You are on page 1of 14

Appl. Math. Mech. -Engl. Ed.

, 33(1), 1–14 (2012)


DOI 10.1007/s10483-012-1529-x
°Shanghai
c University and Springer-Verlag
Berlin Heidelberg 2012

Dynamic analysis of two-degree-of-freedom airfoil with freeplay and


cubic nonlinearities in supersonic flow∗

Hu-lun GUO (HmÔ), Yu-shu CHEN (ƒL)


(School of Astronautics, Harbin Institute of Technology, Harbin 150001, P. R. China)

(Contributed by Yu-shu CHEN)

Abstract The nonlinear aeroelastic response of a two-degree-of-freedom airfoil with


freeplay and cubic nonlinearities in supersonic flows is investigated. The second-order
piston theory is used to analyze a double-wedge airfoil. Then, the fold bifurcation and the
amplitude jump phenomenon are detected by the averaging method and the multi-variable
Floquet theory. The analytical results are further verified by numerical simulations.
Finally, the influence of the freeplay parameters on the aeroelastic response is analyzed
in detail.
Key words freeplay, cubic nonlinearity, averaging method, limit cycle oscillation,
amplitude jump phenomenon
Chinese Library Classification V215.3+ 4, O322
2010 Mathematics Subject Classification 34C15, 70K50

1 Introduction
Strong interactions may occur between high speed flow fields and aerospace structural com-
ponents, resulting in several remarkable aeroelastic phenomena. Such phenomena dramatically
influence the performance of the flight vehicles. Moreover, the tendency of the next genera-
tion aeronautical and space vehicles is of low weight and high structural flexibility[1–2] . These
requirements make the vehicles subject to complex nonlinearities.
The most prevalent nonlinearities in structures are the cubic stiffening and the freeplay in
control surfaces. The freeplay usually occurs in the actuated degree-of-freedom (DOF), i.e., in
the control surfaces with loose joints or in an advanced generic missile pin that is foldable at
the settled position[3] . The cubic stiffening mainly comes from the large amplitude oscillation
of flexible wings.
With the existence of such nonlinearities, a Hopf bifurcation occurs at the linear flutter
Mach number. The limit cycle oscillation (LCO) may appear as the Mach number increases
instead of an aeroelastic instability phenomenon. Such large amplitudes and high frequencies
associated with flutter phenomena can cause the catastrophic structural failure. Therefore, the
flutter prevention is an important technology in the wing design of flight vehicles.
Recent studies on the aeroelasticity of the airfoil system have focused on the identification
of the LCOs of strongly nonlinear airfoils with freeplay. Most of the studies have focused on the

∗ Received Jun. 2, 2011 / Revised Sept. 18, 2011


Corresponding author Yu-shu CHEN, Professor, E-mail: yschen@hit.edu.cn
2 Hu-lun GUO and Yu-shu CHEN

numerical or analytical algorithms for solving the nonlinear aeroelastic equations, including the
point transformation (PT) method[4] , the numerical continuation algorithm[5] , the center mani-
fold method[6–7] , the perturbation incremental (PI) method[8] , the homotopy analysis method[9] ,
and the incremental harmonic balance method[10] . The describing function method was applied
successfully to the flutter analyses of an aircraft for a variety of structural nonlinearities[11] .
Shen[12] discussed the suitability and application of the Krylov-Bogoliubov (KB) method to
nonlinear flutter problems. Yang[13] used the Krylov-Bogoliubov-Mitropolsky (KBM) method
to analyze the LCOs of an airfoil with the freeplay nonlinearity.
Furthermore, some studies have shown the effects of various structural nonlinearities on
aeroelastic characteristics over the past decades. Kim and Lee[14] investigated the dynamics of
a flexible airfoil with the freeplay nonlinearity. They observed that the LCO and the chaotic
motion were highly influenced by the pitch-plunge frequency ratio. A complete methodology was
presented by Dimitriadis[15] , which successfully predicted the bifurcations and post-bifurcations
for the nonlinear subsonic aircraft. The Poincaré mapping method and the Floquet theory[16]
were adopted to analyze the LCO flutter and chaotic motion of a two-DOF airfoil system with
a combination of freeplay and cubic nonlinearities. Conner et al.[17] , Tang and Dowell[18] , Tang
et al.[19–20] , and Liu and Dowell[21] conducted both experimental and analytical studies to reveal
the nonlinear flutter behaviors of a three-DOF typical-section airfoil with the control surface
freeplay. In these studies, both the direct time integration and the describing function (harmonic
balance) methods were used to predict the LCO amplitude, and the results were compared with
the experimental data. A fold bifurcation followed by a supercritical Hopf bifurcation, which
caused a jump phenomenon, was detected in both the analytical results and the experimental
data.
It should be noted that for an aeroelastic system with the freeplay nonlinearity, the time
integration method with the uniform time step cannot precisely predict the switching point,
where the variation in linear regions occurs. The importance of capturing switching points was
illustrated by Lin and Cheng[22] and Conner et al.[17] . However, the time integration method is
very costly and time consuming, and the unstable LCO cannot be found by the time integration
method. To address the above-mentioned issues, in this study, we adopt the averaging method to
analytically determine the switching point. The averaging method can well predict the dynamic
response for a nonlinear system. After averaging the governing equation over a period, we can
obtain both the stable and the unstable periodic solutions and predict the switching point.
In this paper, a two-DOF plunging/pitching double-wedge airfoil with freeplay and cubic
nonlinearities in the pitch DOF in supersonic flows is studied. The unsteady aerodynamic lift
and moment are obtained by the piston theory aerodynamics (PTA). Then, the LCO amplitude
is calculated by the averaging method. Also, the multi-variable Floquet theory is used to analyze
the stability of the LCOs. Furthermore, the influence of the freeplay parameters on the LCO
response is discussed in detail.

