You are on page 1of 8

1

Aeroelastic Optimisation of an Alternative High


Altitude Long Endurance Wing Stiffening Structure
Nicholas F. Giannelis and Gareth A. Vio

Abstract: This paper considers an alternative internal wing stiffening structure to improve the volumetric efficiency
of a generic High Altitude Long Endurance aircraft wing. Finite element models for this structure and an equivalent
conventional configuration are developed in NASTRAN. A Binary Genetic Algorithm optimisation routine is employed
under transient and dynamic aeroelastic constraints to determine the optimum weight for each model. A quasi-static
discrete gust model and linear flutter analysis is used for the transient and dynamic aeroelastic formulation respectively.
Results find that although qualitatively similar aeroelastic behaviour is observed in both configurations, the alternative
topology is structurally inefficient in high aspect ratio applications.

Introduction
The past decade has seen a substantial increase in research
and development for High Altitude Long Endurance (HALE)
unmanned aerial systems [1]. The obvious driving force for
such systems has been the defence sector, where intelligence
and reconnaissance operations may be implemented economically and without the risk of human casualty [2]. Nonetheless,
commercial and scientific applications, including communication network modules, earth and atmospheric remote sensing
and long endurance flight research, are becoming commonplace [35].
With highly complex, multiobjective mission profiles, the
volumetric efficiency of HALE structures is of pivotal concern.
An effective topology is necessary to house the plethora of subsystems required by military or science specifications. [6] has
explored the improvement of high aspect ratio wing structural
efficiency through the redistribution of wing stiffening members of a conventional wing box configuration, with advantages
observed in both weight and flutter stability of the platform. In
[7], the issue of volumetric efficiency of a stored HALE wing is
addressed, with a skin integrated stiffening structure permitting
a 25-35% wing weight reduction compared to a baseline. An
integrated multi-lobe wing stiffening structure is proposed in
[8], with a reduction of 22% wing weight relative to correlated
weights of advanced transportation systems. This analysis is
furthered in [9], where the flutter stability of the multi-lobe
stiffening structure exceeded an equivalent rib and spar wing
by 13.9%. In performing structural optimisation with the large
parameter sets required by HALE aircraft, global heuristic optimisation methods, such as the Genetic Algorithms used in
[10] and [11], have proven effective in locating globally optimum solutions.
The long endurance aspect of HALE missions necessitates
a high aspect ratio planform to achieve the required aerodynamic performance [12]. Consequently, such vehicles typically exhibit low structural weight with highly flexible lift-

Nicholas F. Giannelis. School of AMME, Building J11, The University of Sydney, NSW, 2006, Australia
Gareth A. Vio. School of AMME, Building J11, The University
of Sydney, NSW, 2006, Australia
Fourth Australasian Unmanned Systems Conference : 18 (2014)

ing surfaces, making them susceptible to significant transient


aeroelastic loads from atmospheric gusts and turbulence [13].
Substantial research efforts have thus been devoted to the modeling of aeroelastic phenomena in high aspect ratio wings [14,
15]. Traditionally, a quasi-static, discrete gust formulation,
based on a worst case one-minus-cosine waveform has been
applied to model transient aeroelastic loads [16]. Statistical
modeling techniques, including a Statistical Discrete Gust and
continuous Power Spectral Density functions have since been
developed [17], however evidence suggests that the worst case
gust is an application of an extreme excursion under the statistical methods [18, 19]. More recently, high fidelity, transient
Computational Fluid Dynamics (CFD) simulations have been
employed in the aeroelastic analysis of HALE vehicles [20].
Although the computational overhead of such methods may
be mitigated through Reduced Order Modeling, as in [21], the
computational effort remains significant compared to a discrete
formulation. A number of recent studies have also focused on
the significance of nonlinearities in the aeroelastic response of
high aspect ratio wings [5, 15, 22], with [22] observing a 50%
reduction in flutter stability when accounting for nonlinear effects. Nonetheless, [23] concluded that although nonlinearities
are significant for HALE aircraft, quasi-static linear analysis
remains a primary tool in the preliminary design phase.
In this paper, an alternative internal wing stiffening structure
with improved volumetric efficiency will be developed for a
generic HALE wing. An equivalent conventional rib and spar
model is also presented as a baseline. The structures will be
optimised for weight under dynamic and transient aeroelastic
constraints through use of a Binary Genetic Algorim (BGA).
The formulation of the optimisation routine, along with linear
aeroelastic system and discrete gust model will be described.
The optimal structures for each configuration will then be compared in terms of structural sizing and aeroelastic performance.

