Sunteți pe pagina 1din 14

52nd AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference<BR> 19th

4 - 7 April 2011, Denver, Colorado

AIAA 2011-1916

Nonlinear Aeroelastic Formulation for Flexible


High-Aspect Ratio Wings via Geometrically
Exact Approach
Andrea Arenaand Walter Lacarbonara
Sapienza University, Rome, Italy, 00184
Pier Marzocca
Clarkson University, Potsdam, New York, 13676
The nonlinear aeroelastic modeling and behavior of HALE wings, undergoing large deformations and exhibiting dynamic stall, are presented. A fully nonlinear three-dimensional
structural model, based on an exact kinematic approach, is coupled with the incompressible
unsteady aerodynamic model obtained via a reduced-order indicial formulation accounting
for viscous eects, in term of dynamic stall and ow separation. To this end, a modied
Beddoes-Leishman model is employed. Aeroelastic simulations are performed by reducing
the governing equations to a form amenable to numerical integration. Space and time integrations are conducted using a numerical scheme that includes PDE, associated with the
equation of motion of the exible wing, and ODEs, associated with the lag-state formulation
pertinent to the unsteady aerodynamic loads, in a hybrid solution form. The numerical
investigations show that the proposed approach is suitable for studying the aeroelastic
behavior of highly nonlinear wings, for an improved understanding of the nonlinear phenomena occurring particularly in the neighborhood of the utter boundary and in the
post-critical regime.

I.

Introduction

In recent years, the use of High-Altitude Long-Endurance (HALE) aircrafts has been explored for ight
missions including environmental monitoring, military reconnaissance, and telecommunication relay, to name
only a few. The aspect ratio of HALE aircraft wings is usually very high, in order to increase their specic
mission performance with a high aerodynamic eciency. The inherent exibility is such that large deections
are possible during normal ight operation which includes slow maneuvers at high altitudes and low speeds.
In these conditions, operation at angles of attack close to stall may lead to aeroelastic instabilities that are
associated with dynamic stall. Clearly these aircrafts are susceptible to dynamic instabilities, such as utter,
which can excite large-amplitude Limit Cycle Oscillations (LCO).
In general, structural nonlinearities are governed by elastodynamic deformations that aect the whole
structure. Concentrated nonlinearities, on the other hand, act locally and are commonly found in control
mechanisms or in the connecting parts between wing, pylon, engine, or external stores. While the most
prevalent nonlinearities in aircraft structures are concentrated structural nonlinearities, usually represented
by cubic stiness and the freeplay,1, 2 for HALE wings moderate- to large-amplitude deformations and the
associated nonlinear behaviors are certainly more common.35
Structural nonlinearities, coupled with aerodynamics eects, can induce LCOs in the ight envelope.
An extensive review of the state-of-the-art of nonlinear aeroelasticity is provided in Dowells monograph6
and the references therein. The importance of accounting for large-amplitude deformations from the stable
equilibrium state is documented in.4, 5
PhD

Candidate, Structural and Geotechnical Engineering Department, via Eudossiana 18 00184 Rome, Italy
Professor, Structural and Geotechnical Engineering Department, via Eudossiana 18, 00184 Rome, Italy
Associate Professor, Mechanical and Aeronautical Engineering Department, 8 Clarkson Ave, CAMP 234, Potsdam, New
York 13699-5725, AIAA Senior Member
Associate

1 of 14
American
Institute
Aeronautics
Copyright 2011 by the American Institute of Aeronautics and
Astronautics,
Inc. All of
rights
reserved. and Astronautics

The unsteady aerodynamic nonlinearities may arise from ow separation and stall at high angles of
attack, shock waves in the transonic, supersonic and hypersonic speeds where phenomena like ionization and
plasma may arise.712 In the present work, the considered aerodynamic nonlinearities are associated with
the presence of aerodynamic stall. There exists an abundance of dynamic stall models, some simple, some
more sophisticated, however, to date no model can be considered superior to the others. In the literature one
can nd models from B-L,13 ONERA,14 DLR,15 and their derivatives, Goman and Khrabrov,16 Truong,17
ye,18 Ris,19 and Snel.20 In HALE wing aerodynamics, the dynamic stall model has been considered in
very few studies, primarily using the helicopter rotor blade dynamic stall ONERA model based on the initial
works by Tran and Petot21 and Dat and Tran22 who considered large unsteady motions. The model was
subsequently improved by Dunn23 and applied to study the nonlinear aeroelastic behavior of exible rotor
blades as well as long-span wings by Tang and Dowell.3

II.

A Fully Nonlinear Aeroelastic Model for High-Aspect-Ratio Wings

One of the most critical aspects in nonlinear aeroelasticity is the understanding of how nonlinearities
aect the system dynamics. Due to their exibility, HALE wings exhibit structural and aerodynamic nonlinearities that can potentially initiate aeroelastic instabilities both above and below the utter speed predicted
by the linear theory.24 In this section a structural nonlinear beam model will be presented along with the
pertinent nonlinear aerodynamic description. It is worth mentioning that the literature is paved with several descriptions in a geometrically exact form, such as the mixed variational formulation based on exact
intrinsic equations for the dynamics of the beams by25, 26 and the nonlinear dynamic theory of heterogeneous
anisotropic elastic rods by.27 In this work we are presenting an alternative approach based on the Antman
original work.28
Kinematic Formulation. The three-dimensional parametric model is based on a geometrically exact
semi-intrinsic theory which yields the equations of motion, here obtained within the context of an Updated
Lagrangian Formulation (ULF).28, 29 In the orthonormal basis of a xed inertial reference frame (e1 , e2 , e3 ),
whose origin is in the elastic center of the root wing section, the reference (stress-free) conguration of the
wing is described by the position of the elastic line, r = xe3 , where x is the reference coordinate while
the orientation of the reference principal inertial axes of the wing cross sections is given by the unit vectors
(b1 , b2 , b3 ), where (b1 , b2 ) are lying in the cross-section plane and are rotated with respect to the axes (e1 , e2 )
by angle 3 and b3 is collinear with e3 (see Fig.2). The angle-of-attack W at the wing root section is also
considered in the model. Within the context of ULF, two dierent wing congurations are taken into account

b1

b2
e2

W + 3 + 3(x,t)
b02CE

e1

b3

b1

rS
CE

rS,0

W + 3 + 03(x)

e3
b03

rS

Figure 1. 3D wing congurations: stress-free B, static B0 and dynamic B.

