Notes Aeroelasticity
Notes Aeroelasticity
Aeroelasticity
Notes from prof. Quaranta’s lessons
This book has been written using our notes taken in class.
This is not meant to substitute the following classes or the book Professor Quaranta
suggested.
2 Static Aerolasticity 11
2.1 Rigid wing model: typical section . . . . . . . . . . . . . . . . . . . . . . . 11
2.1.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2 Divergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.1 Equilibrium approach . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.2 Comparison of aeroelastic and rigid lift . . . . . . . . . . . . . . . 13
2.2.3 Torsional divergence as a stability problem . . . . . . . . . . . . . 14
2.2.4 Eigenvalue approach problem . . . . . . . . . . . . . . . . . . . . . 14
2.2.5 Divergence and Mach number . . . . . . . . . . . . . . . . . . . . . 15
2.3 Control surface effectiveness . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3.1 Control Elastic efficiency and control reversal . . . . . . . . . . . . 16
2.3.2 Plot control surface elastic efficiency . . . . . . . . . . . . . . . . . 16
2.3.3 Control Surface Load Distribution . . . . . . . . . . . . . . . . . . 17
2.3.4 Matrix form of control reversal problem . . . . . . . . . . . . . . . 17
2.4 Control surface stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.5 Rolling of a straight wing . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.5.1 Aerodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.5.2 Twist . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.5.3 Control reversal problem . . . . . . . . . . . . . . . . . . . . . . . . 22
2.5.4 Consistent performance problems . . . . . . . . . . . . . . . . . . . 22
6 Structural dynamics 52
6.0.1 Continuous deformable structure: hypothesis . . . . . . . . . . . . 52
6.1 Introduction and beam modeling . . . . . . . . . . . . . . . . . . . . . . . 52
6.2 Euler-Bernoulli’s beam model . . . . . . . . . . . . . . . . . . . . . . . . . 52
6.2.1 Axial load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
6.2.2 Bending . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
6.2.3 Torsion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
6.3 Timoshenko’s beam model . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.4 Free vibration for the beam bending . . . . . . . . . . . . . . . . . . . . . 54
6.5 Free vibration for the beam torsion . . . . . . . . . . . . . . . . . . . . . . 55
6.6 Properties of the modal forms . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.6.1 Orthogonality with respect to mass distribution . . . . . . . . . . . 56
6.6.2 Orthogonality with respect to stiffness distribution . . . . . . . . . 56
6.6.3 Inner product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.7 Modal decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
7 Modal Analysis 60
7.1 State-space form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
7.2 Normalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
7.3 Decoupling through eigenmodes . . . . . . . . . . . . . . . . . . . . . . . . 61
7.4 State space form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
7.4.1 Decoupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
7.5 Coincident eigenvalues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
7.6 Rigid body modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
7.7 Gram-Schmidt orthogonalization . . . . . . . . . . . . . . . . . . . . . . . 63
7.8 Computation of the solution to initial conditions . . . . . . . . . . . . . . 63
7.9 Rayleigh quotient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
7.9.1 Error on the eigenvalue . . . . . . . . . . . . . . . . . . . . . . . . 65
7.10 Computations of a group of eigenvectors & eigenvalues . . . . . . . . . . . 65
7.10.1 Bloch-Stodola block iteration . . . . . . . . . . . . . . . . . . . . . 66
8 Damping models 67
8.1 Kevin-Voigd model (Parallel) . . . . . . . . . . . . . . . . . . . . . . . . . 67
8.2 Maxwell model (Series) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
8.3 Equivalent Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
8.4 Energy analysis - meaning of equivalent damping ratio . . . . . . . . . . . 69
8.4.1 Energy dissipation . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
8.5 Modal damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
8.6 Rayleigh damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
8.7 Equivalent damping: Coulomb friction . . . . . . . . . . . . . . . . . . . . 71
8.8 Hysteretic damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
9 Dynamic Response 73
9.1 Laplace transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
9.1.1 State-space Format . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
9.2 Frequency response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
9.2.1 Dynamic stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
9.2.2 One DoF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
9.2.3 Modal computation of the response . . . . . . . . . . . . . . . . . . 76
9.2.4 Contribution of different terms . . . . . . . . . . . . . . . . . . . . 77
9.3 Fourier Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
9.3.1 Non-periodic but periodic-sizable signals . . . . . . . . . . . . . . . 78