2 Model equations

Consider the vibration of a two-DOF plunging/pitching double-wedge airfoil shown in Fig. 1.


The nonlinear aeroelastic governing equations can be written as
{
mḧ + Sα α̈ + ch ḣ + Kh h = −L,
(1)
Sα ḧ + Iα α̈ + cα α̇ + Kα F (α) = MEA ,

where h is the plunging displacement (positive downward), α is the pitch angle (positive nose
up), the superposed dots denote the derivatives with respect to the time t, m is the structural
mass per unit span, Sα is the static unbalance about the elastic axis, Iα is the mass moment of
inertia about the elastic axis of the airfoil, ch , cα , Kh , and Kα are the linear viscous damping
Dynamic analysis of two-degree-of-freedom airfoil 3

and stiffness coefficients, and F (α) is the structural nonlinear function in the pitch DOF. L
and MEA represent the unsteady aerodynamic lift and moment that can be obtained by the
PTA[23] .

Fig. 1 Two-DOF double-wedge airfoil geometry

The PTA is an inviscid unsteady


√ aerodynamic theory, used extensively in the cases of super-
sonic/hypersonic flows (M∞ > 2) and moderate angles-of-attack (|α| 6 π/6), which provides
a point-function relationship between the local pressure on the surface of the vehicle and the
component of the fluid velocity normal to the moving surface[24] .
In this paper, the second-order piston theory is adopted. Thus, the unsteady aerodynamic
lift and moment of a two-DOF plunging/pitching double-wedge airfoil can be expressed as[25]
( ḣ α̇ ) ( α
b )
L = 4p∞ γM∞ b − ba + α − p∞ γ(γ + 1)M∞ b τb
2 2
, (2)
V∞ V∞ V∞
( ( ḣ (b ) α̇ )
MEA = p∞ γM∞ b2 4 a − + ba2 + aα
V∞ 3 V∞
( ḣ α̇ ))
+ (γ + 1)M∞ τb − 2ba +α . (3)
V∞ V∞
Denote the dimensionless parameters as follows:
√ √
h V∞ t Sα Iα Kh Kα ch
ξ= , τ= , χα = 2
, rα = , ω ξ = , ω α = , ζξ = √ ,
b b mb mb2 m Iα 2 Kh m
cα V∞ ωξ m th ρ∞ c2∞
ζα = √ , V∗ = , ω= , µ= 2
, τb = , p∞ = .
2 Kα Iα bωα ωα 4ρ∞ b b γ
The governing equations of a supersonic/hypersonic double-wedge airfoil featuring the plung-
ing/pitching coupled motion can be described as


 ζξ ω ω2
 ξ 00 + χα α00 + 2 ∗ ξ 0 + ∗2 ξ = −L(τ ),
V V
(4)

 χ ζ 1
 2 ξ 00 + α00 + 2 ∗ α0 + ∗2 F (α) = M EA (τ ),
α α
rα V V
where
1
L(τ ) = (4(ξ 0 − aα0 + α) − (γ + 1)b τ M∞ α0 ),
4µM∞
1 ( ( (1 ) ) )
0 0 0 0
M EA (τ ) = 4 aξ − + a2
α + aα + (γ + 1)b
τ M ∞ (ξ − 2aα + α) .
4µM∞ rα2 3
4 Hu-lun GUO and Yu-shu CHEN

Here, the freeplay and cubic nonlinearities are both considered. Assume that the structural
nonlinear function is
F (α) = F1 (α) + F2 (α), (5)

where F1 (α) is the freeplay nonlinearity, and F2 (α) is the cubic nonlinearity.
The freeplay nonlinearity is defined by[26]

 M0 + α − αf , α < αf ,

F1 (α) = M0 + Mf (α − αf ), αf 6 α 6 αf + δ, (6)


M0 + α − αf + δ(Mf − 1), α > αf + δ,

where M0 , Mf , αf , and δ are constants. A sketch of the freeplay model is given in Fig. 2.
According to Eq. (6), the displacement in pitch can be divided into three regions, i.e., Φ1 , Φ2 ,
and Φ3 .
The cubic nonlinearity is defined by

F2 (α) = ηbα α3 , (7)

where ηbα is the nonlinear stiffness coefficient.