to fully describe its kinematics and dynamics, those induced by the static aeroelastic loads and those due to
dynamic aeroelastic loads. By considering the conguration under the action of static forces, denoted by B0 ,
2 of 14
American Institute of Aeronautics and Astronautics

let us dene the position vector of the wing elastic line as r 0 (x) = r(x) + u0 (x) while we let the orientation of
the cross sections be described by unit vectors (b01 (x), b02 (x), b03 (x)). On the other hand, the current dynamic
0
conguration
B is described
(
) by the position vector r(x, t) = r (x) + u(x, t) and the local frame is dened by
1 (x, t), b
2 (x, t), b
3 (x, t) . To express the unit vectors of the local frame, b0 and b
i , i = 1, 2, 3, in terms
b
i

are introduced. The total rotation of the local


of the inertial frame, nite rotations matrices R0 and R

R0 (x) R(x, t)
basis with respect to the xed reference frame (e1 , e2 , e3 ) can be dened as R(x,
t) := R
represents the rotation of the section principal axes of inertia with respect to the inertial frame.
where R
The wing generalized total strain parameters are dened for both the static and dynamic conguration, thus
the stretch, the shear strains and the bending and twist curvatures, can be obtained through the following
vectorial relations:
0 := r 0x b03 , 10 := r 0x b01 , 20 := r 0x b02 , x b0k = 0 b0k
3 , 1 := r x b
1 , 2 := rx b
2 , x b
k =
k
b
:= r x b

in B 0
in B

(1)

where subscript x, here and henceforth, denotes partial dierentiation with respect x. In Eq.(1), 0 and
1 , b
2 directions,
are the wing stretches, (10 , 20 ) and (
1 , 2 ) are the shear strains along the b01 , b02 and b
0
0 0
0 0
respectively. The components of the curvature vectors in the local bases ( = 1 b1 + 2 b2 + 03 b03 and
1 +
2 +
3 ) denote the twist curvatures (0 and
=

1 b
2 b
3 b
3 ) and the bending curvatures (01 , 02 and
3

1 ,
2 ), respectively.
b2

e2
f

cI

fA
b1

C
d

3
C e1

dE

cA
W

CA
b/2

2b

Figure 2. 2D lifting surface, reference frames.

Dynamic Formulation. By enforcing the balance of linear and angular momentum, the static (equilibrium) aeroelastic governing equations and the dynamic aeroelastic equations for the wing are obtained in
the form of (2) and (3), respectively. The stress and moment resultants are dened on the wing cross section

at the reference position x and time t. The vectors n0 (x) and n(x,
t) denote the total stress resultants

(referred to as contact forces) whereas m0 (x) and m(x,


t) denote the moments resultants (referred to as

contact couples), respectively. The balance equations are obtained for the current congurations (B0 , B)
and written in terms of total forces and moments and incremental kinematic parameters, u0 (x), 0i (x) and
u(x, t), i (x, t), respectively.
Equilibrium Equations. The static equilibrium equations are written as
x n0 (x) + f 0 (x) = o

(2)

x m0 (x) + x r 0 (x) n0 (x) + c0 (x) = o


where the external forces f 0 (x) and couples c0 (x) include the wing and stores weights, and the static
aeroelastic loads.
In this study we are primarily concerned with the behavior of the wing, therefore any additional static
contribution to weight and aerodynamic loads will be discarded. The trim ight condition has also been
discarded, although it has been shown in recent studies5, 30, 31 that will aect the utter characteristics of
exible wings when aeroelasticity is coupled with ight dynamics. On the other hand, by considering their
importance on the aeroelastic response and instabilities,32, 33 wing stores have been included in the model.
Stores, such as fuel tanks, pod with logistic payload, missiles, etc., can be arbitrarily located in various

3 of 14
American Institute of Aeronautics and Astronautics

span-wise and chord-wise wing positions. The position vector of the stores (whose mass is mS ) in the inertial
reference frame is r S = r S + xS e3 , where rS = xS1 e1 + xS2 e2 .
Equations of Motion. The equations of motion are written here referring to the elastic center C E . Since
it is not coincident with the wing cross-section center of mass, the inertial forces and angular momentum
are modied so that the equations of motion can be written as
(
)

x n(x,
t) + f (x, t) = Autt + t i + i
(3)

(x, t) = J E t + (J E ) + i utt
x m(x,
t) + x r(x, t) n(x,
t) + c
where subscript t denotes dierentiation with
respect to)time t; i is the vector listing the area static moments
(
E
2 , b
3 and is the mass density; J E is the inertia tensor of
of inertia with respect to the local frame C , b1 , b
the wing cross sections per unit length referred to the elastic center C E ; is the incremental angular velocity
k = b
k . In addition,
vector of the wing cross sections dened in the local reference frame such that t b

(x, t) denote the total external forces and couples per unit length, respectively, including the
f (x, t) and c
aerodynamic loads, acting on the wing in the current dynamic conguration and the damping forces assumed
proportional to the section velocity ut and angular velocity .
Finally, linearly elastic constitutive equations are assumed in order to describe the relations between
internal forces and couples and the strain parameters; moreover, a cantilevered beam scheme is considered
so as to model the clamping of the wing to the aircraft body and, for the dynamic incremental problem,
natural initial conditions are taken into account.
Nondimensional Form. The governing static and dynamic equations (2) and (3), and the associated
boundary
conditions, are cast in nondimensional form by using the wing span l as characteristic length and