9.4 Fourier Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
9.5 Limited band excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
9.6 Inertia Relief . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
10 Random Vibrations 88
10.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
10.2 Statistic quantities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
10.2.1 Probability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
10.2.2 Probability distribution . . . . . . . . . . . . . . . . . . . . . . . . 88
10.2.3 Cumulative probability function (CPF) . . . . . . . . . . . . . . . 89
10.3 Probability density function . . . . . . . . . . . . . . . . . . . . . . . . . . 89
10.3.1 Expected values . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
10.3.2 Variance and standard deviation . . . . . . . . . . . . . . . . . . . 89
10.3.3 Gaussian process . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
10.3.4 Gaussian process: joint distributions . . . . . . . . . . . . . . . . . 89
10.4 Process function of time . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
10.4.1 Auto-correlation and cross-correlation . . . . . . . . . . . . . . . . 90
10.4.2 Stationary process . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
10.4.3 Time shift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
10.4.4 Properties of the correlation . . . . . . . . . . . . . . . . . . . . . . 91
10.5 Ergodic process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
10.5.1 Auto-covariance and Cross-covariance . . . . . . . . . . . . . . . . 92
10.6 Auto-covariance of a difference . . . . . . . . . . . . . . . . . . . . . . . . 92
10.6.1 Auto-covariance as a mean to de-noise a signal . . . . . . . . . . . 93
10.7 Power Spectral Density (PSD) . . . . . . . . . . . . . . . . . . . . . . . . 93
10.7.1 Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
10.8 White noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
10.8.1 Colored Noise (Band-limited white noise) . . . . . . . . . . . . . . 94
10.9 Exponential correlation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
11 Random response 95
11.1 Computation of the mean . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
11.1.1 Computation of the mean knowing the mean of the input . . . . . 95
11.1.2 Computation of the mean with the second order formulation . . . 95
11.1.3 State space formulation . . . . . . . . . . . . . . . . . . . . . . . . 96
11.2 Computation of the variance of the output using the impulse response . . 96
11.3 Computation of the response in frequency domain . . . . . . . . . . . . . 96
11.4 White noise approximation . . . . . . . . . . . . . . . . . . . . . . . . . . 97
Aerospace structures differ from other structures due to the combined effect of the re-
quest to be lightweight and at the same time extremely performing. This, as a conse-
quence, leads to thin-walled structures, often based on the usage of composite or other
multi-functional materials, characterized by large flexibility.
Considering also controls and thermal effects we have the so-called ”Aero-Servo-Thermo-
Elasticity”.
Normally there is minimal effect of these two modes on structural behavior, with only a
slight vibration being seen for each motion. The bending mode shows up as a relatively
low-frequency flapping effect while the twisting mode is found to be a much higher
frequency vibration. However, with the application of high-speed airflow as a source of
excitation energy, these two modes can produce motions with will severely distort or
break the wing.
1.3.1 Divergence
In the divergence case, the moment produced by the air load is greater than the structural
torsional stiffness and thus it will be twisted off the vehicle. The threshold speed for this
type of failure to occur is called divergence speed, particular problems occur with swept
forward wings as these have a relatively low divergence speed.
1.3.2 Flutter
In this case, there is a synchronized interaction between both modes so that energy
is absorbed from the airflow in one mode to increase the amplitude of the other. At
this point, the frequency of each mode has converged to the same value so that only
one combined mode is possible. The wing will absorb energy from the airflow and will
behave as an ever-increasing bending and torsion flexure until sufficient displacement is
reached and the wing breaks. When the airflow is increased to the critical point to cause
this failure, it is called the flutter speed.
Still today Control surface flutter is the most common problem found in design. i.e.
right and left stabilizers are connected to avoid flutter.
Monoplanes were made of wood and fabric: with very low torsional stiffness. i.e.
Fokker D-VIII had wing failures at high speeds.
• 1916-now: Control surface flutter. i.e. H.P. Type 0 had a tail with twist moment:
oscillatory divergence. This was solved by joining the two surfaces.
• 1940-now: Control reversal. The presence of a flexible wing and flexible aileron
means that there is a certain speed at which the rotation of the control surfaces
produces no effects. This is a very dangerous situation since the pilot can’t control
the plane anymore.