The nonlinear function with freeplay and cubic nonlinearities versus the pitch displacement
is shown in Fig. 3 when the nonlinear parameters are Mf = 0.1, αf = 0.1, δ = 0.1, and ηbα = 10.
It is shown that the main nonlinearity is the cubic nonlinearity, and the freeplay nonlinearity
can change the moment slightly in values. The effect of the freeplay on the dynamic response
characteristic is mainly analyzed in this paper.

Fig. 2 General sketch of freeplay stiffness Fig. 3 Freeplay and cubic nonlinearities

Let x1 = ξ, x2 = ξ 0 , x3 = α, and x4 = α0 . Equation (4) can be written as a set of four


first-order ordinary differential equations x0 = f (x), i.e.,
 0

 x1 = x2 ,



 1 χα F (x3 )

 x02 = (N11 x1 + N12 x2 + N13 x3 + N14 x4 ) + ,
rb rbV ∗2 (8)

 x03 = x4 ,





 x04 = 1 (N21 x1 + N22 x2 + N23 x3 + N24 x4 ) − F (x3 ) .
rb rbV ∗2
Dynamic analysis of two-degree-of-freedom airfoil 5

The coefficients rb and Nij (i = 1, 2 and j = 1, 2, 3, 4) are the functions of the system
parameters and the Mach number, and their expressions are given as

χ2α ω2
rb = 1 − 2
, N11 = − ∗2 ,
rα V
ωζξ 1
N12 = −2 ∗ − (4aχα + 4rα2 + χα (γ + 1)b
τ M∞ ),
V 4µM∞ rα2
1
N13 = − (4aχα + 4rα2 + χα (γ + 1)b
τ M∞ ),
4µM∞ rα2
χα ζα 1 χα
N14 = 2 + (4a + (γ + 1)b
τ M∞ ) + (2(1 + 3a2 ) + 3a(γ + 1)b
τ M∞ ),
V∗ 4µM∞ 6µM∞ rα2
χα ω 2 χα ωζξ 1
N21 = , N22 = 2 + (4χα + 4a + (γ + 1)b
τ M∞ ),
rα2 V ∗2 rα2 V ∗ 4µM∞ rα2
1
N23 = (4χα + 4a + (γ + 1)b
τ M∞ ),
4µM∞ rα2
ζα χα 1
N24 = −2 − τ M∞ ) −
(4a + (γ + 1)b (2(1 + 3a2 ) + 3a(γ + 1)b
τ M∞ ).
V∗ 4µM∞ rα2 6µM∞ rα2

3 Analytical analysis

The applications of averaging techniques to free, forced, and self-excited vibration problems
with a variety of nonlinearities, including some problems with relay characteristics, were given
by Nayfeh and Mook[27] and Nayfeh[28] . In this section, the averaging method is used to solve
the periodic solution of the airfoil system. Note that the LCOs exist in the nonlinear airfoil
system. Thus, the solution of Eq. (8) takes the periodic form, i.e.,


 x1 = a1 cos ϕ1 ,




 x2 = −ωa1 sin ϕ1 ,
(9)




x3 = a2 cos ϕ2 ,



x4 = −ωa2 sin ϕ2 ,

where ϕ1 = ωτ + θ1 , and ϕ2 = ωτ + θ2 .
Substituting Eq. (9) into Eq. (8) and simplifying the equations yield

 rbω ȧ1 = −b
rω 2 a1 sin ϕ1 cos ϕ1 − N11 a1 cos ϕ1 sin ϕ1 + N12 ωa1 sin2 ϕ1



 χα

 − N13 a2 cos ϕ2 sin ϕ1 + N14 ωa2 sin ϕ2 sin ϕ1 − ∗2 F sin ϕ1 ,

 V



 rbωa1 θ̇1 = −b
rω 2 a1 cos2 ϕ1 − N11 a1 cos2 ϕ1 + N12 ωa1 sin ϕ1 cos ϕ1



 χα

 − N13 a2 cos ϕ2 cos ϕ1 + N14 ωa2 sin ϕ2 cos ϕ1 − ∗2 F cos ϕ1 ,
 V
(10)

 rbω ȧ2 = −b
rω 2 a2 sin ϕ2 cos ϕ2 − N21 a1 cos ϕ1 sin ϕ2 + N22 ωa1 sin ϕ1 sin ϕ2



 1

 − N23 a2 cos ϕ2 sin ϕ2 + N24 ωa2 sin2 ϕ2 + ∗2 F sin ϕ2 ,

 V



 rbωa2 θ̇2 = −b 2 2


 r ω a2 cos ϕ 2 N 21 a1 cos ϕ 1 cos ϕ 2 + N22 ωa1 sin ϕ1 cos ϕ2



 1
− N23 a2 cos2 ϕ2 + N24 ωa2 sin ϕ2 cos ϕ2 + ∗2 F cos ϕ2 .
V
6 Hu-lun GUO and Yu-shu CHEN

When the oscillation of the airfoil processes the LCO, the slow dynamics of a1 , θ1 , a2 , and
θ2 can be governed by