Method
The method is broadly divided into three sections. The first
details the development of the two structural models under a
Finite Element scheme. The aeroelastic formulation applied in
this paper is then presented, with definitions of both the dynamic and transient phenomena to be analysed. The structural
optimisation routine is then described, with focus given to the

Fourth Australasian Unmanned Systems Conference , 2014

optimisation constraints, fitness function and the Binary Genetic Algorithm used.

y/c

Structural Model
The wing planform has been developed to reflect a generic
HALE vehicle structure. A rectangular section of aspect ratio
7 (2.5 m chord and 17.5 m span) represents the base wing planform for both conventional and multi-lobe stiffening configurations. A typical high lift-to-drag aerofoil profile for HALE
applications [7], the NASA LRN1015 shown in Figure 1, gives
the aerodynamic section in both structures.

continuity between the vertical webs and external skin. The


upper and lower surface lobe arcs are also comprised of quadrilateral shell elements and provide the improved volumetric efficiency for the multi-lobe structure. The minimum internal
volume in the alternative configuration is 1.12 m3 , a 65% increase from the 0.68 m3 maximum internal volume of the conventional structure. Each of the lobe arcs, webs and the skin
is again grouped independently for sizing, where the lobe arc
groups are not segmented in the spanwise direction and a single group runs along the entirety of the span.

0.1
0
0.1
0

0.2

0.4

0.6

0.8

x/c
Fig. 1. LRN1015 Aerofoil Profile

The finite element models for analysis are realised in NASTRAN. The external wing skin is identical for both stiffened
configurations, and is represented by quadrilateral shell elements. The conventional model internal structure, given in
Figure 2, employs a rib and spar configuration. Four spars
extend from wing root to tip, with eight equally spaced ribs
stationed along the span, a representation of the Global Hawk
internal wing structure [24]. Both ribs and spars are comprised
primarily of quadrilateral shell elements, with triangular elements used to close the rib sections at the leading edge. This
work does not attempt to optimise the topology of either the
conventional or alternative structure, but rather seeks to ascertain the efficacy of the multi-lobe stiffening structure for high
aspect ratio application. Each spar and rib is grouped individually for sizing of the shell thickness, with the wing skin also
grouped separately.

Fig. 3. Multi-Lobe Internal Structure

In both configurations cantilever boundary conditions are


applied at the wing root to the internal stiffening structure.
Nodes at the wing root are constrained in six degrees of freedom to represent attachment to the fuselage. Aluminium 6061
is used throughout each model, with material properties relevant to the analysis given in Table 1. The base models developed are seeded with a uniform shell thickness of 5 cm for each
structural member, representing the heaviest possible structure
in each case.
Table 1. Al-6061 Material Properties
Property
Elastic Modulus (GPa)
Shear Modulus (GPa)
Poissons Ratio
Density (kg/m3 )

Value
68.9
26.0
0.33
2700

Aeroelastic Formulation
Optimisation of the two wing configurations is performed
under dynamic aeroelastic constraints. Specifically, this paper
considers the flutter instability and the transient wing response
to a quasi-static discrete gust load under quasi-static aerodynamics.