Al4 /EJ1 as characteristic time, where A is the area of the wing cross section, E the Young modulus and
J1 the moment of inertia about the local axis b01 . It is worth mentioning that the loads due to the presence
of the stores on the wing are here modeled as Dirac functions along the span, thus the nondimensional
(
= MS /(Al) and MS is the
equivalent distributed mass can be written as m(
e x) = 1 + M
xx
S ) where M
S
store mass at nondimensional position x
. Furthermore, also the equivalent structural damping coecients
turn out to be varying along the span as they can be expressed as follows:
[
]
di (
x) = 2i i Jm + JMS (
xx
S )
where i is the critical damping ratio of the i th mode, i is the nondimensional natural frequency of
the i th mode of the wing, with the presence of the store, Jm is the nondimensional mass moment of the
wing and JMS is the nondimensional equivalent mass moment of the store. The vector-valued equations
(2) and (3), projected into the global basis (e1 , e2 , e3 ), yield six nonlinear partial-dierential equations in
six independent kinematic unknowns, respectively. The obtained equations govern the elasto-static and
elasto-dynamic problem of the wing.

III.

Unsteady Aerodynamics for the Wing Section Undergoing Dynamic Stall

An unsteady aerodynamic model based on a modied B-L formulation,13 which takes into account considerations and recommendations from RIS,19 has been adopted. This model is capable of capturing the
nonlinear dynamic stall eect that may be typically associated with HALE wings. In the linear fully-attached
ow regime, the model is capable of reproducing the results obtained by solving the unsteady aerodynamic
problem represented through the indicial formulation. On the other hand, the selected model is capable of
predicting the typical hysteresis behavior in the critical stall regime.
It is assumed that the lift, drag, and aerodynamic moment are located at the aerodynamic center C A
lying at one-fourth of the chord from the leading edge (see Fig.2). Those resultants have then to be reduced
to the elastic center C E .
Lift force assuming fully attached flow. The unsteady lift expressed in terms of dynamic variables
can be written in terms of its noncirculatory and its circulatory components provided by the indicial formulation.24 In the unsteady lift expression the added mass contribution for moderate reduced frequencies is
negligible.19 The indicial function used to compute the circulatory part of the unsteady lift, for thin airfoils,
is given by the well-known Wagner function.
4 of 14
American Institute of Aeronautics and Astronautics

The unsteady aerodynamic lift coecient24 accounting for the pitch rate contribution, for the fully
attached ow, can be written as CLFA (x, t) = CL ,3 (E3 (x, t) 0 ) + t 3 (x, t)b/U , where CL ,3 is the
slope of the lift curve in the linear region of attached ow, b is the wing half-chord dimension, U the freestream ight speed and 0 is the zero lift angle-of-attack for cambered airfoils while E3 (x, t) is the eective
angle-of-attack which, consistent with the indicial formulation, can be dened by two further additional
lag-state variables Wi (x, t):
(
)
2
2

1
E
W
Bi Ai Wi (x, t)
(4)
Ai +
3 (x, t) = 3 (x, t) 1
TU i=1
i=1
t Wi (x, t) +

Bi
Wi (x, t) = W
3 (x, t) i = 1, 2
TU

(5)

0
where W
3 (x, t) = 3 (x) + w(x, t)/U and w(x, t) is the downwash at the three-quarter chord assuming
positive counterclockwise angles of attack (see Fig.2), represented as

w(x, t) = t u2 (x, t) + dW t 3 (x, t) + U 3 (x, t)

(6)

Herein 3 (x, t) is the incremental torsional rotation of the wing cross-section, t u2 (x, t) is the incremental
velocity of the elastic center C E , dW is the positive distance from the elastic center C E and the three-quarter
chord, while t () indicates dierentiation with respect to time t and the time constant TU is dened as
TU := b/U .
Aerodynamic forces with trailing edge flow separation. The eect of the leading edge separation
will be neglected in this study, a reasonable assumption for the particular class of ying vehicles considered
and their corresponding ight speed regimes and associated Reynolds numbers. As postulated in the B-L
model19 of the trailing edge separation, the distance of the separation point from the trailing edge f ST , and
the lift coecient for fully separated ow, CLFS , will be dened as a function of the static lift curve.
(
)
v
2
ST
3 0 f ST (3 )
u
(

C
ST

3
L
L
,3
CL (3 )
u
(
) 1 , CLFS (3 ) =
(7)
f ST (3 ) = 2t

1
f ST (3 )
C

3
L ,

where CLST (3 ) is the static lift curve for the considered wing section, see Fig. 3.
It is worth noticing that the lift coecient expression in (7), at low angles of attack (f ST (3 ) = 1), imply
that the lift coecient approach to the value CLST (3 )/2; moreover, for fully separated ows (f ST (3 ) = 0),
CLFS (3 ) corresponds to the static lift curve.
In order to dene the dynamics of trailing edge separation, two additional auxiliary state variables are
introduced, as also suggested in the B-L model. The rst state variable W3 (x, t) describes the separation as
a function of the pressure distribution over the wing section in terms of the lift on the airfoil; on the other
hand, the state variable W4 (x, t) is introduced to describe the dynamics of the boundary layer determining
the separation to lag behind the quasi-steady value f FS (FS
3 ). The state equations governing the phenomena
described above can be written as:
t W3 (x, t) +

1
1 FA
W3 (x, t) =
C (x, t) ,
TP
TP L

t W4 (x, t) +

1
1 FS FS
W4 (x, t) =
f (3 )
TF
TF

(8)

where TP is the time constant for the pressure lag and TF is a time constant for the lag in the boundary
layer; such constants are experimentally evaluated and depend on the considered wing section prole. In
this work, a NACA 6315 is assumed as aerodynamic surface and the values suggested in13 are taken into
account, namely, TP = 1.7 s and TF = 3 s. The equivalent quasi-steady separation point distance f FS (FS
3 ) is
computed as in19 where FS
is
an
equivalent
angle
of
attack
that
gives
the
same
quasi-steady
lift
evaluated
3
as FS
3 + 0 .
3 (x, t) = W3 (x, t)/CL ,
The drag force and aerodynamic moment are modeled in accordance with what is suggested in.19 The
drag accounts for the shift in the angle of the lift force and the one caused by the viscous boundary layer
along with the unsteady lift force components in the drag direction, which is perpendicular to the direction
given by E3 (x, t), while the aerodynamic moment accounts for the eect of the unsteady TE separation on
the travel of the center of pressure.
5 of 14
American Institute of Aeronautics and Astronautics

