Goizueta
Goizueta
Goizueta
Paper for Special Session: Progress of the 3rd Aeroelastic Prediction Workshop Large Deflection Group
Two different nonlinear aeroelastic tool sets, SHARPy and the Modal Rotation Method (MRM),
have been employed to predict and design a wind tunnel flutter test campaign of a very flexible
wing, the Pazy Wing, as part of the 3rd Aeroelastic Prediction Workshop. The first method,
SHARPy, uses geometrically exact beams coupled with an Unsteady Vortex Lattice, which is
linearised about a deformed configuration, reduced by means of Krylov subspaces and analysed
to compute the stability boundaries of the wing. The MRM is based on structural modal data,
from either beam models or finite element models, coupled with a doublet-lattice aerodynamic
model from ZAERO of the straight wing configuration. The excellent agreement between
numerical and experimental data for structural-only and static aeroelastic analyses paves the
way for predicting the stability boundaries of the Pre-Pazy wing with sufficient confidence for
the safe design of a flutter wind tunnel test campaign.
I. Introduction
olar-powered, high-altitude, long endurance aircraft (HALE) have become a reality in the aerospace industry with
S the development of aircraft such as Airbus’ Zephyr [1], Boeing’s Odysseus [2], BAE Systems’ PHASA 35 [3] or
Skydweller [4]. All these concepts seek perpetual endurance, stratospheric flight for communication or surveillance — in
essence, pseudo-satellite — missions. Perpetual, year-round endurance particularly imposes constraining requirements
on the design, demanding extreme aerodynamic efficiency and very lightweight structures. The former leads to designs
with very high aspect ratio, that coupled with modern lightweight materials result in wings that are capable of achieving
deformations comparable to their span.
The large deformations induced by the aerodynamic forces make it indispensable to design the system from an
aeroelastic perspective, with tools and methods capable of capturing nonlinear effects caused by such large deformations.
This was a clear outcome of the Helios mishap in 2003 [5], yet is still an area where active research is required, as
shown by the Google Solara crash in 2015 [6].
Although full three-dimensional finite element models and computational fluid dynamics solvers that apply full
Navier-Stokes equations can be coupled together to offer a complete aeroelastic solution, the computational cost is
prohibitive. Thus, the need exists to turn to medium-fidelity models that can offer sufficient accuracy at a fraction
of the cost, in particular if control-related applications, which are highly sensitive to model size, are desired down
the line. Therefore, provided that structural members are often very slender in these aircraft, one-dimensional beam
approximations are an attractive alternative for the structural model. In addition, the laminar, mostly-attached nature of
the flow and the prevalence of thin lifting surfaces make potential flow based aerodynamics a viable option [7].
However, as with all computational methods, and in particular for these medium-fidelity methods, the question
remains as to how accurate are they and whether they are able to capture complex phenomena. Few aeroelastic benchmark
cases for which experimental data is available exist and thus most comparisons between methods have been purely
numerical [8]. In terms of experimental flutter data sets, the Dowell wing [9] in 2001 is a notable example. However, the
data available is reduced and limited to a few angles of attack. Recently, a novel benchmark case has been presented by
the research group at Technion [10] of a very flexible clamped wing for which a flutter test campaign has been planned.
∗ PhD Student, CAGB 308, South Kensington Campus. (norberto.goizueta13@imperial.ac.uk)
† Senior Lecturer in Control, CAGB 340, South Kensington Campus. (a.wynn@imperial.ac.uk)
‡ Professor of Computational Aeroelasticity, CAGB 338, South Kensington Campus. AIAA Associate Fellow (r.palacios@imperial.ac.uk)
§ PhD Student, Faculty of Aerospace Engineering, Technion, Haifa 32000, Israel. (arikdra@gmail.com)
¶ Professor, Faculty of Aerospace Engineering, Technion, Haifa 32000, Israel. AIAA Associate Fellow (daniella@technion.ac.il)
1
This wing, named the Pazy wing, consists of an aluminum spar with Nylon, evenly-spaced, NACA0018-shaped ribs and
a wing tip rod. To give it the external aerodynamic shape, the wing is covered with an Oralight skin.
This paper aims to compare and assess the performance of two medium-fidelity aeroelastic models — SHARPy [11]
and a Modal Rotation Method (MRM) [12] — against the currently available experimental results and provide predictions
for flutter that may be used in the design of further wind tunnel experiments. We present in Sec. II both methodologies,
which are applied to the Pazy wing that is thoroughly described in Sec. III alongside the relevant approximation models.
