowest estimate of flutter dynamic pressure (101 KEAS), followed by nonlinear w/o ACF
(110 KEAS) and then linear w/ ACF (120 KEAS). The nonlinear analysis w/ ACF does not have a zero
damping crossing, and therefore flutter instability is not predicted. This analysis contains the highest fidelity
representation of the physics involved and provides a significantly different prediction (no flutter) than the
result shown in Fig. 11 (120 KEAS) for the unloaded case.
Figure 13 shows the iteration history of the flutter dynamic pressure for the cases of 1 degree and -1
degree AoA with Mach number 0.82 using both linear and nonlinear analysis, w/ and w/o ACF applied
to the unsteady aerodynamic forces. It was found that a lesser number of iterative loops is needed for the
aeroelastic analysis w/o ACF applied to the unsteady aerodynamic forces than for the analyses w/ ACF
applied to the unsteady aerodynamic forces. The use of fewer loops in the nonlinear aeroelastic analysis w/o
ACF applied to the unsteady aerodynamic forces is important because that is the case that is preferred based
on physics. A least 10 initial guesses of velocities were used for the discrete frequency method to satisfy the
accuracy of results when linearly interpolating the data in the V − g plot. The iterative method is found to
be more efficient and more accurate than the discrete frequency method.

12 of 22

American Institute of Aeronautics and Astronautics


Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

(a) AoA=1degree, w/ACF (b) AoA=1degree, w/o ACF

(c) AoA=-1degree, w/ACF (d) AoA=-1degree, w/o ACF

Figure 13: Iteration history of flutter dynamic pressure using linear and nonlinear analysis

13 of 22

American Institute of Aeronautics and Astronautics


Table 1 shows the flutter dynamic pressures using both linear and nonlinear analysis, w/ and w/o ACF
applied to the unsteady aerodynamic forces, from both of the iterative method and the discrete frequency
method at Mach number 0.82 and 1 degree AoA. It is seen that the flutter dynamic pressure calculated
from the two methods are almost same. Another case, with -1 degree AoA, is chosen to validate the flutter
dynamic pressure from the iterative method.

Table 1: Flutter dynamic pressure(psf) using both linear and nonlinear analysis under Mach=0.82, AoA=1degree, ballast mass of 2.92 lbs.
w/ and w/o ACF applied to the unsteady aerodynamic pressure

Method Linear w/ACF Nonlinear w/ACF Linear w/o ACF Nonlinear w/o ACF
Iterative method 110.71 N/F* 78.36 93.06
Discrete frequency method 110.69 N/F 78.36 93.34

*N/F means No Flutter predicted


Table 2 shows the flutter dynamic pressures for the case with -1 degree AoA, and we can see that the
two sets of results obtained from the discrete frequency method and the iterative method are also close to
Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

each other. Generally, parallel computing can be utilized for the discrete frequency method provided that
the critical mode which passes through zero structural damping in the V − g plot remains unchanged. For
instance, mode 3 could be the critical mode under all test flight conditions for the TBW WTM studied here.
However, as the compressive stress and tensile stress generated in wing components vary with AoA and
dynamic pressure, the mode information and flutter response would also be changed. So, it is hard to know
a priori which two modes of the pre-stressed model would couple to cause flutter in the discrete frequency
method. So, the iterative method is chosen for all the following flutter analysis.

Table 2: Flutter dynamic pressure(psf) using both linear and nonlinear analysis at Mach=0.82, AoA=-1degree, ballast mass of 2.92 lbs.
w/ and w/o ACF applied to the unsteady aerodynamic forces

Method Linear w/ACF Nonlinear w/ACF Linear w/o ACF Nonlinear w/o ACF
Iterative method 100.71 N/F* 72.67 86.02
Discrete frequency method 99.87 N/F 72.62 86.15

*N/F means No Flutter predicted at this analysis type


Figure 14 provides results for flutter velocity across the range of AoA examined. It was found that for
the case of nonlinear analysis w/ ACF applied to the unsteady aerodynamic pressure, no flutter instability
was observed at any AoA (so that it is not shown in the plot). The model flutter speed for the unloaded
wing is shown by a horizontal line, since the pre-stress is not included and Vf is not a function of AoA. It
can be seen that the variation of the predicted flutter speed increases with AoA regardless of analysis type
(linear or nonlinear, with or without ACF used). It can also be seen that analysis with ACF results in a
higher predicted flutter velocity.