1.5

f ST( )
0.5

FA

CL ( ), CL ( ), CL ( ), f ST ( )

ST

CL ( )
1

ST

FS

-0.5
FA

CL ( )

-1

FS

CL ( )
-1.5
-30

-20

-10

10

20

30

[deg]

Figure 3. Static lift coecients and separation point distance.

Finally, the unsteady aerodynamic loads due to separation at trailing edge can be written in terms of the
four state variables Wi (x, t) as:
[
]
1
2
(2b)U
CL ,3 (E3 0 )W4 (x, t) + CLFS (E3 ) (1 W4 (x, t)) + t 3 TU ,
2
)
(
{
1
2

D(x,
t) = (2b)U
CDST (E3 ) + 3 E3 CLUS
2
(
)2 (
)2

1 f ST (E3 ) }
1 W4 (x, t)
0
E
ST
+ (CD (3 ) CD )
,

2
2
{
}

2
ST
A (x, t) = 1 (2b)2 U
M
CM
(E3 ) + CLUS [ ST (W4 ) ST (f ST (E3 ))] TU t 3
2
2
t)
L(x,

(9)

where ST represents the arm of the lift force and it is a function of the separation point distance f ST on the
upper wing surface.
When the governing aeroelastic equations are solved simultaneously, the pitching motion is itself one
of the states variables and its dynamics result coupled with the other aeroelastic variables. The coupled
problem will be solved by time-marching simulations, where space-time mixed nite elements are used to
obtain a set of nonlinear algebraic equations which can be solved with the given initial conditions to get
the solution at each time step. It is worth mentioning that the steady-state equilibrium for a given ight
condition will be evaluated rst. The equations of motion can then be perturbed about this steady-state
condition to calculate the eigenvalues of the system, i.e., frequency and damping, that will be used for the
utter analysis. Furthermore, the pre-stress eect induced by the aeroelastic loads can be taken into account
in order to study the response of the wing to the aerodynamic forces.

6 of 14
American Institute of Aeronautics and Astronautics

IV.

Model Validation

The complete aeroelastic model is obtained by


coupling the structural and aerodynamic models;
the nonlinear aeroelastic equations are then implemented and solved by a Finite Element code. This
section presents initial validation results associated
with the dynamic stall model and the structural
nonlinear response to the static aeroelastic loads.
The NACA 6315 aerodynamic surface is assumed
in the simulations since for this particular airfoil all
modeling data required are readily available. The
drag, lift and moment coecients of such airfoil to
an oscillatory pitch (3 (t) = 03 3 (t)) are presented in Figure 4.

1.6
1.4
1.2

US

CL ( )

1
0.8
0.6
0.4

(t) = 3
(t) = 12

0.2

(t) = 21
0
-0.02

US
CD ( )

US

CM ( )

Unsteady Aerodynamics. The reduced fre-0.04


quency is k = b/U = 0.1, the mean value of
-0.06
the angle of attack is assumed 03 = 3 , 12 and

21 , respectively, and the amplitude of the pitch-0.08


ing oscillation is 3 = 4 . The behavior of the
-0.1
airfoil sectional model undergoing pitching motion
-0.12
and exhibiting full dynamic stall is represented by
the well-known dynamic loops centered at selected
-0.14
values of the mean angles. The comparison of the
0
5
10
15
20
25
30
aerodynamic characteristic responses from oscilla[deg]
tory pitching gives identical results to the one of
0.35
the RIS report,19 thus ensuring that the aerody(t) = 3
0.3
namic results are consistent and accurate. It is also
(t) = 12
0.25
important to note that the dynamic model gives re(t) = 21
sults identical to the unsteady inviscid ow solution
0.2
within the accuracy of the approximation of the in0.15
viscid response function, while, in the stall region,
0.1
the model exhibits the expected large dynamic loop
response. Clearly the eective slope of the coun0.05
terclockwise lift loop in the linear region is reduced
0
as compared to the dynamic loops in the stall and
-0.05
post-stall regions, which are clockwise, while the
0
5
10
15
20
25
30
drag loop is mainly due to inviscid eects the the
[deg]
loop on the aerodynamic moment is due to the upwash velocity, clearly aecting the eective angle- Figure 4. Dynamic lift, drag and moment coecients
of-attack. it is also evident that the dynamic loop for an harmonic input function, natural frequency =
in the lift associated with the stall is more promi- 0.1 rad/s.
nent that the one for the drag and moments, as they
revolve around the static curves. To evaluate the
aerodynamic characteristics undergoing an arbitrary pitching motion, the pitch angle has been driven by a
function similar in shape to a Mexican hat wavelet34 In this case the dynamic oscillations are provided from
an assumed angle-of-attack corresponding to the steady condition, 3 = 12 . 3 , providing the maximum
variation in the angle-of-attack, is given as 4 deg. The time parameters tw and t0 are assumed to be 15 sec
and 100 sec, respectively. The initial conditions in this case are set up as: W1 (0) = TBU1 3 , W2 (0) = TBU2 3 ,
(
)
W3 (0) = CL ,3 3 0 , W4 (0) = f ST (3 ). Fig.5(a) shows the Mexican hat input for dierent values of
mean angle 3 and the corresponding normalized responses in terms of unsteady coecients are presented in
Fig. 5b. The dynamic coecients appears to be a distorted version of the input 3 (t). This is due to both
a memory eect, the phase lag, and the presence of dynamic stall. In fact, if the variations of 3 (t) were
slow enough that the lift could be computed statically, and if the ow remained in the attached regime, then