Section IV compares the numerical models against the static experimental results. The strong correlation between both
sets paves the way for a on-going flutter prediction exercise, discussed in Sec. V.
This activity is part of the Large Deflection working group of the 3rd Aeroelastic Prediction Workshop (AePW3) led
by NASA. This project aims to bring together different numerical and experimental knowledge from academia to solve
relevant and modern aeroelastic problems that may be used as benchmark for validation of future tool sets. For such
purpose, we provide all models, scripts and codes used to obtain the results presented herein available as open source.
Details on how to obtain them can be found in the Research Data Appendix.
A. SHARPy
SHARPy (Simulation of High Aspect Ratio aeroplanes and wind turbines in Python) is a nonlinear aeroelastic simulation
toolbox available under open-source licence [11]. At its core, SHARPy couples a nonlinear beam solver with an
unsteady aerodynamic model based on potential flow assumptions. Additional modules focus on the linearisation of
these nonlinear models around nonlinear static equilibria and model reduction. SHARPy has tens of modules that
can be combined to offer various analysis capabilities and interested readers can find details in the extensive online
documentation∗ . For brevity, we will only summarize here those modules that need to be employed to predict flutter of
highly-flexible, clamped structures. For this we will (i) solve for the static aeroelastic equilibrium condition of the wing;
(ii) linearize the structural and aerodynamic systems about this reference; (iii) reduce the system using modal reduction
and Krylov-based methods, respectively; and (iv) analyze the stability of the resulting reduced linear state-space.
The structural model is based on a geometrically-exact 1D beam formulation, with linear constitutive relations
and nonlinear velocity and displacement kinematic relations. The formulation is parametrised in displacements and
rotations and applied by discretizing the beam in quadratic (3-node) finite elements [13, 14]. The nonlinear equations
that result from the application of Hamilton’s principle take the form of [8]
where M is the mass matrix and 𝑸 represents the discrete external (aerodynamic), gyroscopic and stiffness forces. The
flexible degrees of freedom expressed in a body-attached frame, 𝜼 ∈ R6×(𝑛nodes −1) , include displacements and rotations,
the latter parametrised through a Cartesian rotation vector (CRV).
These equations can be linearised about a reference equilibrium condition (𝜼0 , 𝜼¤ 0 ) where 𝜼¥ 0 = 0, under the
assumption of small perturbations to the flexible degrees of freedom [15]. The resulting equations
adopt the familiar form, where C and K are the constant, tangent damping and stiffness matrices, respectively. The
system can then be projected onto modal coordinates (which will be dependent of the equilibrium point) and truncated
to retain the modes that capture the most significant dynamics.
In SHARPy, the aerodynamics are solved using an Unsteady Vortex Lattice Method (UVLM) which is based on
the assumption of potential flow [7, 8, 16, 17]. Vortex panels are laid out over lifting surfaces, with their spanwise
location coincident with that of the underlying structural elements. In turn, the wake can be modelled by either infinitely
long horseshoe vortices (in static simulations) or with discrete panels that are influenced by the bound and other wake
vortices (for unsteady simulations). The flow field itself is solved enforcing the non-penetration boundary condition
∗ http://imperial.ac.uk/aeroelastics/sharpy
2
at the centre of the bound vortex rings by adjusting their circulation strength.The formulation is nonlinear given the
dependency of the aerodynamic influence coefficients (which contain the information on how each vortex affects others)
on the instantaneous deformation of the structure and shape of the wake sheet. Once the circulation of the vortices has
been solved for, the aerodynamic forces are calculated attending to steady and unsteady contributions: the former is
calculated using the Joukowsky theorem [18] whereas the latter uses Bernoulli’s unsteady equation based on the time
derivative of the circulation, which is calculated by finite differences. An important underlying assumption is that all
viscous effects are confined to thin boundary layers whose effects can be neglected, including viscous drag contributions
and flow separation [7]. Stall has not been modeled.
The linearization of the UVLM equations is performed analytically about the deformed shape under the assumption
of constant aerodynamic influence coefficients (i.e. small structural deformations) and a frozen, yet arbitrary, wake
shape [19]; the linearization includes steady load effects [20]. To achieve convergence and capture unsteady effects a
highly discretized bound vortex lattice and a long wake are required [19, 21, 22], thus it is highly convenient to reduce
the system to alleviate the computational burden of linear analysis tools (i.e computing eigenvalues, etc.). Thus, we
turn to Krylov subspace methods to reduce the dimension of the aerodynamic system [23]. This is a computationally
efficient model order reduction method based on matching the transfer functions and their derivatives at user-defined
frequencies, which by the nature of the UVLM method we limit to reduced frequencies below approximately 0.6 [7].