C. Influence of Ballast Mass on Flutter Velocity


The structural dynamics model matches the ”as built” model, with lumped masses installed in the forward
wing, strut, and jury used to conduct flutter and loads analysis prior to the first TDT entry. These lumped
masses are the ballast masses. Six ballast mass locations were developed for the TBW WTM in order to
improve modal similitude with the full-scale FEM. The locations of these six masses are shown in Fig. 15.
Each mass is nominally 2.92 lbs.
Both the mass distribution and magnitude have a significant influence on mode frequencies and, therefore
flutter boundaries. The results presented in Section V.B indicate that the highest fidelity analysis (nonlinear
w/ ACF on unsteady aerodynamic forces) did not predict any flutter instability. Therefore, we examined the
flutter response as the ballast mass (6 locations at 2.92 lbs. each) was first increased by 50% (6 locations at
4.38 lbs. each) and then by 100% (6 locations at 5.84 lbs. each). The results in this section were generated
at Mach 0.82 and included ACF employed on unsteady aerodynamic forces.
Figure 16 illustrates the flutter speed prediction for the unloaded model at three different ballast mass
conditions. It can be seen that flutter speed is reduced by nearly 20% by doubling the ballast mass from 2.92

14 of 22

American Institute of Aeronautics and Astronautics


Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

Figure 14: Flutter Speed vs. AoA for analyses w/ and w/o ACF applied to the unsteady aerodynamic pressure. Both linear and
nonlinear results are provided for the case of Mach number 0.82 and 2.92 lbs. ballast mass

Figure 15: Ballast mass added to the WTM to improve modal similitude with the full-scale aircraft.

15 of 22

American Institute of Aeronautics and Astronautics


Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

Figure 16: Flutter speed of the unloaded TBW WTM at three different ballast mass, Mach=0.82

Figure 17: Flutter speed of the pre-stressed model using linear analysis w/ACF at three different ballast masses, Mach=0.82

16 of 22

American Institute of Aeronautics and Astronautics


Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

Figure 18: Flutter speed of the pre-stressed model using nonlinear analysis w/ACF at three different ballast masses, Mach=0.82

to 5.84 lbs. Although this is a significant change in the ballast mass, it only represents a 3.36% change to
the TBW WTM weight, as indicated in the legend of Fig. 16. It is the ballast mass located aft of the elastic
axis which decreases the bending and torsion frequencies that reduce the flutter speed. This illustrates that
the flutter behavior of the model is very sensitive to the changes in the ballast mass.
Figure 17 provides flutter speed predictions of the pre-stressed model using linear analysis for the three
ballast mass conditions. The responses of all three ballast mass cases indicate the previously seen behavior
for flutter speed to increase with increasing AoA. Similar to the results of the unloaded model in Fig. 16,
flutter speed is reduced by approximately 20% by doubling the ballast mass.
Figure 18 provides the flutter speed predictions of the pre-stressed model using nonlinear analysis for
the three ballast mass conditions. As noted previously for a ballast mass of 2.92 lbs., nonlinear analysis w/
ACF does not predict a flutter instability. Therefore, the results in Fig. 18 only include the ballast mass
cases of 4.38 lbs. and 5.84 lbs., which showed that the flutter instability happened within the planned test
envelope. The response of these two mass cases indicates the previously seen behavior for flutter speed to
increase with increasing AoA. The magnitude of flutter speed reduction due to the increased ballast mass is
similar to the previous cases for the unloaded model and the linear analyses.

D. Pre-stressed Modal Information


1. Change of Modal Frequencies
In order to better understand the underlying mechanisms by which pre-stress influences flutter speed, we
examine the modal frequencies across the range of AoA. The fundamental mechanism influencing flutter
speed prediction involves the change in the modal frequency of the critical flutter mode 3 (2nd bending
mode) and to a lesser extent mode 4 (1st torsion mode). When the natural frequency of mode 3 increases,
a corresponding increase in flutter speed can be seen. We examined the variation of modal frequencies
of modes 3 and 4 with AoA at Mach number 0.82. Variations of the normalized modal frequencies with
AoA using both linear and nonlinear analysis were investigated, and the results are shown in Fig. 19. The
frequency here was normalized with respect to the corresponding mode frequency of the unloaded model.
The natural frequency of the third mode increased by about 2.5% compared to that of the unloaded model
up to 5 degrees AoA, whereas, the natural frequency of the fourth mode increased by approximately 1% up
to 5 degrees AoA.

17 of 22

American Institute of Aeronautics and Astronautics


Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

(a) Mode 3 (b) Mode 4

Figure 19: Variation of the normalized modal frequencies with AoA using both linear and nonlinear analysis with steady ACF; with
ballast mass 2.92 lbs.