7 of 14
American Institute of Aeronautics and Astronautics

2.25

3.5

1.75

2.5

(t) = 3 4
(t) = 12 4
(t) = 21 4

0
3

C D (t)/C D

2.5

US

(t) /

1.5
1.25

1.5

1
1
0.75
0.5
0.5
0
0.25

1.2

1.6

(t) = 3 4

1.5

(t) = 12 4

(t) = 12 4

1.4

1.1

(t) = 21 4
0

1.2

US

US

C M (t)/C M

1.3

C L (t)/C L

(t) = 3 4

1.15

1.1

(t) = 21 4

1.05
1

1
0.95

0.9
0.8

0.9

0.7
0.85
0

20

40

60

80

100

120

140

160

180

200

20

40

60

80

100

120

140

160

180

200

t [s]

t [s]

(a)

(b)

Figure 5. Mexican hat input functions and corresponding dynamic lift coecients by the nonlinear B-L model.
Angles of attack 3 : lift (a), drag and moment (b) coecients.

(
)
the coecients would linearly follow 3 (t) 0 . For the parameters used in this analysis, the maximum
unsteady CLUS is attained before the maximum angle-of-attack is reached. Although the input is symmetric,
the output is not. Furthermore, a third peak, of limited amplitude, occurs, before the system nally settles
to the steady state CLST . Similarly, drag and moments coecient exhibit similar patterns, as depicted in
Fig.5(b).
Modal Identification and Static Aeroelastic Behavior. The structural wing parameters are taken
from the experimental wing model proposed by Tang and Dowell3537 and reported in.38 The experimental
modal characteristics are compared with the results obtained via the proposed geometrically exact model.
As shown in Tab. 1, the natural frequencies, with and without the presence of a store mass at the tip, are in
very good agreement; the percent dierence with respect to the experimental model is also given. In Figs.
Table 1. Wing lowest ve natural frequencies with and without the store at the tip; comparison between the proposed nonlinear model (NL model) and the experimental results
(Exp).

Selected natural mode


First apping mode
Second apping mode
Third apping mode
First lagging mode
First twisting mode

Without store
NL model Exp35, 38
f [Hz]
f [Hz]
3.682
3.675
23.019
23.03
64.411
64.50
24.110
24.39
120.109
119.5

[%]
0.19
-0.05
-0.14
-1.14
0.51

With store at the tip


NL model Exp35, 38

f [Hz]
f [Hz]
[%]
2.316
2.625
-11.8
18.096
17.88
1.21
54.769

14.362
14.13
1.64
23.833
22.88
4.16

8 of 14
American Institute of Aeronautics and Astronautics

6 (a) and 6 (b) the equilibrium paths for the apping displacement u02 and the twisting rotation 03 are shown
assuming the lifting surface NACA0012 which was employed in the experiments carried out by Tang and
Dowell.35 The experimental and numerical response curves are in excellent agreement for the assumed steady
angle of attack W at the wing root. Pre-critical load paths were obtained for the lifting surface NACA6315
0.01

0.7

0.6
0.5

[deg]

-0.02

0.4
0.3

0
3

u 02 [m]

-0.01

-0.03

0.2

-0.04

0.1

-0.05
W

= 1

-0.06

10

15

20

25

30

35

10

15

U [m/s]

20

25

30

35

U [m/s]

(a)

(b)

0.12

1.8

0.1

1.6
1.4

0.08

1.2

[deg]

0.06
0.04
0.02

1
0.8

0
3

u 02 [m]

= 1

-0.1

0.6
0

0.4

-0.02

0.2

-0.04

= 2.2

-0.06

= 2.2

-0.2
0

10

15

20

25

30

35

40

10

U [m/s]

15

20

25

30

35

40

U [m/s]

(c)

(d)

Figure 6. Equilibrium paths under static aeroelastic loads for a NACA0012 lifting surface with store mass at
the wing tip: wing tip apping displacement u02 and twisting angle 03 for a root angle of attack W = 1 (a)
and (b), for a root angle of attack W = 2.2 (c) and (d).

considered in the present work, and assuming the same structural characteristics of the experimental wing.35
Figures 7 (a) and 7 (b) show the dierence between a linear and a nonlinear structural model in the static
aeroelastic response of the wing. The nonlinear model, that takes into account the variation of the wing
stiness arising from the considered geometric nonlinearities, shows a hardening eect both in the apping
displacement u02 and in the twisting angle 03 . These dierences with respect to the linear model appear to
be of about 17% already at a free-stream velocity of 35 m/s which is close to the value of the wind speed
at the onset of utter measured for the experimental wing. It is thus evident the importance of using a
nonlinear framework also for a better understanding of the dynamical aeroelastic response of the wing.3, 35, 36
Figures 7 (c) and 7 (d) propose the variation of the lowest lagging, apping, and twisting frequency, given
by the imaginary part I of the nondimensional eigenvalue associated with the considered modes at several
prestress levels induced in the wing at dierent free-stream velocities U and for dierent positions xs of
the store mass along the wing span. These curves are obtained by linearizing the eigenvalue problem about
the deformed conguration B 0 of the wing under the static aeroelastic loads, attained at the velocity U ,
whose expressions are given by
L0 (x) =

1
2
(2b)U
CLST (03 ) ,
2

1
2
D0 (x) = (2b)U
CDST (03 ) ,
2

MA0 (x) =

9 of 14
American Institute of Aeronautics and Astronautics

1
2
ST
(2b)2 U
CM
(03 )
2

(10)

1
0.26

Linear wing model


Nonlinear wing model

0.22

0.8

0.18

[deg]

0.1

0.4

0
3

u02 [m]

0.6
0.14

0.06

0.2

0.02

Linear wing model


Nonlinear wing model

-0.02
-0.06

-0.2
0

10

15

20

25

30

35

40

10

15

20

U [m/s]