 1 1 1 1

 rbω ȧ1 = N12 ωa1 + N13 a2 sin θ + N14 ωa2 cos θ + χα Z(a2 )a2 sin θ,

 2 2 2 2



 1 1 1 1 1
 rbωa1 θ̇1 = − rbω 2 a1 − N11 a1 − N13 a2 cos θ + N14 ωa2 sin θ − χα Z(a2 )a2 cos θ,
2 2 2 2 2 (11)

 rbω ȧ2 = − 1 N21 a1 sin θ + 1 N22 ωa1 cos θ + 1 N24 ωa2 ,



 2 2 2



 rbωa2 θ̇2 = − rbω a2 − N21 a1 cos θ − N22 ωa1 sin θ − 1 N23 a2 + 1 Z(a2 )a2 ,
1 2 1 1
2 2 2 2 2

where θ = ϕ2 − ϕ1 = θ2 − θ1 , and
∫ 2π ∫ 2π
2 1 2 1
Z(a2 ) = ∗2
F cos ϕ2 d(ωτ ) = ∗2
(F1 + F2 ) cos ϕ2 d(ωτ ).
a2 V 2π 0 a2 V 2π 0

The integration of the part of the cubic nonlinearity is


∫ 2π ∫ 2π
1 1 3
F2 cos ϕ2 d(ωτ ) = ηbα a32 cos4 ϕ2 d(ωτ ) = ηbα a32 .
2π 0 2π 0 8

The integration of the freeplay function can be discussed according to the following three cases.
Case (i) a2 > αf + δ
In this case, the LCO exists in the regions Φ1 , Φ2 , and Φ3 (see Fig. 2). Thus,
∫ 2π ∫ ϕf+δ
1 1
F1 cos ϕ2 d(ωτ ) = (M0 + a2 cos ϕ2 − αf + δ(Mf − 1)) cos ϕ2 d(ωτ )
2π π
0 0
∫ ϕf
1
+ (M0 + Mf (a2 cos ϕ2 − αf )) cos ϕ2 d(ωτ )
π ϕf+δ
∫ π
1
+ (M0 + a2 cos ϕ2 − αf ) cos ϕ2 d(ωτ )
π ϕf
(Mf − 1)a2 ( 1 ) 1
ϕf − ϕf+δ − (sin(2ϕf ) − sin(2ϕf+δ )) + a2 ,
=
2π 2 2
( α +δ )
where ϕf = arccos αa2 , and ϕf+δ = arccos
f f
a2 .
Case (ii) αf 6 a2 6 αf + δ
In this case, the LCO exists in the regions Φ1 and Φ2 , and ϕf+δ = 0. Thus,

1 2π
(Mf − 1)a2 ( 1 ) 1
F1 cos ϕ2 d(ωτ ) = ϕf − sin(2ϕf ) + a2 .
2π 0 2π 2 2

Case (iii) a2 < αf


In this case, the LCO only exists in the region Φ1 , and ϕf = ϕf+δ = 0. Thus,
∫ 2π
1 1
F1 cos ϕ2 d(ωτ ) = a2 .
2π 0 2

Note that the freeplay does not exist in this case. Therefore, the integrated nonlinear
Dynamic analysis of two-degree-of-freedom airfoil 7

function Z(a2 ) can be expressed as



 1 3

 + ηb a2 , a2 < α f ,

 V ∗2 4V ∗2 α 2



 1 + 3 ηb a2 − 1−M ( f 1 )
α 2 ϕf − sin(2ϕ f ) , αf 6 a2 6 αf + δ,
Z(a2 ) = V ∗2 4V ∗2 πV ∗2 2 (12)

 1 − Mf ( )
 1 + 3 ηb a2 −
 − −
1 1

 α 2 ϕ f ϕ f+δ sin(2ϕ f ) + sin(2ϕ f+δ ) ,
 V ∗2
 4V ∗2 πV ∗2 2 2

a2 > αf + δ.

It is important to note that M0 does not exist in the integrated nonlinear function. Thus,
we can conclude that M0 is independent of the LCOs of an airfoil.
When the motion of the airfoil is an LCO, we can assume a1 , θ1 , a2 , and θ2 to be constants,
i.e., ȧ1 = θ̇1 = ȧ2 = θ̇2 = 0. Thus, Eq. (11) can be rewritten as follows:


 N12 ωa1 + N13 a2 sin θ + N14 ωa2 cos θ + χα Z(a2 )a2 sin θ = 0,


 − rbω 2 a − N a − N a cos θ + N ωa sin θ − χ Z(a )a cos θ = 0,
1 11 1 13 2 14 2 α 2 2
(13)

 − N21 a1 sin θ + N22 ωa1 cos θ + N24 ωa2 = 0,



− rbω 2 a2 − N21 a1 cos θ − N22 ωa1 sin θ − N23 a2 + Z(a2 )a2 = 0.

It is convenient to use the following transformation:


{
b1 = a1 cos θ,
(14)
b2 = a1 sin θ.