Fig. 2. Conventional Internal Structure

An equivalent alternative wing stiffening structure is given


in Figure 3. The vertical webs are stationed at equivalent chordwise positions as the conventional model and consist entirely
of quadrilateral shell elements. Rigid bar elements provide

Dynamic Aeroelasticity
Under the assumption of linearity, the dynamic aeroelastic
problem is governed by the system described by:




1 2
1
U B + D q +
U C + E q = Q
[1]
A
q+
2
2
where A is the inertial matrix, B and C are the aerodynamic

Transient Aeroelasticity
Converse to the intrinsic flutter instability, the transient aeroelastic response is concerned with the time domain solution due
to non-zero external excitation, namely, a gust load. A discrete
gust disturbance is modeled as the generalised excitation force
in Equation 1 to characterise the transient aeroelastic response
of the linear system. The gust load is deterministic and assumes a vertical disturbance, constant along the wing span with
a small magnitude relative to the freestream velocity. Consequently, the discrete gust velocity acts to increase the effective
angle of attack, and hence the aerodynamic forces experienced
by the wing.
A one-minus-cosine waveform is used to model the magnitude of the gust velocity in this analysis. In the time domain,
the gust amplitude is given by:

Wg (t) =

1
2


Wref

H
12.5

15

1 cos

2t
tg

ti t < tf

[2]

where Wg (t) is the gust amplitude at time t, ti is the gust start


time with a gust duration of tg and end time of tf . Wref is the
reference gust velocity for the flight condition and H is the gust
gradient, which defines the time taken for the gust to achieve
maximum amplitude. FAA regulations FAR 25 recommend a
gust gradient of 12.5c for certification, where c represents the
chord length [20].
In the present analysis, a discrete gust loading at the cruise
condition is considered. For this flight condition Wref = 15.24

0.5

1
1.5
Time (sec)

Fig. 4. Discrete Gust Disturbance Velocity Profile

Optimisation Routine
The objective of this work is to determine the minimum
weight wing configuration that conforms to both a minimum
flutter speed and maximum bending moment constraints. As
such, the problem is formulated as a constrained optimisation, looking to determine the vector of design variables x =
T
(x1 , x2 , ...xn ) that minimises the fitness function f (x) subject to the inequality constraints contained in g(x):
g(x) =

10

1/7 ! 
for

m/s, tg = 0.286 s and H = 31.25 m. The resulting gust input


waveform, beginning at time 1 s, is given in Figure 4.
The transient aeroelastic analysis is performed under NASTRANs 146 Solution, with analogous aerodynamics and flight
conditions to the flutter solution. A time step of 0.01 s is
used for a simulation time of 2.5 s, allowing sufficient time to
observe decay in the wings displacement amplitude. Control
points at the wing tip leading and trailing edges are monitored
for displacement. The FAR 25 regulations specify that under
discrete gust loading, the aircraft be constrained to plunge motion only [26]. As such, the critical parameter categorising
the the transient aeroelastic response of the models is the root
bending moment about the aircraft longitudinal axis. Rigid
body elements are used to connect the constrained wing root
nodes to the centroid of the root aerofoil section, where the
total root bending moment is computed.

Gust velocity (m/s)

and structural damping matrices respectively, D and E are the


aerodynamic and structural stiffness matrices respectively, is
the air density, U is the free stream velocity, q are the generalised displacements, q is the first time derivative, q is the second time derivative, and Q are the generalised external forces
applied to the system. For a given structure, the inertial, damping and stiffness matrices are constant, reducing the stability
of the system to a function of dynamic pressure.
The flutter instability is a product of the interaction of orthogonal modes of vibration. The coupling of elastic modes in
a linear system produces self-excited oscillations of the structure, typically resulting in catastrophic failure. As identification of the flutter condition is concerned with the underlying
stability of Equation 1, the generalised excitation force, Q, is
taken to be zero. The flutter analysis then proceeds through
NASTRANs 145 Solution, with the inertial, damping and stiffness matrices of the system derived from the FE solution. Lanczos eigenvalue extraction is used to determine the first ten mass
normalised eigenvectors and eigenfrequencies for flutter analysis. Subsonic Doublet-Lattice theory is employed to determine the aerodynamic damping and stiffness matrices, with 25
spanwise and 10 chordwise panels across the wing planform.
The aerodynamics are determined at a nominal flight condition of 160 m/s TAS at 20 km altitude, representing the cruise
condition of the popularised HALE vehicle Global Hawk [25].
Further, matched frequencies are used to calculate the flutter
velocity.

gi (x) 0,
with,

for i = 1, 2, ..., m
lxu

[3]

where l and u are the vectors of lower and upper limits of the
design variables respectively. The design variables specify the
panel thickness of each structural member, with a search space
bounded by li = 0.001 m and ui = 0.05 m for the ith panel
thickness.
Although the objective of the optimisation routine is to determine the minimum weight structure for each configuration,
the aeroelastic stability and the severity of the transient response impose significant constraints to the optimal solution.
These conditions are imposed as inequality constraints g1 (x)
and g2 (x), where g1 (x) represents the flutter mechanism and
g2 (x) the root bending moment of the respective structures.