25

30

35

40

U [m/s]

(a)

(b)

1.1
1
1.05
0.8

0
I

0
I

/
I

0.6
0.95

0.4

0.9

Twisting mode
Lagging mode
Flapping mode

0.85

Twisting mode
Lagging mode
Flapping mode

0.2

0.8

0
0

10

15

20

25

30

35

40

0.1

0.2

0.3

0.4

U [m/s]

0.5

0.6

0.7

0.8

0.9

xs / l

(c)

(d)

Figure 7. NACA6315 lifting surface: equilibrium path under static aeroelastic loads for a root angle of attack
W = 0 , tip apping displacement u02 (a) and twisting angle 03 (b) with store mass at the wing tip. Variation
of the lowest nondimensional frequency of the apping, lagging and twisting modes for dierent wind speeds
and with the store mass at the tip (c) and for dierent store positions at zero wind speed U = 0 m/s (d).
In part (c), I0 is the natural wing frequency of the associated mode at zero velocity U = 0; in part (d) it
represents the frequency of the wing without the presence of the store mass.

Note that the static aeroelastic coecients are considered in their nonlinear expressions obtained by assuming
the static curves shown in Fig. 4.
It is worth noticing that the eigenvalue analysis is carried out in terms of the incremental kinematic
parameters u and i . In Figure 8 the 3D aeroelastic wing congurations are shown at dierent free-stream
speeds: 10 m/s, 20 m/s and 30 m/s.

V.

Critical Flutter Speed Evaluation

The critical velocity at the onset of the utter condition is rst evaluated considering the lift, drag,
and aerodynamic moment curves linearized around a zero angle of attack. To this aim, a complex-valued
eigenvalue problem is formulated and solved for a certain range of ight speeds. In particular, the prestressed
wing conguration, under the action of the static aeroelastic loads given by (10), is evaluated at dierent
values of U0 . The eigenvalue problem is formulated for each ight conguration, where the wing stiness
properties are modied by the eects of the geometric nonlinearities, and the lowest frequencies of the
apping, lagging, and twisting modes are evaluated. The critical utter speed has been computed according
to four dierent approaches. In the rst approach, a fully linear structural model is considered along with

10 of 14
American Institute of Aeronautics and Astronautics

(a)

(b)

(c)

Figure 8. NACA6315 lifting surface with the store mass at the wing tip: equilibrium congurations at dierent
free-stream velocities U0 = 10 m/s (a), 20 m/s (b), 30 m/s (c).

the aerodynamic coecients CAE linearized about the stress-free conguration (Linear Model ) as

w(x, t)
CAE
0

CAE = CAE
+
3 3 =0
U

(11)

This approach, which fully neglects the prestress due to the wing/store weights and to the static part of the
aeroelastic forces, leads to a utter speed of 32.5 m/s. With the second approach, the nonlinear structural
model accounts for the wing prestressed conguration under self-weight and store weight (S-W ), while the
aerodynamic coecients are linearized about such conguration according to

w(x, t)
CAE
W
0
0
W
(12)
CAE = CAE +
3 (x, t) where 3 (x, t) = 3 (x) +

U
3 3 =03
The critical velocity is slightly higher, Ucr = 33.8 m/s.
With the third approach, the fully nonlinear structural and aerodynamic models are considered whereby
the wing nonlinear static equilibrium induced by the free-stream speed U is rst computed and then
the eigenvalues of the perturbed problem are calculated to determine the critical condition at which the
eigenvalues cross the imaginary axis. This approach results in a utter speed of 40.1 m/s.
Finally, to better demonstrate the eects of the evaluation of the nonlinear prestress condition on utter
and appreciate the dierences with respect to the former (exact) approach, a dierent evaluation of the
critical condition is performed. Each equilibrium conguration B0 attained under the ow velocity U0 is
considered as an initial conguration about which the problem is linearized. This approach leads to dierent
eigenvalue problems in the incremental velocity from U0 in the range [U0 , 45 m/s] thus determining the
incremental velocity which causes the onset of utter as that at which the eigenvalues traverse the imaginary
axis. Clearly, this approach is in a sense an intermediate approach between the second and third approaches
and consists in neglecting the further updates in the wing stiness and aerodynamic coecients during the
incremental increase of ow speed from U0 . Variation of the utter speed of the twisting mode with the
initial speed U0 calculated via the fourth approach is shown in Fig. 9 (d).
The paths of the critical eigenvalue = R I , corresponding to the lowest apping and twisting
mode, are shown in Figs. 9 (a) and 9 (b), respectively, and are obtained linearizing the equations of motion
around three dierent prestress conditions, attained at the ight speeds U0 = 10 m/s, 20 m/s, 30 m/s.
The paths are obtained incrementing the speed but keeping the static conguration xed to that attained
under the ow speed U0 . The crossing of the imaginary axis signals the utter condition. Figure 9 (c)
shows variation of the real part of the eigenvalues associated with the apping, lagging, and twisting modes
obtained with the incremental approach starting from the prestress conditions attained at the ight speeds
U0 = 10 m/s, 20 m/s, 30 m/s.
It is interesting to note that the utter speed evaluated by accounting for the full nonlinearities is higher
than the linear prediction (stress-free) by 23.4% while it is higher than the critical velocity obtained by
accounting for the wing and store weights only by 18.6%. On the other hand, the evaluation of the utter
speed obtained by an incremental approach (based on linearization about each conguration B 0 at U0 ) leads
to higher or lower values with respect to the utter speed calculated by the linearized approach about the
stress-free conguration.
More extensive results, addressing the trim ight dynamic condition as well as the post-utter analysis
of the considered HALE wing excited by unsteady aerodynamic loads and exhibiting dynamic stall will be
presented elsewhere.
11 of 14
American Institute of Aeronautics and Astronautics