• 2D strip theory;
• linear behavior since only low amplitude pitch oscillations are considered
2.1.1 Definitions
Shear center (SC)
It is a point of a 2D section where if you apply a transverse shear load it produces zero
change (i.e. rate) of twist of the section. It is a property of the cross-section
2.2 Divergence
kα kα
qD = =
eSCLα kA
θ 1
=
θ0 1 − qqD
It is possible to reach a dynamic pressure such that the numerator goes to zero: it means
that the rotation of the flap isn’t producing any change of lift. The rotation of the flap
produces an increase of lift CLβ and consequently negative pitching, the wing submitted
to CMβ twist with θ < 0, this causes a reduction of the angle of attack and consequently
a reduction of lift that (in this specific case), balances the increase of lift due to the flap.
Control reversal condition:
kα CLβ
qR = −
ScCLα CMβ
In the control reversal condition, the flap deflection produces an effect that is opposite
to the one for which the flap was designed. The more I increase q the less efficient is the
control. If you want an aircraft that is more controllable at a higher speeds, you must
meet divergence earlier.
The elastic twist reduces the incidence and torsional loads, favorable redistribution of
loads.
Find qR by imposing:
det(A) = 0
This is again an eigenvalues problem that gives us an additional solution at zero
dynamic pressure any rotation of beta produces no change of lift.
We are adding another equation. Our system is now in equilibrium around EA and flap
hinge.
kβ
θ = qS (eCLβ + cCmβ )β0
detA
kβ
β= (kα + qSeCLα )β0
detA
Substituting in the lift L = qS(CLα θ + CLβ β):
L = qS (CLα θ + CLβ β)
kβ
L = qS (CLα qS (eCLβ + cCmβ ) + CLβ kα − CLβ qSeCLα ) β0
det A
kβ kα qSc
L = qS CLβ + Cmβ CLα β0
det A kα
Reversal is possible for the same dynamic pressure as for the infinite flap hinge
stiffness. There is a change in divergence speed.
We allow the simple typical section model to rotate freely about the roll axis. The
proposed model allows the dynamics of the rigid roll movement caused by a rotation of
the aileron to be investigated
Hypothesis:
1. The motion is considered the sum of a rigid movement plus the deformation of the
wing structure.
2.5.1 Aerodynamics
py py
∆α = tan ≈
U∞ U∞
The lift is:
py
L = qc(y) CLα θ − + CLβ β
U∞
MAC = qc2 (y)Cmβ β
We have also inertia forces in G:
Fi = −mṗy
The twist generated around the EA is:
Mi = mdṗy
Integrating:
Z b Z b
Ly dy − my 2 dy ṗ = 0
0 0
Z b
py
−Ixx ṗ + qc CLα θ − + CLβ β dy = 0
0 U∞
2.5.2 Twist
Z b Z b
kα θ = (Le + MAC dy + myd dy ṗ
0 0
Substituting:
mb2
1 pb
kα θ = qS eCLα θ − + eCLβ + Cmβ β + dṗ
2 U∞ 2
Or also:
b2
eCLα pb
(kα − qSeCLα )θ = qS − + eCLβ + Cmβ β + mT dṗ
2 U∞ 2
We have a system of two equations:
py
Ixx ṗ = qSb Clθ θ + Clp U + Clβ β (1)
∞
b2
eC
(kα − qSeCLα )θ = qS − Lα f racpbU∞ + eCLβ + Cmβ β + mT dṗ (2)
2 2
Matrix form of the problem:
" #
C
Ixx −qSb L2 α ṗ bCℓβ bCℓp pb
= qS β − qS eCLα
−mT db
2 (k α − qkA ) θ eC Lβ
+ cCm β 2 U∞
If d = 0, then:
(eCLβ + cCmβ ) eCLα pb
θ = qS β − qS
kα − qkA 2(kα − qkA ) U∞
Otherwise:
(eCLβ + cCmβ ) eCLα pb mT db
θ = qS β − qS + ṗ
kα − qkA 2(kα − qkA ) U∞ 2(kα − qkA )
Or
pb
Ixx ṗ = qSb −(Clp )e + (Clβ )e β
U∞
Kα CLβ
qR = −
ScCLα Cmβ
p̄b (Clβ )e
= β
U∞ (Clp )e
qSb(Clβ )e
ṗ0 = β
Ixx
The model is composed of several rigid wing portions connected through torsional
springs. For each rigid portion the simple pseudo-2D aerodynamic model is used. No
mutual aerodynamic influence between portions is considered.
Where δri and φj are virtual movements: do not violate any constraints.