Fourth Australasian Unmanned Systems Conference , 2014

Mathematically, the constraints are represented by:



F SV g1 (x) 0, for flutter speed
g(x) =
BMV g2 (x) 0, for bending moment

[4]

where the values of F SV and BMV represent the minimum


allowable flutter speed and maximum bending moment respectively. The optimal solution thus aims to satisfy simultaneously
a minimum bending stiffness to resist large root bending moments, as well as a torsional stiffness requirement to delay the
onset of orthogonal mode coupling, and hence, the flutter instability.
In this implementation of the optimisation, the flutter speed
constraint F SV is chosen to be 200 m/s, equivalent to the dive
speed of typical HALE aircraft. Solutions producing a flutter
speed above this value are deemed to exhibit exhibit the best
possible flutter performance as the instability occurs outside of
the flight envelope. The bending moment constraint, BMV ,
due to the gust load is taken to be -180000 Nm, the equivalent
bending moment of that generated by an elliptical lift distribution in steady level flight.
Fitness Function
The fitness function, f (x), for minimisation is the weight of
the respective wing structures. The constrained optimisation
is converted to an unconstrained problem by augmenting the
fitness function to include penalty terms for each of the aeroelastic constraints. The cost function for minimisation, , is
thus given by:
= (x, r) = f (x) +

m
X



rj Gj gj (x)

various permutations of the design variables examined. The


optimal gene from a particular set is carried forward to the next
iteration, where it may seed a new set of solution genes for the
next generation. The process proceeds until an optimal gene is
obtained.
BGAs represent the design variables as a discrete set through
chromosomes, essentially strings of binary digits. The creation
of new solutions from a particular gene set is governed by the
principals of:
Crossover. A section of a pair of genes is swapped
0110
0011

100001
0110
=
010000
0011

Mutation. The value of a cell within a gene is randomly


swapped
0 1 1 0 0 1 0 0 0 0 = 0 1 1 0 1 1 0 0 0 0
Translation. The order within the gene is randomly
swapped
0 1 1 0 0 1 0 0 | 0 0 = 0 0 | 0 1 1 0 0 1 0 0
New Blood. Introduction of new genetic characteristics,
independent of the current genes.
The length of each gene is determined by the range of values in the design variable bounds, with each chromosome a
concatenation of each of structural groups panel thicknesses,
represented in binary form. In this paper the optimisation algorithm follows the process given in Figure 5.

[5]

j=1

{z

penalty term

where the term rj represents an arbitrarily large positive constant penalty parameter for the j th constraint and Gj is the
penalty function for each of the constraints, as given by:

|F SV g1 (x)|, for j = 1
G(x) =
[6]
|BMV g2 (x)|, for j = 2
The penalty term only contributes to the cost function if the
constraints have not been satisfied. The penalty constants are
chosen to impart similar magnitudes for both the flutter mechanism and bending moment. The large magnitudes of these constants are intended to severely penalise solutions that do not
satisfy the aeroelastic criteria, ensuring the optimal structure
exhibits satisfactory aeroelastic performance. The modulus of
each constraint is included to ensure penalty functions increase
the cost of solutions not meeting minimum flutter speed or
maximum bending moment. The cost function of the optimal
solution should thus represent the weight of the structure.
Binary Genetic Algorithm
In order to determine the most efficient structure for each
configuration a Binary Genetic Optimisation (BGA) routine is
applied in the present work. As a whole, genetic algorithms are
founded upon Darwins principal of natural selection, survival
of the fittest. Design variables are expressed as genes, with