Then, Eq. (13) can be rewritten as

rbω 2 b2 + N11 b2 − N12 ωb1 − N14 ωa2 = 0, (15)


rbω b1 + N11 b1 + N12 ωb2 + N13 a2 + χα Z(a2 )a2 = 0,
2
(16)
− N21 b2 + N22 ωb1 + N24 ωa2 = 0, (17)
rbω 2 a2 + N21 b1 + N22 ωb2 + N23 a2 − Z(a2 )a2 = 0. (18)

Substituting Eq. (17) into Eqs. (15), (16), and (18) and eliminating b2 yield
( ) ( )( )
N21 N12 b1 + N14 a2 − rbω 2 + N11 N22 b1 + N24 a2 = 0, (19)
(( ) ( ) ) ( )
N21 rbω 2 + N11 b1 + χα Z + N13 a2 + N12 ω 2 N22 b1 + N24 a2 = 0, (20)
(( 2 ) ) ( )
rbω − Z + N23 a2 + N21 b1 N21 + N22 ω 2 N22 b1 + N24 a2 = 0. (21)

From Eqs. (20) and (21), the square dimensionless frequency can be expressed as

R1 + R2 Z(a2 )
ω2 = . (22)
R3

From Eqs. (19) and (20), the following equation is obtained:

R4 ω 4 + (R5 + R6 Z(a2 ))ω 2 + R7 + R8 Z(a2 ) = 0. (23)


8 Hu-lun GUO and Yu-shu CHEN

The expressions for the coefficients Ri (i = 1, 2, · · · , 8) are given as

R1 = N14 N21 − N11 N24 + N22 N13 − N12 N23 ,

R2 = N12 + χα N22 , R3 = rb(N12 + N24 ),

R4 = rb2 N22 , R5 = rb(N11 N22 + N22 N23 − N21 N12 − N21 N24 ) + N22 (N14 N22 − N12 N24 ),

R6 = −b
rN22 , R7 = N23 (N11 N22 − N12 N21 ) + N21 (N14 N21 − N11 N24 ),

R8 = N12 N21 − N11 N22 .

Substituting Eq. (22) into Eq. (23) yields

Q1 Z 2 (a2 ) + Q2 Z(a2 ) + Q3 = 0. (24)

The expressions for the coefficients Qi (i = 1, 2, 3) are given as

Q1 = R22 R4 + R2 R3 R6 ,

Q2 = 2R1 R2 R4 + R1 R3 R6 + R2 R3 R5 + R32 R8 ,

Q3 = R12 R4 + R1 R3 R5 + R32 R7 .

Thus, the integrated nonlinear function can be solved from Eq. (24) as follows:

−Q2 ± Q22 − 4Q1 Q3
Z(a2 ) = . (25)
2Q1

Substituting Eq. (25) into Eq. (12) yields the amplitude of the LCO in the pitch DOF. Then,
by substituting a2 back into Eq. (13), the amplitude of the LCO in the plunge DOF can also
be calculated analytically.

4 Stability analysis

The T -period steady state solution X0 of Eq. (8) is determined by Eq. (9). The parameters
of Eq. (9) are solved in the previous section. The local stability of X0 can be determined by the
multi-variable Floquet theory. Substitute the perturbation solution X = X0 + δX into Eq. (8),
and ignore the higher order terms in δX. Then, the linear variational equation with periodic
coefficients is obtained as
δX 0 = A(τ )δX, (26)
where A(τ ) = DX f (X0 ).
Therefore, the stability of the periodic solution of the system (8) is equivalent to the stability
of the zero solution of the system (26). The stability of the zero solution of the system (6) is
determined by the transition matrix Φ according to the Floquet theory. The actual transition
matrix can be approximated by an approximate expression[29–30] .
Each period T = 2π T
ω is divided into N intervals. The size of each interval is ∆ = N , and
the kth interval is denoted by ((k − 1)∆, ∆). In the kth interval, the periodic coefficient matrix
A(τ ) can be replaced by a constant matrix Bk defined by
 
∫ k∆ 0 rb 0 0
1 1  N11 N12 N13 + χα R N14 
Bk = A(τ )dτ =  , (27)
∆ (k−1)∆ rb  0 0 0 rb 
N21 N22 N23 − R N24
Dynamic analysis of two-degree-of-freedom airfoil 9
ηα a22
3b (N (( ) 2π ) )
where R = 2V ∗2 2π cos 2k − 1 N + 2θ2 sin 2π
N +1 + f
V ∗2 , and
 ( )
 2π

 1, a cos (2k − 1) + θ 2 < αf ,


2
N
 ( )

f = Mf , αf 6 a2 cos (2k − 1) + θ2 6 αf + δ,

 N

 ( )

 1, a2 cos (2k − 1) 2π + θ2 > αf + δ.
N
Note that the aeroelastic model of the airfoil described in Eq. (8) is a nonlinear autonomous
system. We can assume that θ2 = 0. Thus, the coefficient R can be simplified as

ηα a22 ( N
3b ( 2π ) 2π ) f
R= ∗2
cos (2k − 1) sin + 1 + ∗2 ,
2V 2π N N V
where
 (
 2π )

 1, a2 cos (2k − 1) < αf ,

 N
 ( 2π )
f = Mf , αf 6 a2 cos (2k − 1) 6 αf + δ,

 N

 ( )

 1, a2 cos (2k − 1) 2π > αf + δ.
N
Hence, the approximate transition matrix Φ is given in the following form:
N (
∏ ∑ (∆Bi )j )
Nj
Φ= I+ , (28)
i=1 j=1
j!

where Nj is the number of the terms in the approximation of the constant exponential matrix
Bi , and I is the identity matrix.
The eigenvalues of the monodromy matrix Φ are also called the Floquet multipliers, which
can be used to determine the stability of a steady state solution. If all of the moduli of the
eigenvalues are smaller than unity, the solution will be stable. If the modulus of one of the
eigenvalues is larger than unity, the solution will be unstable.