Figure 20: Mode shapes of Mode 3 and Mode 4 of the unloaded model, the linear prestressed model and the nonlinear prestressed
model at -1 degree AoA and Mach number 0.82.

18 of 22

American Institute of Aeronautics and Astronautics


2. Change of Mode Shapes
The doublet-lattice method (DLM)-based aerodynamic forces also depend on the structural mode shapes.
The change in mode shapes also influences the flutter boundary. Figure 20 shows comparisons of mode
shapes for mode 3 and mode 4 of the unloaded model, linearly pre-stressed model and nonlinearly pre-
stressed model. It was found that mode shapes of modes 3 and 4 of the unloaded model and the linearly
pre-stressed model are similar. The mode shapes of the nonlinearly pre-stressed model are very different
from the mode shapes of the unloaded model and the linearly pre-stressed model. These changes of mode
shapes and natural frequencies of the nonlinearly pre-stressed model cause the change in the flutter boundary
of the TBW WTM when using a nonlinear analysis for different AoAs.

E. Flutter Boundaries of Updated Models and Test Results


The finite element model of the SUGAR TBW WTM was updated to FEM v20 based on ground vibration test
(GVT) results. The steady ACF were also updated based on the wind tunnel tests. The flutter boundaries
for the latest FEM with updated steady ACF were computed. It was found that there is no flutter predicted
when employing steady ACF as the ACF for unsteady aerodynamic pressure while performing a nonlinear
Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

flutter analysis(seen Fig. 14), which is not consistent with the phenomenon observed in the wind tunnel
test. So, no ACF were applied to the unsteady aerodynamic forces in the pre-stressed flutter analysis while
performing aeroelastic analysis of the TBW WTM for FEM v20. There are two subsections in this section,
the first subsection discusses how the flutter dynamic pressure varies with AoA ranging from -3 degrees to 3
degrees for FEM v20 using both linear and nonlinear analysis w/o ACF applied to the unsteady aerodynamic
forces. The next subsection shows the comparisons between theoretical predictions and experimental results
for different AoA.

1. Flutter Predictions of Updated FEM and Updated ACF

Figure 21: Variation of the flutter dynamic pressure with AoA for FEM v20 using both linear and nonlinear analysis, the updated ACF
are only used to modify the steady aerodynamic pressure.

Flutter dynamic pressure for FEM v20 was calculated with the Boeing updated steady ACF applied
to the steady aerodynamic pressure, as shown in Fig. 21. The flutter dynamic pressure for FEM v20
increases linearly with AoA using both linear and nonlinear analysis with updated ACF only applied to the

19 of 22

American Institute of Aeronautics and Astronautics


steady aerodynamic pressure. It was found that the flutter dynamic pressure decreases as the Mach number
increases, which agrees with the transonic ”dip” often seen in the relationship of flutter dynamic pressure
with Mach number, caused by the aft shift of the Cp in a transonic flow.

2. Comparisons between Theoretical Predictions and Test Results


Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

Figure 22: Comparison of the experimental results with the theoretical predictions for FEM v20 using nonlinear analysis with the
updated ACF applied to the steady aerodynamic forces only

Wind tunnel tests of the scaled SUGAR TBW WTM, for which the test results are presented here, was
finished during the month of February 2014. Some open-loop flutter test results were obtained for -3, -1,
1 and 3 degrees AoA at Mach numbers ranging from 0.70 to 0.79. Comparisons between the experimental
results and theoretical predictions are shown for the latest FEM version of the SUGAR TBW WTM (FEM
v20)6 are shown in Fig. 22
Figure 22 shows the comparison between the theoretical predictions and the experimental results for the
latest finite element model, FEM v20. The flutter dynamic pressure was calculated with both the original
ACF and the updated ACF applied only to the steady aerodynamic forces and using a nonlinear analysis.
No ACF was applied to the unsteady aerodynamic forces in the flutter analysis. It was observed from the
test results that the flutter speed is sensitive to the AoA regardless of Mach number, which is similar with
those obtained from the theoretical predictions as shown in Fig. 21. Flutter dynamic pressure calculated
with the updated ACF are less than those obtained using the original ACF by 2–10 psf (3%–13%) depending
on AoA for the same analysis type.
It is also seen that there exists a ”dip” in the experimental results of flutter dynamic pressure with Mach
number for each AoA. However, no flutter ”dip” was found in the theoretical predictions in the given test
flight envelope regardless of the analysis type. Theoretical predictions for FEM v20 using nonlinear analysis
with both the original ACF and the updated ACF applied to the steady aerodynamic forces did not match

20 of 22

American Institute of Aeronautics and Astronautics


well with the experimental results. Once again, it is felt that the unsteady ACF, including time lag correction
factors, to correct DLM should be investigated to capture the transonic effects in transonic flow.