010000
100001

Fig. 5. BGA Solution Process

Results
Results are given with respect to the structural weight optimisation for each structure. A discussion of the optimal member sizing is provided with regards to the aeroelastic constraints.
The dynamic behaviour and aeroelastic performance of the optimal structures is then presented.
Structural Weight
The BGA optimisation routine progresses through 100 generations for each of the structural configurations. Figure 6
gives the convergence history of the three superior genes of the
conventional model. An optimal solution is achieved within
56 generations. The multi-lobe structure exhibits an analogous
history, with the optimal parameter set realised in 59 generations. For each structure the optimal gene results in a cost function equivalent to the structural weight, indicating that both
aeroelastic constraints are satisfied.
5

x 10

Cost

0
0

20

40
60
Generation

80

100

Fig. 6. Conventional Model Convergence Plot

Given the optimal panel thicknesses from the BGA routine,


the conventional structure sees an 83% weight reduction from
the original model, whereas the multi-lobe configuration experiences an 86% decrease. The rib and spar model sizes significantly lighter than the alternative structure, with final weights
of 495 kg and 1920 kg respectively. Although the multi-lobe
structure provides a minimum 65% improved volumetric efficiency, the weight penalty is substantial. The heavy sizing
of the multi-lobe model is attributed to the single grouping of
the lobe arcs along the span, where the greater thickness required at the root is carried throughout the entirety of the span.
With the alternative structure sizing to 288% that of the conventional model, the application of this stiffening structure to
the generic HALE wing is not warranted. Although [9] also
applied a multi-lobe topology that sized heavier than an equivalent conventional model, the 14% weight increase for a low
aspect ratio wing was a justifiable trade-off for the improved
volumetric efficiency. Further, the conventional configuration
results in a structural wing weight of 11.3 kg/m2 . This is 15%
above the maximum HALE structural wing weight proposed
in [27], and indicates the need for further design iteration with
a possible application of composite members.
Considering the sizing of individual stiffening members, it
can be deduced that the topology of each configuration is not

optimised for the aeroelastic constraints considered. In the


conventional model, the front two spars and third rib from the
root size to the minimum thickness bounds. This is also observed in the second and third webs (from the leading edge),
as well as the front upper and lower lobe arcs in the multilobe structure. As such, further progression of this project will
require a topology optimisation prior to sizing, such as that
implemented in [28] and [29].
Although no structural members size to the upper bound of
the search space, considering the heaviest sized components
gives insight into the critical constraint for this particular geometry. For the conventional model, the two outermost rib
sections yield a plate thickness of 200% that of the thickest
inboard section. Despite [30] finding a concentrated mass towards a wing tip is detrimental to flutter stability, the gust response is deemed the critical aeroelastic consideration in this
study. The increased mass at the wing tip is thus an expected
outcome, as it acts to provide the greatest restoring bending
moment to oppose the vertical gust. Under both configurations
the spanwise stiffening members (spars, webs and lobe arcs)
at the trailing edge size significantly thicker than those at the
middle and leading edge stations. This is a likely result of the
critical bending moment constraint, with the routine generating a mass distribution offset furthest from the flexural axis to
provide the largest restoring moment.
Natural Frequencies
Although the flutter analysis considers the first ten natural
modes, only the first five are presented here for ease of illustration. Table 2 gives the natural frequencies and mode shape
characteristic for the conventional configuration. Typical of
high aspect ratio wings [3133], first wing bending is excited at
low frequencies (< 5 Hz), with the presence of in-plane bending in the first five modes.
Table 2. Natural Frequencies of Optimal Conventional Structure
Mode Frequency (Hz)
Character
1
1.04
1st Out-of-Plane Bending
2
5.50
1st In-Plane Bending
nd
6.60
2 Out-of-Plane Bending
3
4
13.36
1st Torsion
rd
5
18.25
3 Out-of-Plane Bending