Another way to write the PWC is:
δWi = δWe
∂U
= Qi
∂θi
(Ks − qKA )θ = M0
(Ks − qKA )θ = 0
det(Ks − qKA ) = 0
Note: in general, we will have as many divergences dynamic pressures as many degrees
of freedom.
For our purposes we are actually interested in the lowest one (the most restricting dy-
namic pressure).
1
We can simply define λ̂ = λ and look for the biggest λ̂.
Let’s compute them using the PCVW (Principle of Complementary Virtual Works)
through a virtual load δMt applied in yi+1 :
Z b Z yi+1
∗ Mt Mt Mt δMt
δ Wi = δMt dξ = δMt dξ = yi+1
0 GJ 0 GJ GJ
δ ∗ We = θi+1 δMt
Mt δMt
yi+1 = θi+1 δMt
GJ
Mt
δMt θi+1 − yi+1 = 0 ∀δMt (1)
GJ
And a virtual load δ M̂t applied in yi :
Mt
δ M̂t θi − yi = 0 ∀δ M̂t (2)
GJ
Go on until:
n n
1 X X
θ̂1 = mj = C1 mj
k1
j=1 j=1
Note: if a unit point moment is applied as y = γ i.e m(ξ) = δ(ξ − γ) where δ is the Dirac
delta we have: Z L
θ(y) = C θθ (y, ξ)δ(ξ − γ) dξ = C θθ (y, γ)
0
So C θθ represents the twist at y due to a unit moment in γ. How do we proceed?
C θθ (yi , yj ) = C θθ (yj , yi )
(cCL )e
But changing the unknowns: θ =
cCLα
Substituting:
Z L
(cCL )e
=q C θθ (y, ξ)(ecCL0 + e(cCL )e + c2 CmAC ) dξ
cCLα 0
(cCL )e is taken as the unknown and discretized as constant values on wing strips, making
the hypothesis that the deformation does not depend on how loads are distributed on a
strip.
And
Cij = C θθ (yj , yi )
The integral is:
C θθ (yj , yi )∆bi
Where ∆bi = yi+1/2 − yi−1/2 . The sum can be also written as:
C11 C12 . . . ∆b1 0 0
C21 C22 . . . 0 ∆b 0
F= 2
.. .. .. ..
. . . 0 0 .
Note: F = K−1
Mt − (Mt + dMt ) = mt dy
dMt
+ mt = 0
dy
Then the constitutive law for torsion can be expressed as GJθ′ = Mt .
Substituting:
(GJθ′ )′ + mt = 0
Apply BCs: θ(0) = 0 and θ′ (L) = 0.
We can also get the same result by using the PVW. Let’s write the PVW for torsion:
Z Z L Z L
′
δWi = σij δεij dv = δθ Mt dy = δθ′ GJθ′ dy
V 0 0
Integrate by parts:
Z L Z L
δWi = [δθGJθ′ ]L
0 − δθ(GJθ ) dy ′ ′
δWe = δθmt dy
0 0
Write the differential equation for the wing case (all properties uniform):
Deriving:
qceCLα qc dN mg
θ′′ + =− (e(CL0 + cCmAC ) +
GJ GJ GJ
Let’s non-dimensionalize the problem by defining:
y d2 θ d2 θ d2 ŷ θ∗∗
ŷ = −→ θ′′ = 2 = 2 2 = 2
L dy dŷ dy L
Our equation then becomes:
Where:
qecCLα L2 qcL2 dN mgL2
µ2 = QA = − (eCL0 + cCmAC )) Qm =
GJ GJ GJ
µ cos µ = 0
This happens for µ = 0 and µ = ±(2n + 1) π2 , n ∈ N. From here we can retrieve the n
divergence pressure as:
π 2 GJ
qD = (2n + 1)2
2 ecL2 CLα
We have an infinite number of eigenvalues or divergence dynamic pressures.
From the same result, we can also find the associated deformation modes or eigenfunc-
tions: πy
θ̂n = sin (2n + 1)
2L
If we neglect the inertial effects (i.e. d = 0) and consider a symmetric airfoil (CmAC = 0),
then:
qcL2 eCL0 QT CL
QT = QA = − =⇒ 2 = − 0
GJ µ CLα
Due to the twist of the wing, we will have a higher angle of attack at the tip of the wing.
This means that we will have an overall higher lift than the one expected for the rigid
wing.