22.5

Flapping

2.2

Twisting
22

2
21.5
1.8

21

20.5

1.6

Linear Model
U0 = 10 m/s

1.4

U0 = 20 m/s
1.2

Linear Model
U0 = 10 m/s

20

U0 = 20 m/s

19.5

U0 = 30 m/s

U0 = 30 m/s
19

-2

-1.8

-1.6

-1.4

-1.2

-1

-0.8

-0.6

-0.4

-0.2

-0.75

-0.6

-0.45

-0.3

-0.15

0.15

0.3

0.45

(a)

(b)

42

Twisting

0.5

Nonlinear Model

40.1 m/s

40

Ucr [m/s]

38
-0.5

Lagging

-1

Linear Model
U0 = 10 m/s

-1.5

33.8 m/s

S-W
34

Flapping

U0 = 20 m/s

-2

U0

36

32

32.5 m/s

Linear Model

U0 = 30 m/s
-2.5

30
5

10

15

20

25

30

35

40

13

17

21

25

29

33

37

41

U0 [m/s]

U [m/s]
(c)

(d)

Figure 9. Critical eigenvalues paths: lowest apping mode u2 (a) and lowest twisting mode 3 (b). Flutter
(gray zone) and stability (white zone) regions: variation of the twisting lowest eigenvalue real part increasing
the free-stream velocity U (c). Critical utter speed U for dierent aeroelastic congurations at U0 (d).

VI.

Concluding Remarks and Future Outlook

A geometrically exact wing model is proposed for high-aspect-ratio wings coupled with an incompressible
unsteady aerodynamic formulation accounting for ow separation and dynamic stall which can result in
large-amplitude motions. The goal of this research program is to develop tools that can provide realistic
aeroelastic simulations for HALE wings exhibiting large deformations during their ight. To assess the
delity of the developed aeroelastic model, several initial tests have been performed including comparison
with experimental ndings for the modal analysis of a wing carrying an external store mounted at the
wing tip. Parametric studies illustrating how the geometric nonlinearities aect the static and modal wing
responses are summarized.
The primary objective of this work is to seek a better understanding of the nonlinear wing aeroelastic
behavior and how the static aerodynamic loads aect the wing dynamic response to external disturbances,
as well as the utter behavior. Successive studies based on the physical based modied nonlinear state-space
Beddoes-Leishman model will address dynamic stall and LCO in the post-utter regime. Particular eort is
devoted to investigate the sub-critical behavior of such wings in the proximity of the utter boundary due to
the presence of both aerodynamic and structural nonlinearities. This is possible by implementing the exact
nonlinear relationships, without ad-hoc approximations, between the structural and aerodynamic models
thus overcoming errors associated with the static oset and the large structural deformations. The proposed
fully nonlinear model of HALE wings becomes a suitable parametric framework within which aeroelastic limit
states, associated with utter and post-utter, can be suitably investigated; moreover, ecient structural
optimization studies can be eectively conducted. Parametric studies will be carried out, in terms of wing

12 of 14
American Institute of Aeronautics and Astronautics

(a)

(b)

Figure 10. Static conguration at the utter velocity (a), utter mode shape (b).

aspect ratio variations, stiness ratios, and masses distribution, along with dynamic variables such as gust
wave-front magnitude, and comparison with experimental data, when available.

VII.

Acknowledgments

The authors would like to thank the Sapienza PHD Fellowship Program which supports AA and the PRIN
Shape-memory alloy advanced modeling for industrial and biomedical applications from MIUR granted to
WL. Moreover, the authors gratefully acknowledge the ARMY Research Oce, grant W911NF-05-1-0339,
and The National Science Foundation, grant NSF-CMMI-1031036, for providing partial support for this
research.

References
1 Woolston, D. S., Runyan, H. L., and Andrews, R. E., An Investigation of Eects of Certain Types of Structural
Nonlinearities on Wing and Control Surface Flutter, Journal of Aeronautical Sciences, Vol. 24, 1957, pp. 5763.
2 Shen, S. F., An Approximate Analysis of Nonlinear Flutter Problems, Journal of Aeronautical Sciences, Vol. 26, 1959,
pp. 2532.
3 Tang, D. M. and Dowell, E. H., Eects of geometric structural nonlinearity on utter and limit cycle oscillations of
high-aspect-ratio wings, Journal of Fluids and Structures, Vol. 19, 2004., pp. 291306.
4 Patil, M. J. and Hodges, D., On the importance of aerodynamic and structural geometrical nonlinearities in aeroelastic
behavior of high-aspect-ratio wings, Journal of Fluids and Structures, Vol. 19, 2004, pp. 905915.
5 Frulla, G., Cestino, E., and Marzocca, P., Critical Behavior of Slender Wing Congurations, Proceedings of the Institution of Mechanical Engineers, Part G: Journal of Aerospace Engineering, Vol. 224, 2009.
6 et alt., E. H. D., A Modern Course in Aeroelasticity. Fourth Revised and Enlarged Edition, Kluwer Academic Publishers,
4th ed., 2004.
7 Lighthill, M. J., On Boundary Layers and Upstream Inuence. II. Supersonic Flows without Separation, Proceedings
R. Soc. Lond., Vol. 217, May 1953, pp. 478507.
8 Librescu, L., Aeroelastic Stability of Orthotropic Heterogeneous Thin Panels in the Vicinity of the Flutter Critical
Boundary, Part I, Journal de Mcanique, Vol. 4, 1965, pp. 5176.
9 Librescu, L., Aeroelastic Stability of Orthotropic Heterogeneous Thin Panels in the Vicinity of the Flutter Critical
Boundary, Part II, Journal de Mcanique, Vol. 6, 1967, pp. 133152.
10 E. H. Dowell, M. I., Studies in Nonlinear Aeroelasticity, New York, Springer, Verlag, 1988.
11 Abbas, L. K., Chen, Q., Marzocca, P., ODonnell, K., and Valentine, D., Aeroelastic Behavior of the Lifting Surfaces with
Free-Play and Aerodynamic Stiness and Damping Nonlinearities, International Journal of Bifurcation and Chaos, Vol. 18,
No. 4, 2008, pp. 11011126.
12 Librescu, L., Chiocchia, G., and Marzocca, P., Implications of Cubic Physical / Aerodynamic Nonlinearities on the
Character of the Flutter Instability Boundary, International Journal of Nonlinear Mechanics, Vol. 38, 2003, pp. 173199.
13 Leishman, J. G. and Beddoes, T. S., A semi-empirical model for dynamic stall, J. Am. Helicopter Soc., Vol. 7, 1989,
pp. 317.
14 Petot, D., Dierential equation modeling of dynamic stall, Rech. Aerosp., Vol. 5, 1989, pp. 5972.