The natural frequencies of the multi-lobe structure are given


in Table 3. Again, the low frequency in-plane bending mode
is present. The dynamic behaviour of the two configurations is
qualitatively similar, however the first in-plane bending mode
of this structure occurs at higher frequency than the second
out-of-plane bending mode. In both models the observed low
frequency mode shapes exhibit either pure bending or pure torsion characteristics, with no cross-coupling throughout the first
ten natural frequencies.
Table 3. Natural Frequencies of Optimal Multi-Lobe Structure
Mode Frequency (Hz)
Character
1
0.95
1st Out-of-Plane Bending
2
5.87
2nd Out-of-Plane Bending
3
6.58
1st In-Plane Bending
4
10.45
1st Torsion
rd
5
15.96
3 Out-of-Plane Bending

Fourth Australasian Unmanned Systems Conference , 2014

Flutter
The linear flutter boundary of both configurations lies well
outside the flight envelope of a HALE mission profile. In Figure 7 the V -g and V - plots for the conventional model are
given. Across the velocity range considered, little change is
observed in the natural frequencies and the flutter instability
is not present. The first torsion frequency begins to approach
the third out-of-plane bending at higher velocities, with these
modes likely to coalesce at supersonic speeds to produce flutter.

Frequency (Hz)

Mode 1

Mode 2

Mode 3

Mode 4

Mode 5

15
10
5
0
0

50

100

150
Velocity (m/s)

200

250

Damping Ratio

0
0.05
0.1
0.15
0

50

100

150
Velocity (m/s)

200

250

Fig. 7. Conventional Model V -g and V - Plots

Similar dynamic behaviour is observed for multi-lobe structure in Figure 8. The natural frequencies of each mode remain
relatively constant throughout the velocity range, with a small
exception again in the first torsion mode, which begins to approach the third out-of-plane bending frequency. This results
in the damping of mode 4 reverting to approach the zero damping condition, however this exists well outside the flight envelope of subsonic HALE vehicles.

Frequency (Hz)

Mode 1

Mode 2

Mode 3

Mode 4

Mode 5

15
10
5
0
0

50

100

150
Velocity (m/s)

200

250

Damping Ratio

Transient Aeroelasticity
The transient root bending moment response for each configuration is given in Figure 9. In line with the results of the
member sizing, the bending moment constraint appears critical for both configurations. The maximum root bending moment is marginally achieved in each case with -179.6 kN.m
and -173.4 kN.m for the conventional and multi-lobe models
respectively. There is little difference between the two as once
the constraint is satisfied it no longer contributes to the fitness
function, at which point the structural weight drives the optimisation routine.
200
Conventional
MultiLobe

150
100
50
0
50
100
150
200
0

0.5

1.5
Time (s)

2.5

Fig. 9. Transient Root Bending Moment Response

0.05

0.1
0

A small wind-off frequency spacing is evident between the


first in-plane bending and second out-of-plane bending modes
in both models. Although a narrow frequency spacing between
orthogonal modes is indicative of a structure prone to the flutter
instability, [34] found that in-plane bending modes have a negligible effect on free flight flutter. This is evident in this study
for both configurations, with no apparent interactions with the
in-plane bending modes.
The high flutter stability of the wings can, in part, be attributed to the geometry. The rectangular planform produces
pure low frequency mode shapes. In the absence of orthogonal
characteristics within each mode, cross-coupling is delayed to
a higher energy state. The addition of sweep or taper to the
generic planform is likely to increase the significance of the
flutter speed to the structural optimisation. Moreover, linearity within the aeroelastic system may be a source of overestimation of the stability boundary. If, in an extreme case, the
50% reduction in flutter velocity observed through the inclusion of structural nonlinearities in [22] pertains to the models considered in this paper, instability at operational velocities
may emerge. A nonlinear aeroelastic analysis of these configurations is proposed for future work.

Root Bending Moment (kN.m)

50

100

150
Velocity (m/s)

200

Fig. 8. Multi-Lobe Model V -g and V - Plots

250

The maximum tip deflection experienced during the gust


loading is on the order of 20 cm for each structure. This is
well within the 30% span tip deflection proposed as a maximum bounds by [5]. This is an indication that the bending
moment constraint applied is too conservative and the resultant wing is overdesigned. As an area for further work, a tip
displacement constraint of 30% span is to be implemented and

the impact on root bending moment and structural weight will


be analysed.