Nonetheless, since this reduction method is based on transfer function matching, the reduced order system is prone to be
of significant size when the input/output dimensionality is large and this is the case of the linear UVLM, where the
inputs (displacement and velocities) and outputs (forces) are defined at each of the lattice vertices. Therefore, under
the assumption of chordwise rigid rotations of the beam elements, the UVLM input/output space is projected onto the
modal coordinates of the underlying beam element, which reduces the number of inputs and outputs to the number
of chosen structural modes. Then, the Krylov subspace reduction can be performed efficiently, leading to substantial
model size reductions: full order, converged UVLM systems are typically in the range of O (105 ) states and they are
reduced to a dimension of O (10) states.
The coupling between the linear structural and aerodynamic models is trivial, since the inputs and outputs of both
systems are expressed in the structural modal coordinates. The aeroelastic system is also of a sufficiently small size that
stability and other linear analyses can be performed very efficiently.
and their summation over the structure (over the reference line) using nonlinear kinematics:
where [𝑅𝑖 ] is a rotation matrix of the 𝑖-th segment which is based on the summation of the local curvatures (represented
via Euler angles) from the root to the segment. To accurately model the internal bending moments, the MRM applies an
iterative procedure that updates the location and orientation of the applied external loads based on the deformed shape
via external correction moments:
𝑛
Õ
{Δ 𝑀𝑖 } = {𝑀loc𝑖 } − {𝑀0𝑖 } − {Δ 𝑀 𝑗 } (5)
𝑗=𝑖+1
where {𝑀0𝑖 } is the internal moment acting on the undeformed configuration. The MRM is coupled with an aerodynamic
model, based on the rigid configuration, to compute static aeroelastic deformations and the flutter velocity about the
static equilibrium. Two methods are proposed to update the structural data of the deformed structure and to alter the
3
Fig. 1 Illustration of the segmentation and two arbitrary reference lines used in the MRM [12].
unsteady aerodynamic matrices, and to obtain a flutter solution of the deformed wing. One is based on discrete mass
and aerodynamic properties of the wing and the other only on modal data, thus providing a nonlinear solution that is
based only on linear modal data. The curvature based representation yields a constant generalized stiffness matrix and
the updating procedure of the mass and aerodynamic matrices is done by representing the deformed mode shapes as a
combination of the basic mode shapes
𝑚
Õ
{𝜙deformed }𝑖 ≈ 𝐺 𝑖 𝑗 {𝜙} 𝑗 (6)
𝑗=1
or
[𝜙deformed ] ≈ [𝜙] [𝐺] , (7)
thus yielding an expression for the deformed mass matrix that is based directly on the modal mass matrix of the
undeformed structure:
[𝐺 𝑀deformed ] = [𝐺 > ] [𝜙> ] [𝑀] [𝜙] [𝐺] = [𝐺 > ] [𝐺 𝑀] [𝐺]
| {z } | {z }
[ 𝜙> ] [𝜙 ] . (8)
deformed deformed
For the update of the deformed aerodynamic matrices, we assume that, locally, the relation between an incremental
deformation (namely, the local incremental angle of attack) and the local forces do not change with the large deformation.
This is equivalent to the strip theory assumption that the local strip characteristics are unaffected by the deformation. It
is noted that the baseline model does not have to be generated by a strip theory. Under this assumption, updating the
aerodynamic matrix is done similarly to the update of the mass matrix, yielding:
[𝑄 𝐻 𝐻 ,deformed (𝑖𝑘)] = [𝐺 > ] [𝜙> ] [ 𝐴𝐼𝐶 (𝑖𝑘)] [𝜙] [𝐺] = [𝐺 > ] [𝑄 𝐻 𝐻 (𝑖𝑘)] [𝐺]
| {z } | {z }
>
[ 𝜙deformed ] [ 𝜙deformed ] (9)
where 𝐴𝐼𝐶 (𝑖𝑘) is the discrete aerodynamic influence coefficient matrix, assumed to be unknown explicitly, and
𝑄 𝐻 𝐻 (𝑖𝑘) is the generalized aerodynamic coefficient matrix. Given the deformed, linearized stiffness, mass, and
aerodynamic matrices, flutter can be evaluated with any conventional flutter solution technique, such as the 𝑘-method
used in the current case.