VI. Conclusion
We have described an iterative solution scheme to perform nonlinear aeroelastic analysis including struc-
tural nonlinearities and in-plane loads due to both the aerodynamic loads and the self-weight. A NASTRAN
based solution sequence was employed to solve this nonlinear aeroelastic problem. The impact of transonic
aerodynamic correction factors, ballast mass, and AoA on flutter speed predictions were conducted. After
that, nonlinear aeroelastic analyses of the SUGAR TBW WTM with updated finite element models and the
Boeing updated steady ACF were performed. Comparisons of theoretical predictions and test results were
presented and discussed.
It was found that pre-stressed flutter speed of the TBW WTM is a linear function of AoA using both
linear and nonlinear analysis regardless of analysis type. The analysis has shown the importance of including
structural in-plane loads and geometric nonlinearity when predicting the flutter behavior of nonconventional
truss-braced wing aircraft. In the case of the TBW WTM studied here, for high AoA the nonlinear analysis
Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

predicts an increased flutter margin of 20%–30%, or in some cases, does not predict any flutter instability
at all. The linear analysis, in general, doesn’t provide the same prediction as the nonlinear analysis, and
the variation of the flutter dynamic pressure with Mach number doesn’t agree with the experimental results.
Therefore the linear analysis may not be an adequate prediction for this TBW example. The values for
linear analysis typically fell in between the unloaded results and the nonlinear ones.
The inclusion of ACF for unsteady aerodynamic pressure can increase the flutter margin prediction by
20%–30% depending on AoA. However, we do not recommend this practice. Study also shows that steady
ACF were found not to be appropriate to be used as ACF for the unsteady aerodynamic pressure in the
flutter analysis.
Parametric studies show that an strong sensitivity to ballast mass was demonstrated by evaluating several
ballast mass cases. It was shown that even small changes of 3.36% change to the TBW WTM weight can
result in large changes to the flutter speed of around 20%.
The physical mechanisms influencing the flutter boundary were investigated. It was shown that the
inclusion of pre-stress can result in changes to the critical mode frequencies when using linear analysis, and
both natural frequencies and mode shapes in the nonlinear analysis, and that in turn changes the flutter
boundary. The natural frequency of the critical mode, the third mode passes through the zero damping line
in the V − g plot for the case studied here, increased by around 2.5% compared to that of the unloaded
vehicle up to 5 degrees AoA.
Comparisons of theoretical predictions for the latest FEM v20 and test results were conducted. The
updated steady ACF decrease the flutter boundary compared to that calculated using the original steady
ACF. Test results show that the flutter speed is sensitive to the AoA regardless of Mach number. The
same phenomena was found in the theoretical prediction. However, the variation of flutter dynamic pressure
with Mach number for each AoA does not agree well with the experimental results. Comparisons between
the theoretical predictions for FEM v20 and the experimental results show that there was no flutter ”dip”
predicted in the calculated results. The flutter predictions for FEM v20 didn0 t match quite well with the
experimental results.
The jury member in the TBW WTM studied here has weak stiffness and light weight that connects
the inboard wing and the strut. The jury connecting the main wing and the strut that restrains the wing
deflection, changes the load conditions of the TBW WTM. So, the jury can change the TBW WTM vibration
modes and hence influence the aeroelastic behavior. Besides, the connections can introduce induced drag
to the TBW WTM, which influence the aerodynamic forces acting on the TBW WTM. The future work
can consider the influence that the jury and the induced drag at the jury connections can have on the
flutter boundary. Steady aerodynamic forces obtained from the corrected DLM can be easily realized by
utilizing the ACF, which can be obtained from the test or CFD analysis. Studies are also needed to calculate
the appropriate aerodynamic correction factors that must be applied to the unsteady aerodynamic forces
predicted by DLM.

21 of 22

American Institute of Aeronautics and Astronautics


Acknowledgments
Funding for this effort is provided by the NASA Fundamental Aeronautics Program, through the Subsonic
Fixed Wing organization. The contract number is NNL08AA16B task order NNL11AA00T. The authors
wish to thank SUGAR team members from Boeing Research and Technology, Boeing Commercial Airplanes,
Virginia Polytechnic Institute and State University, and NextGen Aeronautics. The authors would also like
to thank Erik Olson of the NASA Langley Research Center for his guidance as the NASA Contracting Officer
Technical Representative (COTR), and the task technical advisors (TAs): Mark Guynn, Doug Wells, Gerry
Brown, Bob Bartels, William Milholen, Rob Scott, and Jim Felder.