Conclusions
In the present paper, an alternative internal wing stiffening
structure for HALE applications has been investigated. This
structure was sized using a Binary Genetic Algorithm under
transient and dynamic aeroelastic constraints, along with an
equivalent conventional wing model. The numerical results indicate the multi-lobe structure is inefficient in a high aspect
ratio wing. Although the intrinsic dynamic behaviour, flutter
stability and maximum root bending moment of the optimal
structures are qualitatively similar, the 288% increase in structural weight of the alternative configuration does not justify
the 65% improvement in volumetric capacity. The results have
however shown the topology of the two configurations is not
optimal. Further investigation will be necessary to develop an
optimal topology governed by nonlinear aeroelastic analysis.

References
1. Yinan Wang, Andrew Wynn, and Rafael Palacios. Robust aeroelastic control of very flexible wings using intrinsic models. 54th AIAA/ASME/ASCE/AHS/ASC
Structures, Structural Dynamics and Materials Conference, Boston, Massachusetts, USA, 2013.
2. Peter Van Blyenburgh. UAVs: An Overview. Air &
Space Europe, 1(5):4347, 1999.
3. Kitt C Reinhardt, Thomas R Lamp, Jack W Geis, and
Anthony J Colozza. Solar-powered unmanned aerial vehicles. In Energy Conversion Engineering Conference,
1996. IECEC 96., Proceedings of the 31st Intersociety,
volume 1, pages 4146. IEEE, 1996.
4. Adam C Watts, Vincent G Ambrosia, and Everett A
Hinkley. Unmanned aircraft systems in remote sensing
and scientific research: Classification and considerations
of use. Remote Sensing, 4(6):16711692, 2012.
5. Carlos ES Cesnik and Weihua Su. Nonlinear aeroelastic
simulation of X-HALE: a very flexible UAV. In 49th
AIAA Aerospace Sciences Meeting including the New
Horizons Forum and Aerospace Exposition, Orlando,
Florida, 2011.
6. Zdobyslaw Goraj. Ultra light wing structure for high
altitude long endurance UAV. In ICAS Congress, pages
1019, 2000.
7. Matthew J Scott, Jamey D Jacob, Suzanne W Smith,
Laila T Asheghian, and Jayanth N Kudva. Development
of a novel low stored volume high-altitude wing design.
In Proceedings of 50th AIAA/ASME/ASCE/AHS/ASC
Structures, Structural Dynamics, and Materials Conference, volume 4, pages 20092146, 2009.
8. Johan Steelant and Marco van Dujin. Structural Analysis of the LAPCAT-MR2 Waverider Based Vehicle. In
17th AIAA International Space Planes and Hypersonics
Systems and Technologies Conference, 2011.

9. Nicholas F Giannelis, Gareth A Vio, Dries Verstraete,


and Johan Steelant. Temperature effect on the structural design of a Mach 8 vehicle. In Applied Mechanics
and Materials, volume 553, pages 249254. Trans Tech
Publ, 2014.
10. Somanath Nagendra, D Jestin, Zafer Gurdal, Raphael T
Haftka, and Layne T Watson. Improved genetic algorithm for the design of stiffened composite panels. Computers & Structures, 58(3):543555, 1996.
11. Sameer B Mulani, David Havens, Ashley Norris,
R Keith Bird, Rakesh K Kapania, and Olliffe Robert.
Design, optimization and evaluation of Al2139 compression panel with integral T-stiffeners. Journal of Aircraft, 50(4):12751286, 2013.
12. Andrea Arena, Walter Lacarbonara, and Pier Marzocca. Post-flutter analysis of flexible high-aspectratio wings. In Collection of Technical Papers53rd
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference. AIAA, 2012.
13. Joseph A Garcia. Numerical investigation of nonlinear
aeroelastic effects on flexible high-aspect-ratio wings.
Journal of Aircraft, 42(4):10251036, 2005.
14. Christopher M Shearer and Carlos ES Cesnik. Nonlinear flight dynamics of very flexible aircraft. Journal of
Aircraft, 44(5):15281545, 2007.
15. Deman Tang and Earl H Dowell. Experimental and theoretical study on aeroelastic response of high-aspect-ratio
wings. AIAA journal, 39(8):14301441, 2001.
16. Bernard Etkin. Turbulent wind and its effect on flight.
Journal of Aircraft, 18(5):327345, 1981.
17. JG Jones. A Relationship between the Power-SpectralDensity and Statistical-Discrete-Gust methods of Aircraft Response Analysis. Royal Aircraft Establishment,
Farnborough, England, UK, RAE TM Space, 347, 1984.
18. P Chud`y. Response of a light aircraft under gust loads.
Acta Polytechnica, 44(2), 2004.
19. Weihua Su and Carlos ES Cesnik. Dynamic response
of highly flexible flying wings. In Proceedings of
the 47th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, pages 412
435, 2006.
20. Guowei Yang and Shigeru Obayashi. Numerical analyses of discrete gust response for an aircraft. Journal of
Aircraft, 41(6):13531359, 2004.
21. Daniella Raveh. CFD-based gust response analysis of
free elastic aircraft. ASD Journal, 2(1):2334, 2010.
22. Mayuresh J Patil and Dewey H Hodges. On the importance of aerodynamic and structural geometrical nonlinearities in aeroelastic behavior of high-aspect-ratio
wings. Journal of Fluids and Structures, 19(7):905915,
2004.