4
covered with Polyester foil which is typically used in radio-controlled drones. A 300 mm long and 10 mm diameter
wing-tip rod is 3D printed as part of the chassis. The rod is used for attaching weights (via drilled holes) that can modify
the dynamic properties of the structure and alter the flutter speed. Overall, the wing weighs 0.32 kg without the base
with which it attaches to the wind tunnel floor.
B. Model Reduction
SHARPy’s structural module requires of one-dimensional beams represented by 6-by-6 stiffness and mass matrices. To
obtain these, the full 3D FEM can be loaded in several directions to obtain the relevant stiffness coefficients that are
applicable to one-dimensional beams. This has been done in [26], where Riso et al. developed the beam model for its
use in UM/NAST (University of Michigan Nonlinear Aeroelastic Simulation Toolbox). The UM/NAST baseline model
has been adapted for SHARPy since in SHARPy the stiffness matrix is a 6-by-6 matrix as opposed to UM/NAST’s
4-by-4. The difference being in the shear terms which have been modelled in the SHARPy model as with a large
number to simulate zero compliance, although the vibration characteristics are not very sensitive to this term. Both the
UM/NAST and SHARPy models share the same beam reference line, located at 44.1% chord.
5
The Pazy wing’s Nylon ribs have a NACA0018 airfoil shape which is symmetric, therefore modeled as a flat plate of
vortex panels in SHARPy. For static simulations, the wake is modeled by infinitely long horseshoe vortices whereas,
for dynamic ones, the circulation shed by the bound vortices is convected downstream from the deformed wing and
retained, resulting in a wake sheet of discrete vortex panels of constant circulation that keep a history of the aerodynamic
unsteady effects. The aerodynamics of the skin-on and skin-off models are identical.
For consistency between the baseline models, the MRM structural model was based on a coupled 6-by-6 beam
model built in Matlab with the stiffness coefficients from [26]. Twenty mode shapes were extracted from 100 FE nodes
along the beam and interpolated to 400 segments along the reference line over the wingspan. A similar analysis based
directly on the FE model, without the beam reduction can be found in [24]
The MRM aerodynamic model is based on the doublet-lattice method. A rigid aeroelastic analysis was conducted in
Zaero [27]. For the static aeroelastic analysis the integrated forces per strip were used to generate a strip model, with the
aerodynamic normal force coefficient slope and center of pressure location varying along the span as shown in Fig. 4.
For the flutter analysis, the generalized aerodynamic coefficient matrices, [𝑄 ℎℎ ], at several reduced frequencies of the
undeformed structure were exported from Zaero and used as a database for the flutter analysis of the deformed wing .
Fig. 4 Span-wise distribution of the sections’ normal force coefficient slope and center of pressure locations,
computed from Zaero for the undeformed structure.
C. Modal Analysis
A modal analysis of the unloaded Pazy wing has been performed to validate the sectional coefficients of the equivalent
beam models against the full 3D FEM. Tables 1 and 2 show the original FEM, UM/NAST [26], SHARPy and MRM
frequencies. Since the SHARPy and MRM models are derived from the UM/NAST model, the relative error between
them and UM/NAST is shown. With a maximum error of 0.7% between SHARPy and UM/NAST models in the 3rd
out-of-plane (OOP) bending mode, the SHARPy model can be considered acceptable, as can the MRM with a maximum
1.96% difference in the 1st torsion mode. Finally, excellent agreement is observed between the full 3D FEM and the
beam models’ frequencies.
6
Table 1 Modal frequencies comparison with the skin fitted.
0.0 0.0
SHARPy
Wing tip vertical displacement, m
UM/NAST
0.1 0.1 MRM
Nastran FEM
Technion Experimental
0.2 0.2
SHARPy
UM/NAST
0.3 MRM 0.3
Technion Experimental
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Wing tip load, kg Wing tip load, kg
(a) Skin on (b) Skin off
Fig. 5 Wing tip, mid-chord vertical displacement comparison under a dead load between SHARPy, MRM,
UM/NAST and Technion’s experimental results.
7
Table 3 Linear bending curve slope and intercept for a linear regression coefficient of 0.999 for the bending
test up to 1 kg load.
SHARPy SHARPy
Wing tip vertical displacement, m
0.2 0.2
0.3 0.3
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Tip Load, kg Tip Load, kg
(a) Skin on (b) Skin off
Fig. 6 Wing tip, mid-chord vertical displacement comparison under a twisting dead load between SHARPy,
MRM, UM/NAST and Technion’s experimental results.