13 of 14
American Institute of Aeronautics and Astronautics

15 van der Wall, B., Analytische Formulierung der instation


aren Prolbeiwerte und deren Anwendung in der Rotorsimulation, Tech. Rep. DLR-FB 90-28, Deutsche Forschungsanstalt f
ur Luft- und Raumfahrt, Braunschweig, 1990.
16 Goman, M. and Khrabrov, A., State-space representation of aerodynamic characteristics of an aircraft at high angles of
attack, J. Aircraft, Vol. 31 (5), 1994, pp. 11091115.
17 Truong, V. K., Prediction of helicopter rotor airloads based on physical modeling of 3-D unsteady aerodynamics, Tech.
rep., ONERA TP, 1996.
18 ye, S., Dynamic stall simulated as time lag of separation, EWEC , Thessaloniki, Greece, October 1994.
19 Hansen, M. H., Gaunaa, M., and Madsen, H. A. A., A Beddoes-Leishman type dynamic stall model in state-space and
indicial formulations, Tech. Rep. R -1354(EN), Ris National Laboratory, 2004.
20 Snel, H., Heuristic modeling of dynamic stall characteristics, Proceedings of the EWEC , 1997, pp. 429433.
21 Tran, C. and Petot, D., Semi-Empirical Model for the Dynamic Stall of Airfoils in View of the Application to the
Calculation of Responses of a Helicopter Blade in Forward Flight, Vertica, Vol. 5, 1981.
22 Dat, D. and Tran, C. T., Investigation of the stall utter of an airfoil with a semi-empirical model of 2-D ow, Vertica,
Vol. 7(2), 1983, pp. 7386.
23 Dunn, P. and Dugundji, J., Nonlinear Stall Flutter and Divergence Analysis of Cantilevered Graphite/Epoxy Wings,
AIAA Journal, Vol. 30, 1992, pp. 153162.
24 Bisplingho, R. L., Ashley, H., and Halfman, H., Aeroelasticity, Addison-Wesley Publishng Company, 1996.
25 Hodges, D. H., A Mixed Variational Formulation Based On Exact Intrinsic Equations For Dynamics Of Moving Beams,
International Journal of Solid and Structures, Vol. 26, No. 11, 1990, pp. 12511273.
26 Hodges, D. H., Geometrically Exact, Intrinsic Theory for Dynamics of Curved and Twisted Anisotropic Beams, AIAA
JOURNAL, Vol. 41, No. 6, 2003, pp. 11311137.
27 Hegemier, G. A. and Nairt, S., A Nonlinear Dynamical Theory for Heterogeneous, Anisotropic, Elastic Rods, AIAA
JOURNAL, Vol. 15, 1976, pp. 815.
28 Antman, S. S., Nonlinear problems of elasticity, New York: Springer-Verlag, 2nd ed., 2005.
29 Lacarbonara, W. and Arena, A., Flutter of an Arch Bridge via a Fully Nonlinear Continuum Formulation, Journal of
Aerospace Engineering, Vol. 24, No. 1, January 2011, pp. 112123.
30 Sotoudeh, Z., Hodges, D. H., and Chang, C., Validation Studies for Aeroelastic Trim and Stability Analysis of Highly
Flexible Aircraft, Journal of Aircraft, Vol. 47, No. 4, JulyAugust 2010, pp. 12401247.
31 Patil, M. J., Hodges, D. H., and Cesnik, C. E. S., Nonlinear Aeroelasticity And Flight Dynamics Of High-altitude
Long-endurance Aircraft, AIAA-99-1470 , 1999.
32 Abbas, L., Chen, Q., Marzocca, P., and Milanese, A., Non-linear Aeroelastic Investigations of Store(s)-Induced Limit
Cycle Oscillations, Journal of Aerospace Engineering, Vol. 222, No. 1, 2008, pp. 6380.
33 Beran, P. S., Strganac, T. W., Kim, K., and Nichkawde, C., Studies of Store-Induced Limit-Cycle Oscillations Using a
Model with Full System Nonlinearities, Nonlinear Dynamics, Vol. 37, No. 4, September 2004, pp. 323339.
34 Brinks, R., On the convergence of derivatives of B-splines to derivatives of the Gaussian function, Comp. Appl. Math.,
Vol. 27, No. 1, 2008.
35 Tang, D. M. and Dowell, E. H., Experimental and theoretical study on aeroelastic response of high-aspect-ratio wings,
AIAA Journal, Vol. 39(8), August 2001, pp. 14301441.
36 Tang, D. M. and Dowell, E. H., Experimental and theoretical study of gust response for high-aspect-ratio wing, AIAA
Journal, Vol. 40(3), March 2002, pp. 419429.
37 Tang, D. M. and Dowell, E. H., Limit-cycle hysteresis for a high-aspect-ratio wing model, Journal of Aircraft, Vol. 39(5),
September-October 2002, pp. 885888.
38 Jaworski, J. W., Nonlinear Aeroelastic Analysis of Flexible High Aspect Ratio Wings Including Correlation With Experiment, Ph.D. thesis, Department of Mechanical Engineering and Materials Science Duke University, 2009.

14 of 14
American Institute of Aeronautics and Astronautics

S-ar putea să vă placă și