5 Results and discussion

In the above sections, the periodic solution and its stability are studied. In the following,
the dynamic response of the aeroelastic system will be investigated.
In this paper, unless otherwise specified, the system parameters under consideration are
rα = 0.5, χα = 0.25, ςξ = ςα = 0.05, c∞ = 300 m/s, b = 1 m, µ = 100, γ = 1.4, a = −0.15,
τb = 0.05, ωα = 80, ω = 0.8, and ηbα = 10, while the values of the freeplay parameters Mf , αf ,
and δ vary.
When the freeplay region is sufficiently small such that its effect can be neglected, i.e., δ = 0,
there is only the cubic nonlinearity in this aeroelastic system. The dynamic response of this
case is shown in Fig. 4. Figure 4(a) is obtained by the averaging method. It can be seen that an
LCO appears after the supercritical Hopf bifurcation (flutter) point P1 at M∞ = 2.1. The LCO
amplitude increases as the Mach number increases. A numerical result is obtained in Fig. 4(b).
It is shown that the bifurcation point P10 (M∞ = 2.1) and the obtained LCO amplitude are in
good agreement with the results obtained by the averaging method.
For the nonlinear system with the freeplay parameters Mf = 0.1, δ = 0.1, and αf = 0.05,
the LCO response is plotted in Fig. 5. In Fig. 5(a), the result is obtained by the averaging
10 Hu-lun GUO and Yu-shu CHEN

Fig. 4 LCO amplitude in pitch vs. Mach number for δ = 0

Fig. 5 LCO amplitude in pitch vs. Mach number for Mf = 0.1, δ = 0.1, and αf = 0.05

method, and in Fig. 5(b), the result is obtained by a numerical method. In Fig. 5(a), the solid
curve depicts the stable LCO, and the dashed curve depicts the unstable LCO. Comparing
Fig. 4 with Fig. 5, we see that the Mach number of the supercritical Hopf bifurcation point
(P1 ) does not change. It means that the freeplay is independent of the flutter Mach number.
One supercritical Hopf bifurcation (P1 ) and two fold bifurcations (P2 and P3 ) exist in this
system with freeplay. An amplitude jump phenomenon followed by the fold bifurcation is
detected. The LCO amplitude jumps from P2 and P3 , where the LCO amplitude increases
abruptly. This situation is dangerous and even catastrophic for an airfoil. Dimitriadis[15] found
the amplitude jump phenomenon and the fold bifurcation in a two-DOF airfoil system with the
cubic nonlinearity. An LCO with a large amplitude that appears before the supercritical Hopf
bifurcation may abruptly jump from the zero point. In the previous studies[2,26] , the LCO can
appear before the flutter speed when the Hopf bifurcation is subcritical. One can conclude that
the LCO may also appear before the flutter speed when the Hopf bifurcation is supercritical
for an airfoil system with freeplay. Therefore, it is necessary to avoid the control surfaces with
loose joints. The numerical result shown in Fig. 5(b) depicts the supercritical Hopf bifurcation,
the fold bifurcation, and the jump phenomenon. It is consistent with the result obtained by
the averaging method.
When Mf = 0.1, δ = 0.1, and αf = 0.1, the LCO response is shown in Fig. 6, where Fig. 6(a)
Dynamic analysis of two-degree-of-freedom airfoil 11

is obtained by the averaging method, and Fig. 6(b) is obtained by a numerical method. In
Fig. 6(a), the solid curve depicts the stable LCO, and the dashed curve depicts the unstable
LCO. One supercritical Hopf bifurcation (P1 ), two fold bifurcations (P2 and P3 ), and the
amplitude jump phenomenon can also be found. In this case, the jump phenomenon happens
between two LCOs as the Mach number increases past the fold bifurcation point. The numerical
result shown in Fig. 6(b) depicts the supercritical Hopf bifurcation, the fold bifurcation, and
the jump phenomenon. It is noted that it agrees with that in Fig. 6(a) very well.