8
SHARPy 0.0 SHARPy
0.0 MRM MRM
Technion Experimental 2.5 Technion Experimental
Twist angle, degrees
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Tip Load, kg Tip Load, kg
(a) Skin on (b) Skin off
Fig. 7 Wing tip twist angle under a tip load (negative twist indicates leading edge down).
Angle of attack 5◦ , 7◦
Free stream velocity 1 − 60 m/s
Density 1.225 kg/m3
Chordwise panels 16
Spanwise panels 32
Wake panels Infinite horseshoe
Figure 8 shows the wing tip out-of-plane displacements expressed in an inertial frame of reference with origin at the
wing root and aligned with the flow (the structure rotates to simulate the angle of attack). The displacement shows a
linear relationship with dynamic pressure below 1000 Pa (40.4 m/s) and in this region the results from all computational
models and the experimental results are in good agreement. Once the geometrical nonlinearities become more apparent,
the MRM displays a slightly more compliant response with the displacement approximately 5% above that of SHARPy
at 60 m/s. The number of experimental data points gathered at high speeds is not sufficient to draw any conclusive
comments on the fidelity of aeroelastic models, although the general trend is enough to have confidence in the results
obtained at these speeds.
V. Dynamic Results
The previous results show good agreement between the SHARPy model, MRM, the FEM and the experimental results
for static cases. To verify the inertia properties of the model, we will compare the vibration modes of the deformed wing
under a follower force. Then, we proceed with a flutter prediction analysis for the wing in the wind tunnel fixed in a
vertical position at various angles of attack to evaluate the variation in flutter speed. For these analysis, we additionally
consider the case with a 10 g mass at the trailing edge of the wing tip section to evaluate its impact. This mass is
modeled as a lumped, point mass with no inertia.
9
Velocity, m/s Velocity, m/s
0 20 40 60 0 20 40 60
0.6
Skin off SHARPy
Skin on MRM
Normalised wing tip vertical displacement, -
0.5 UM/NAST
Experimental
0.4
0.3
0.2
0.1
0.0
0 500 1000 1500 2000 2500 0 500 1000 1500 2000 2500
Dynamic Pressure, Pa Dynamic Pressure, Pa
(a) 5 degrees (b) 7 degrees
Fig. 8 Wing tip out-of-plane deflections at various angles of attack and free-stream velocities.
Fig. 9 Linearised beam’s modes of vibration frequencies at a deformed condition when subject to a wing tip
vertical follower force.
10
Figure 9 compares the deformed structural frequencies between SHARPy, MRM, NASTRAN and UM/NAST [29]
for both the skin on and off cases. The structural change of shape, in this case parametrised by the vertical wing tip
deformation, causes two clear mode switches. Of particular importance is the drop in frequency of the 1st torsional
mode, which eventually crosses the 2nd out-of-plane bending mode, and the drop in frequency of the 1st in-plane
mode. The out-of-plane bending modes remain fairly constant in frequency, with the OOP1 seeing a slight increase in
frequency caused by the stiffening terms arising from the linearisation of the wing tip follower force. We note that these
results are only indicative as aerodynamic loading produces different equilibrium shapes to these and thus would skew
the frequency against wing tip displacement relation.
11
Skin off Skin on 50 100
0.2
90
0.1
Damping ratio, -
40 1st Torsion
80
0.0
Imag, Hz
0.2
50 50
1st Torsion 20
40 2nd OOP Bending
40
Frequency, Hz
30
2nd OOP Bending 10 30
20
1st Torsion + IP 1st OOP Bending
10 1st OOP Bending 20
0 0 10
20 30 40 50 60 70 80 90 100 5 4 3 2 1 0 1 2
Velocity, m/s Real, Hz
Fig. 10 Stability of the Pazy wing at zero angle of attack. The frequency-damping-velocity (a) and eigenvalue
plot (b) with eigenvalues highlighted in 10m/s intervals
40 1st Torsion
80
0.0
0.2
50 50
1st Torsion 20
40 2nd OOP Bending
40
Frequency, Hz
30
2nd OOP Bending 10 30
20
1st Torsion + IP 1st OOP Bending
10 1st OOP Bending 20
0 0 10
20 30 40 50 60 70 80 90 100 5 4 3 2 1 0 1 2
Velocity, m/s Real, Hz
Fig. 11 Stability of the Pazy wing at zero angle of attack. The frequency-damping-velocity (a) and eigenvalue
plot (b) with eigenvalues highlighted in 10m/s intervals
12
obtained using SHARPy and Zaero/MRM.