References
1 Bradley,
M. K., and Droney, C. K. ”Subsonic Ultra Green Aircraft Research,” NASA/CR-2011-216847, 2011.
2 Gur,
O., Bhatia, M., Schetz, J. A., Mason, W. H., Kapania, R. K., and Mavris, D. N. ”Design Optimization of a Truss-
Braced Wing Aircraft,” 9th AIAA Aviation Technology, Integration, and Operations Conference, Hilton Head, SC, AIAA
Paper. Vol. 7114, 2009.
3 Meadows, N. A., Schetz, J. A., Kapania, R. K., Bhatia, M., and Seber, G. ”Multidisciplinary Design Optimization of
Downloaded by VIRGINIA TECH on November 4, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-1173

Medium-Range Transonic Truss-Braced Wing Transport Aircraft,” Journal of Aircraft, Vol. 49, No. 6, 2012, pp. 1844-1856.
4 Bhatia, M., Kapania, R. K., and Haftka, R. T. ”Structural and Aeroelastic Characteristics of Truss-Braced Wings: A

Parametric Study,” Journal of Aircraft, Vol. 49, No. 1, 2012, pp. 302-310.
5 Coggin, J. M., Kapania, R. K., Zhao, W., Schetz, J. A., and Hodigere-Siddaramaiah, V. ”Nonlinear Aeroelastic Analysis

of a Truss Braced Wing Wind Tunnel Model.” 55th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and
Materials Conference, AIAA 2014-0335
6 Kapania, R. K., Zhao, W., Schetz, J. A., Coggin, J. M., Hodigere-Siddaramaiah, V., Allen, T. J., and Sexton, B. W.

”Nonlinear Aeroelastic Analysis of SUGAR Truss Braced Wing (TBW) Wind Tunnel Model (WTM).” Submitted to NASA
report, 2014.
7 Friedmann, P. P. ”Renaissance of aeroelasticity and its future,” Journal of Aircraft, Vol. 36, No. 1, 1999, pp. 105-121.
8 Dowell, E., Edwards, J., and Strganac, T. ”Nonlinear aeroelasticity,” Journal of Aircraft, Vol. 40, No. 5, 2003, pp.

857-874.
9 Demasi, L., and Livne, E. ”Aeroelastic coupling of geometrically nonlinear structures and linear unsteady aerodynamics:

Two formulations,” Journal of Fluids and Structures, Vol. 25, No. 5, 2009, pp. 918-935.
10 Demasi, L., Cavallaro, R., and Mrquez Razn, A. ”Postcritical Analysis of Prandtl Plane Joined-Wing Configurations,”

AIAA Journal, Vol. 51, No. 1, 2012, pp. 161-177.


11 Livne, E. ”Aeroelasticity of Joined-Wing Airplane Configurations: Past Work and Future Challenges−A Survey,” 42nd

AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, April 2001, A01-25399.
12 van Schoor, M. C., and von Flotow, A. H. ”Aeroelastic characteristics of a highly flexible aircraft,” Journal of Aircraft,

Vol. 27, No. 10, 1990, pp. 901-908.


13 Patil, M. J., Hodges, D. H., and S. Cesnik, C. E. ”Nonlinear aeroelastic analysis of complete aircraft in subsonic flow,”

Journal of Aircraft, Vol. 37, No. 5, 2000, pp. 753-760.


14 Patil, M. J., Hodges, D. H., and S. Cesnik, C. E. ”Nonlinear aeroelasticity and flight dynamics of high-altitude long-

endurance aircraft,” Journal of Aircraft, Vol. 38, No. 1, 2001, pp. 88-94.
15 Tang, D., and Dowell, E. H. ”Experimental and theoretical study on aeroelastic response of high-aspect-ratio wings,”

AIAA Journal, Vol. 39, No. 8, 2001, pp. 1430-1441.


16 Bartels, R. E., Scott, R. C., Funk, C. J., Allen, T. J., and Sexton, B. W. ”Computed and Experimental Flutter/LCO

Onset for the Boeing Truss-Braced Wing Wind Tunnel Model ” 44th AIAA Fluid Dynamic Conference. June, 2014.
17 Sulaeman, E., Kapania, R. K., and Haftka, R. T. ”Effect of Compressive Force on a Strut-Braced Wing Response,” 42th

AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, AIAA 2001-1611


18 Aeroelastic Analysis User’s Guide MSC Nastran Version 68: MSC software.

22 of 22

American Institute of Aeronautics and Astronautics

View publication stats

You might also like