Fig. 6 LCO amplitude in pitch vs. Mach number for Mf = 0.1, δ = 0.1, and αf = 0.1

To investigate the effect of the freeplay parameters on the LCO response, the LCO ampli-
tudes in the pitch DOF with different freeplay parameters are plotted in Figs. 7–9.
In Fig. 7, the freeplay parameters are αf = 0.1, δ = 0.1, and Mf = 0.0, 0.5, 1.0, respectively.
The solid line depicts the LCO response of the system without freeplay, i.e., Mf = 1.0. In
this case, the LCO amplitude increases with the increase of the Mach number, and the jump
phenomenon of the LCO amplitude does not exist. The dotted line depicts the LCO response of
the system when Mf is 0.5. There is a turning point (TP) at the start of the freeplay position,
i.e., a2 = αf (αf = 0.1). The LCO amplitude with freeplay becomes bigger than the ampli-
tude without freeplay after the TP. It should be remarked that the jump phenomenon of the
LCO amplitude still does not exist in this case. As the central stiffness term for the freeplay Mf

Fig. 7 LCO amplitude in pitch vs. Mach Fig. 8 LCO amplitude in pitch vs. Mach
number for δ = 0.1 and αf = 0.1 number for Mf = 0.1 and δ = 0.1
12 Hu-lun GUO and Yu-shu CHEN

decreases, i.e., Mf = 0.0, the dashed line depicts the LCO response. In this case, there is the
fold bifurcation at the TP. Moreover, a jump phenomenon of the LCO amplitude appears as
the Mach number is above the fold bifurcation point. The jump phenomenon leads to an LCO
with a large amplitude.
In Fig. 8, the freeplay parameters are Mf = 0.1, δ = 0.1, and αf = 0.05, 0.10, 0.15, respec-
tively. The TP is changed with the start of the freeplay αf . The solid line depicts the LCO
response of the system with αf = 0.15. In this case, the jump phenomenon does not exist. For
αf = 0.10, a jump phenomenon appears, which is depicted by the dotted line. The LCO with
a small amplitude jumps to another larger one. For αf = 0.05, the jump phenomenon also
appears, which is depicted by the dashed line. A more dangerous jump phenomenon appears.
An LCO with a large amplitude can jump from the zero solution before the Hopf bifurcation
point. It means that the LCO can appear before the flutter Mach number. Furthermore, the
LCO amplitude increases after the TP as the start of the freeplay decreases.
In Fig. 9, the freeplay parameters are αf = 0.1, Mf = 0.10, δ = 0.05, 0.10, 0.15, respectively.
The solid line depicts the LCO response of the system with δ = 0.15. The dotted line and the
dashed line depict the LCO responses of the system with δ = 0.10 and δ = 0.05, respectively.
It can be seen that the jump phenomenon always exists in such three cases. However, the LCO
amplitude after jumping decreases. It is indicated that the freeplay angle cannot change the
property of the TP but can enlarge the jumping amplitude.
Therefore, the results obtained from the averaging method give an explicit and sufficiently
accurate prediction for the bifurcations and the LCO response of an airfoil system with freeplay
and cubic nonlinearities. Both the stable and unstable LCOs are obtained.

Fig. 9 LCO amplitude in pitch vs. Mach number for Mf = 0.1 and αf = 0.1

6 Conclusions

In this paper, the nonlinear aeroelastic response of a two-DOF airfoil with freeplay and cubic
nonlinearities in supersonic flows is investigated. The averaging method and the multi-variable
Floquet theory are used to analyze the LCOs of the airfoil. One supercritical Hopf bifurcation,
two fold bifurcations, and the amplitude jump phenomenon are detected in this airfoil system.
The analytical results are verified by the numerical method. Finally, the influence of the freeplay
parameters on the LCO response is analyzed in detail. The results show that the freeplay can
cause the amplitude jump phenomenon, which is dangerous and even catastrophic for an airfoil.
The jump phenomenon may disappear as the central stiffness term for the freeplay or the start
of the freeplay increases. However, the freeplay angle cannot change the property of the jump
Dynamic analysis of two-degree-of-freedom airfoil 13

phenomenon. Moreover, decreasing the freeplay angle can only decrease the jumping value of
the amplitude.