13
= 1.0 = 5.0 MRM = 5 50 100
0.2
90
0.1
Damping ratio, -
40 1st Torsion
80
0.0
Imag, Hz
0.2
50 50
1st Torsion 20
40 40
Frequency, Hz
(a) V-g-omega plot, where colour scale relates to the natural (b) Eigenvalues of the aeroelastic system. Eigenvalues are
frequency of the aeroelastic mode. MRM data shown for the highlighted at 10m/s increments.
5 degree angle of attack case.
Fig. 12 SHARPy stability plot for the skin off wing fitted at 1 and 5 degrees angle of attack.
This is illustrated by the steep “stabilisation boundary” of the first flutter mode, which is fairly constant around 25%
wing tip span wise deformation.
The second flutter speed is also shown to decrease with increased deflection and angle of attack, although this
decrease is not monotonic and appears to even increase beyond 5 degrees angle of attack. These regions for the skin off
model provide the most conservative flutter scenario for a wind tunnel test, and could offer insight into an experiment to
reach the second flutter instability region without excessive deflections and spending minimal time in the first flutter
region.
Turning to the comparison between codes, it is in these deformed, dynamic cases where the difference in aerodynamic
models becomes most obvious. The MRM frequency predictions in Fig. 12a and 13a for the 5 degree angle of attack
show an excellent agreement at the lower velocities, where the deflection is smallest. Beyond 50 m/s, point at which
the deflection reaches 30% of the span, the differences in the T1 and OOP1 start becoming noticeable. This can be
attributed to the change of dihedral effect caused by the deformation, which is captured by the linearised SHARPy
UVLM but not the MRM implementation of the linearised aerodynamic system. This effect has also been previously
observed by the authors on the setting of a different framework [21] that used the UVLM linearised at the straight wing
condition as aerodynamic model. In such framework, it was observed that the OOP1 reduced its frequency to become a
purely real mode as deformation increased, as opposed to using the deformed condition as reference for the linearised
UVLM, case in which the OOP1 frequency increases with deformation as seen here.
The difference is further observed in Fig. 14, where the agreement in the flutter boundary is excellent for the small
deformation conditions (the first flutter instability in Fig. 14a) as opposed to the second flutter mode shown which differs
substantially at the larger wing tip deflections.
VI. Conclusions
We have evaluated and compared between two different aeroelastic toolboxes, SHARPy and Technion’s MRM, the
structural static and dynamic response of the Pre-Pazy wing. The agreement between tools for the different models and
correlation with experimental results give confidence in using these models to support the design of non-destructive
wind tunnel experiments to explore the wing’s flutter mechanisms. The two identified flutter modes’ dependency on the
wing deformation provide an excellent example to showcase the need for nonlinear aeroelastic analysis tools. Both
the MRM and SHARPy capture well these effects, although at high deformations the need for aerodynamic models
linearised at the deformed condition become of paramount importance to predict flutter.
On the other hand, the complex modelling of the Oralight skin introduces some uncertainties into these analyses
which, thus far, have been tackled by using two extreme models. By using skin on and skin off models it is expected that
14
= 1.0 = 5.0 MRM = 5 50 100
0.2
90
0.1
Damping ratio, -
40 1st Torsion
80
0.0
Imag, Hz
0.2
50 1st Torsion
50
20
40 40
Frequency, Hz
(a) V-g-omega plot, where colour scale relates to the natural (b) Eigenvalues of the aeroelastic system with eigenvalues
frequency of the aeroelastic mode. MRM data shown for the highlighted at 10m/s increments.
5 degree angle of attack case.
Fig. 13 SHARPy stability plot for the wing fitted at 1 and 5 degrees angle of attack.
90
1st Torsion + 1st OOP Bending
90 Flutter Region
0.5
0.5 1st Torsion + 1st OOP Bending
80 Flutter Region
1.0
1st Torsion + 2nd OOP Bending 1.0 80
Flutter Region 1st Torsion + 2nd OOP Bending
Flutter Region
70
Velocity, m/s
Velocity, m/s
2.0
2.0 70 3.0
3.0 4.0
4.0
60
5 60 0.25
0.2 5.0
5.0
50 50
SHARPy Flutter Boundary SHARPy Flutter Boundary
MRM Flutter Boundary MRM Flutter Boundary
40 40
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.0 0.1 0.2 0.3 0.4 0.5 0.6
Span-normalised wing tip vertical displacement, - Span-normalised wing tip vertical displacement, -
Fig. 14 Flutter boundaries and regions, with contours showing wing tip deflection at a given root angle of
attack.