Fourth Australasian Unmanned Systems Conference , 2014

23. Anthony P Ricciardi, Mayuresh J Patil, Robert A Canfield, and Ned Lindsley. Evaluation of quasi-static
gust loads certification methods for high-altitude longendurance aircraft. Journal of Aircraft, 50(2):457468,
2013.
24. Guy Norris. Global leader: The success of Northrop
Grummans Global Hawk is helping to transform opinion on the military capability of unmanned air vehicles.
In Flight International, pages 2631. Flight Global, 30
January - 5 February 2001.
25. Anthony Colozza and James L Dolce. High-altitude,
long-endurance airships for coastal surveillance. NASA
Technical Report, NASA/TM-2005-213427, 2005.
26. Frederic M Hoblit. Gust loads on aircraft: concepts and
applications. AIAA, 1988.
27. Glenn C Grimes. Composite materials: testing and design (Tenth volume), volume 1120. ASTM International,
1992.
28. Mark D Sensmeier and Jamshid A Samareh. Automatic aircraft structural topology generation for multidisciplinary optimization and weight estimation. In 46th
AIAA/ASME/ASC Structures and Materials Conference.
Austin, Texas: American Institute of Aeronautics and Astronautics, volume 1893, 2005.
29. Vivek Mukhopadhyay, Su-Yuen Hsu, Brian H Mason,
Mike D Hicks, William T Jones, David W Sleight,
Julio Chu, Jan L Spangler, Hilmi Kamhawi, and Jorgen L Dahl. Adaptive modeling, engineering analysis
and design of advanced aerospace vehicles. In 47th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference; Newport, RI, pages
21822194, 2006.
30. SA Fazelzadeh, A Mazidi, and H Kalantari. Bendingtorsional flutter of wings with an attached mass subjected to a follower force. Journal of Sound and Vibration, 323(1):148162, 2009.
31. Vasily V Chedrik, Fanil Z Ishmuratov, Svetlana I Kuzmina, and Mikhail Ch Zichenkov.
Strength/aeroelasticity research at multidisciplinary
structural design of high aspect ratio wing. In 27th
Congress of the International Council of the Aeronautical Sciences, Nice, France, 2010.
32. Jiri Cecrdle and Ondrej Vich. Eigenvalue and flutter sensitivity analysis of airliner wing. In 27th Congress of
the International Council of the Aeronautical Sciences,
Nice, France, 2010.
33. Anas El Arras, Chan Hoon Chung, Young-Ho Na,
SangJoon Shin, SeYong Jang, SangYong Kim, and
Changmin Cho. Various structural approaches to analyze an aircraft with high aspect ratio wings. International Journal of Aeronautical and Space Sciences,
13:446457, 2012.

34. Weihua Su. Coupled nonlinear aeroelasticity and flight


dynamics of fully flexible aircraft. PhD thesis, University
of Michigan, 2008.

You might also like