References
[1] Abbas, L. K., Qian, C., Marzocca, P., Zafer, G., and Mostafa, A. Active aerothermoelastic control
of hypersonic double-wedge lifting surface. Chinese Journal of Aeronautics, 21(1), 8–18 (2008)
[2] Librescu, L., Chiocchia, G., and Marzocca, P. Implications of cubic physical/aerodynamic nonlin-
earities on the character of the flutter instability boundary. International Journal of Non-Linear
Mechanics, 38(2), 173–199 (2003)
[3] Hyun, D. H. and Lee, I. Transonic and low-supersonic aeroelastic analysis of a two-degree-of-
freedom airfoil with a freeplay non-linearity. Journal of Sound and Vibration, 234(5), 859–880
(2000)
[4] Liu, L. and Song, Y. S. Nonlinear aeroelastic analysis using the point transformation method,
part 1, freeplay model. Journal of Sound and Vibration, 253(2), 447–469 (2002)
[5] Roberts, I., Jones, D. P., Lieven, N. A. J., Bernado, M. D., and Champneys, A. R. Analysis of
piecewise linear aeroelastic systems using numerical continuation. Journal of Aeronautical Engi-
neering, 216(1), 1–11 (2002)
[6] Chen, Y. M. and Liu, J. K. Supercritical as well as subcritical Hopf bifurcation in nonlinear
flutter systems. Applied Mathematics and Mechanics (English Edition), 29(2), 181–187 (2008)
DOI 10.1007/s10483-008-0207-x
[7] Liu, L., Wong, Y. S., and Lee, B. H. K. Application of the center manifold theory in nonlinear
aeroelasticity. Journal of Sound and Vibration, 234(4), 641–659 (2000)
[8] Chung, K. W., Chan, C. L., and Lee, B. H. K. Bifurcation analysis of a two-degree-of-freedom
aeroelastic system with freeplay structural nonlinearity by a perturbation-incremental method.
Journal of Sound and Vibration, 299(3), 520–539 (2007)
[9] Chen, Y. M. and Liu, J. K. Homotopy analysis method for limit cycle oscillations of an airfoil
with cubic nonlinearities. Journal of Vibration and Control, 16(2), 163–179 (2010)
[10] Raghothama, A. and Narayanan, S. Nonlinear dynamics of a two-dimensional airfoil by incremen-
tal harmonic balance method. Journal of Sound and Vibration, 226(3), 493–517 (1999)
[11] Gordon, J. T., Meyer, E. E., and Minogue, R. L. Nonlinear stability analysis of control surface
flutter with freeplay effects. Journal of Aircraft, 45(6), 1904–1916 (2008)
[12] Shen, S. F. An approximate analysis of nonlinear flutter problems. Journal of the Aerospace
Sciences, 25(1), 25–32 (1959)
[13] Yang, Y. R. KBM method of analyzing limit cycle flutter of a wing with an external store and
comparison with a wind-tunnel test. Journal of Sound and Vibration, 187(2), 271–280 (1995)
[14] Kim, S. H. and Lee, I. Aeroelastic analysis of a flexible airfoil with a freeplay nonlinearity. Journal
of Sound and Vibration, 193(4), 823–846 (1996)
[15] Dimitriadis, G. Bifurcation analysis of aircraft with structural nonlinearity and freeplay using
numerical continuation. Journal of Aircraft, 45(3), 893–905 (2008)
[16] Zhao, D. M. and Zhang, Q. C. Bifurcation and chaos analysis for aeroelastic airfoil with freeplay
structural nonlinearity in pitch. Chinese Physics B, 19(3), 1–10 (2010)
[17] Conner, M. D., Tang, D. M., Dowell, E. H., and Virgin, L. N. Nonlinear behavior of a typical
airfoil section with control surface freeplay: a numerical and experimental study. Journal of Fluids
and Structures, 11(1), 89–109 (1997)
[18] Tang, D. and Dowell, E. H. Flutter and limit-cycle oscillations for a wing-store model with freeplay.
Journal of Aircraft, 43(2), 487–503 (2006)
[19] Tang, D., Dowell, E. H., and Virgin, L. N. Limit cycle behavior of an airfoil with a control surface.
Journal of Fluids and Structures, 12(7), 839–858 (1998)
14 Hu-lun GUO and Yu-shu CHEN

[20] Tang, D., Conner, M. D., and Dowell, E. H. Reduced-order aerodynamic model and its application
to a nonlinear aeroelastic system. Journal of Aircraft, 35(2), 332–338 (1998)
[21] Liu, L. and Dowell, E. H. Harmonic balance approach for an airfoil with a freeplay control surface.
AIAA Journal, 43(4), 802–815 (2005)
[22] Lin, W. B. and Cheng, W. H. Nonlinear flutter of loaded lifting surfaces (I) and (II). Journal of
the Chinese Society of Mechanical Engineers, 14(5), 446–466 (1993)
[23] Ashley, H. and Zartarian, G. Piston theory — a new aerodynamic tool for the aeroelastician.
Journal of the Aeronautical Sciences, 23(12), 1109–1118 (1956)
[24] Abbas, L. K., Chen, Q., O’Donnell, K., Valentine, D., and Marzocca, P. Numerical studies of
a nonlinear aeroelastic system with plunging and pitching freeplays in supersonic/hypersonic
regimes. Aerospace Science and Technology, 11(5), 405–418 (2007)
[25] Friedmann, P. P., McNamara, J. J., Thuruthimattam, B. J., and Nydick, I. Aeroelastic analysis
of hypersonic vehicles. Journal of Fluids and Structures, 19(5), 681–712 (2004)
[26] Lee, B. H. K., Price, S. J., and Wong, Y. S. Nonlinear aeroelastic analysis of airfoils: bifurcation
and chaos. Progress in Aerospace Sciences, 35(3), 205–334 (1999)
[27] Nayfeh, A. H. and Mook, D. T. Nonlinear Oscillations, 1st ed., Wiley, New York (1979)
[28] Nayfeh, A. H. Perturbation Methods, 1st ed., Wiley, New York (1973)
[29] Friedmann, P., Hammond, C. E., and Woo, T. H. Efficient numerical treatment of periodic systems
with application to stability problems. International Journal of Numerical Methods in Engineering,
11(7), 1117–1136 (1977)
[30] Ge, Z. M. and Chen, H. H. Bifurcations and chaotic motions in a rate gyro with a sinusoidal
velocity about the spin axis. Journal of Sound and Vibration, 200(2), 121–137 (1997)

You might also like