15
the results show the two limits between which the flutter boundary is expected in the wind tunnel experiments.
Further work will focus on the analysis of an updated computational model of the Pazy wing, the comparison against
dynamic wind tunnel tests and other numerical tools that also consider the Pazy example [29, 30].
Research Data
With the objective of making this a benchmark case that can be used to validate and test future aeroelastic simulation
tools, the SHARPy model, scripts and results are all made available online. The results presented herein can be
downloaded from the Zenodo archive.
URL
AePW3 Results https://doi.org/10.5281/zenodo.4311606
SHARPy simulation scripts https://doi.org/10.5281/zenodo.4299819
Acknowledgements
The authors would like to gratefully acknowledge the Pazy Foundation for their support of the Pazy-Wing experimental
campaign.
The authors would also like to thank Markus Ritter from the DLR for leading and organising the Large Deflection
working group within the 3rd Aeroelastic Prediction Workshop.
Norberto Goizueta would like to thank Mark Korchma from Imperial College for deriving the original equivalent
beam model for SHARPy that started this work and Cristina Riso for her help providing the UM/NAST model.
The Department of Aeronautics at Imperial College has sponsored Norberto Goizueta’s research under EPSRC
grant EP/R513052/1.
References
[1] “Airbus Zephyr Solar High Altitude Pseudo-Satellite flies for longer than any other aircraft during its successful maiden
flight,” , 2018. URL https://www.airbus.com/newsroom/press-releases/en/2018/08/Airbus-Zephyr-Solar-
High-Altitude-Pseudo-Satellite-flies-for-longer-than-any-other-aircraft.html, Accessed 05/06/2020.
[2] “Odysseus - Global Reach, Airborne for Months, Powered by the Sun,” , 2020. URL https://www.aurora.aero/odysseus-
high-altitude-pseudo-satellite-haps/, Accessed 05/06/2020.
[3] “Persistent High Altitude Solar Aircraft (PHASA-35®) has the potential to stay airborne for a year,” ,
2020. URL https://www.baesystems.com/en/article/ground-breaking-solar-powered-unmanned-aircraft-
makes-first-flight, Accessed 05/06/2020.
[5] Noll, T., Brown, J., Perez-Davis, M., Ishmael, S., Tiffany, G., and Gaier, M., “Investigation of the Helios Prototype Aircraft
Mishap,” Tech. Rep. January, NASA, 2004. https://doi.org/10.1109/MWSYM.2002.1012130, URL https://www.nasa.gov/pdf/
64317main_helios.pdf.
[6] d’Oliveira, F. A., de Melo, F. C. L., and Devezas, T. C., “High-Altitude Platforms - Present Situation and Technology Trends,”
Journal of Aerospace Technology and Management, Vol. 8, No. 3, 2016, pp. 249–262.
[7] Murua, J., Palacios, R., and Graham, J. M. R., “Applications of the unsteady vortex-lattice method in aircraft aeroelasticity and
flight dynamics,” Progress in Aerospace Sciences, Vol. 55, 2012, pp. 46–72. https://doi.org/10.1016/j.paerosci.2012.06.001.
[8] del Carre, A., Teixeira, P., Palacios, R., and Cesnik, C. E., “Nonlinear response of a very flexible aircraft under lateral gust,”
International Forum on Aeroelasticity and Structural Dynamics 2019, IFASD 2019, 2019, pp. 1–27.
[9] 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. https://doi.org/10.2514/2.1484, URL https://doi.org/10.2514/2.1484.
16
[10] Avin, O., Drachinsky, A., Ben-Shmuel, Y., and Raveh, D. E., “Design of an Experimental Benchmark of a Highly Flexible
Wing,” 60th Israel Annual Conference on Aerospace Sciences, 2020, pp. 1–18.
[11] Carre, A., Muñoz-Simón, A., Goizueta, N., and Palacios, R., “SHARPy : A dynamic aeroelastic simulation toolbox
for very flexible aircraft and wind turbines,” Journal of Open Source Software, Vol. 4, No. 44, 2019, p. 1885. https:
//doi.org/10.21105/joss.01885.
[12] Drachinsky, A., and Raveh, D. E., “Modal Rotations: A Modal-based Method for Large Structural Deformations,” AIAA
Journal, 2020. https://doi.org/10.2514/1.J058899.
[13] Geradin, M., and Cardona, A., Flexible multibody dynamics: a finite element approach, John Wyley and Sons, Chichester, 2001.
[14] Simpson, R. J. S., and Palacios, R., “Numerical aspects of nonlinear flexible aircraft flight dynamics modeling,” 54th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, 2013, pp. 1–25. https://doi.org/10.
2514/6.2013-1634, URL http://arc.aiaa.org/doi/10.2514/6.2013-1634.
[15] Hesse, H., Palacios, R., and Murua, J., “Consistent structural linearisation in flexible aircraft dynamics with large rigid-body
motion,” AIAA Journal, Vol. 52, No. 3, 2014, p. 528. https://doi.org/10.1016/j.compstruc.2012.05.011.
[16] Katz, J., and Plotkin, A., “Unsteady Incompressible Potential Flow,” Low Speed Aerodynamics, Cambridge University Press,
2001, Chap. 13, 2nd ed., pp. 369–433.
[17] Del Carre, A., and Palacios, R., “Efficient Time-Domain Simulations in Nonlinear Aeroelasticity,” AIAA Scitech Forum, 2019,
pp. 1–20. https://doi.org/10.2514/6.2019-2038.
[18] Simpson, R. J. S., Palacios, R., and Murua, J., “Induced-Drag Calculations in the Unsteady Vortex Lattice Method,” AIAA
Journal, Vol. 51, No. 7, 2013, pp. 1775–1779. https://doi.org/10.2514/1.j052136.
[19] Maraniello, S., and Palacios, R., “State-Space Realizations and Internal Balancing in Potential-Flow Aerodynamics with
Arbitrary Kinematics,” AIAA Journal, Vol. 57, No. 6, 2019, pp. 1–14. https://doi.org/10.2514/1.J058153, URL https:
//arc.aiaa.org/doi/10.2514/1.J058153.
[20] Murua, J., Martínez, P., Climent, H., van Zyl, L., and Palacios, R., “T-tail flutter: Potential-flow modelling, experimental
validation and flight tests,” Progress in Aerospace Sciences, Vol. 71, 2014, pp. 54–84. https://doi.org/https://doi.org/10.1016/j.
paerosci.2014.07.002, URL http://www.sciencedirect.com/science/article/pii/S0376042114000669.
[21] Artola, M., Goizueta, N., Wynn, A., and Palacios, R., “Modal-Based Nonlinear Estimation and Control for Highly Flexible
Aeroelastic Systems,” AIAA Scitech Forum, 2020, pp. 1–23. https://doi.org/10.2514/6.2020-1192.
[22] Maraniello, S., and Palacios, R., “Parametric Reduced-Order Modeling of the Unsteady Vortex-Lattice Method,” AIAA Journal,
Vol. 58, No. 5, 2020, pp. 2206–2220. https://doi.org/10.2514/1.j058894.
[23] Goizueta, N., Wynn, A., and Palacios, R., “Parametric Krylov-based order reduction of aircraft aeroelastic models,” AIAA
Scitech Forum, 2021.
[24] Drachinsky, A., and Raveh, D. E., “Nonlinear Aeroelastic Analysis of Highly Flexible Wings Using the Modal Rotation Method,”
AIAA Scitech Forum, 2021.
[25] Avin, O., Drachinsky, A., Ben Shmuel, Y., and Raveh, D. E., “An Experimental Benchmark of a Very Flexible Wing,” AIAA
Scitech Forum, 2021.
[26] Riso, C., and Cesnik, C. E. S., “Equivalent Beam Distributions of the Pazy Wing,” Tech. rep., University of Michigan, 2020.
[27] Zaero Theoretical Manual V. 9.2, Scottsdale, AZ, 3rd ed., 2017.
[28] Avin, O., Drachinsky, A., Ben-Shmuel, Y., and Raveh, D. E., “An Experimental Benchmark of a Highly Flexible Wing,” AIAA
Scitech Forum, 2021.
[29] Riso, C., and Cesnik, C. E. S., “UM/NAST Experimental Validation Using the Pazy Wing Aeroelastic Benchmark,” AIAA
Scitech Forum, 2021.
[30] Ritter, M., Hilger, J., and Zimmer, M., “Static and Dynamic Simulations of the Pazy Wing Aeroelastic Benchmark by Nonlinear
Potential Aerodynamics and detailed FE Model,” AIAA Scitech Forum, 2021.
17