This will also change the shape of our distribution. i.e. if we wanted an elliptical lift
distribution for better performances the twist would change this behaviour.
Where:
N
X
θ(y) ≈ θ̃(y) = Ni (y)qi = N(y)q
i=1
Where:
N = [N1 , N2 . . . , NN ] q = [q1 , q2 . . . , qN ]T
Note: If the method is applied to a strong formulation, (i.e. differential) the chosen
shape functions must satisfy all boundary conditions.
Each shape function should be p times differentiable where p is the highest spatial
derivative present in the formulation. At least one of the shape function p-th order
derivatives should be not zero.
To get a solvable system with N equations we can adopt two different strategies:
1. Weighted integrals: Z L
wi (y)ε(y) dy = 0
0
Where wi are weight functions.
2. Collocation (spectral method): imposing null the error in N points:
ε(Yi ) = 0 i = 1, 2, 3, . . . , m
Usually, the Ritz approach is applied to the weak formulation. In our case the PVW.
In this case there are two functionals to be approximated: the displacement field (trial
functions), and the virtual displacement field (test function). If the same series (func-
tional space) is used for both, the approach is called Ritz-Galerkin.
• The boundary conditions on forces and moments, called “natural” (Neumann BCs)
are not necessary. Their satisfaction will be obtained. naturally at convergence
(by using a sufficiently large number of shape functions). However, satisfying also
these conditions can speed up convergence.
4.1 Introduction
4.1.1 Reason to use a sweep angle
There are many reasons to use swept wings in an aircraft: firstly, we want to reduce the
local Mach. Secondly, we want to increase stability for subsonic flight. We know that
in the transonic range, we have a higher CD due to compressibility effects. By sweeping
the wing the velocity seen by the wing will be reduced by a factor of cos Λ. This means
we can go faster and still have a lower local Mach on the wing, preventing in this way
the added drag.
Let’s look at the image below (fig 17) (source: Hoerner’s Fluid Dynamic Drag, section
XV).
4.1.2 Torsion
Aerodynamic forces are developed by airfoils aligned with the asymptotic speed.
Define a structural reference system frame (x̄, ȳ) aligned with the axis of the wing.
Rotated by a Λ angle wrt the aerodynamic reference (x, y). Torsion and bending are
along the structural reference system. We can write the rotation matrix as:
x cos Λ − sin Λ x̂
=
y sin Λ cos Λ ŷ
Because of this only a portion of the torsion θ results in a change in angle of attack:
∆αT = θ̄ cos Λ
φ = θ̄ sin Λ
This effect can be neglected since is small.
4.1.3 Bending
When a swept-back wing bends up them angle of attack of streamwise sections will
reduce.
∆ȳ = c sin Λ
z(ȳ1 ) − z(ȳ2 ) 1 dz
∆αB = =− ∆ȳ = −z ′ sin Λ
c c dy
α = α0 + ∆α = α0 + θ̄ cos Λ − z ′ sin Λ
S(y) = qcCL − mN g
The wing bending is caused by the shear force load S(ȳ) and the moment mx̄ . This
second term may be neglected (as a first approximation).
The lift L works for the virtual displacement of the AC: δwAC = δ(θ̄e + φ̄y).
Z b̄ Z b̄
δWe = δ θ̄T eL d(ȳ) + δ φ̄T eȳL d(ȳ)
0 0
Where:
1 α
0
α0 + θ̄ cos Λ − z ′ sin Λ = q̄c̄C̄Lα + θ̄ − z ′ tan Λ
L = qcCLα α = q̄c̄C̄Lα
cos Λ cos Λ
Note:
c̄ α CLα
dy = dȳ cos Λ , c = , q̄ = q cos2 Λ , ᾱ = , C̄Lα =
cos Λ cos Λ cos Λ
We obtain:
θ̄ Q̄ ē
(Ks − Q̄KA ) = α0
φ̄ cos λ b̄/2
Where:
k 0 ē − tan Λē
Ks = θ̄ KA =
0 kφ̄ b/2 − tan Λb̄/2
And Q̄ = q̄c̄b̄C̄Lα = q̄S C̄Lα = qS cos2 ΛC̄Lα
It depends on Λ:
1. Λ > 0 (backward sweep angle). It is possible to identify a critical sweep angle
ΛCRIT for which no divergence exists, because qD < 0.
2. Λ < 0 (forward sweep angle). The higher is |Λ| the lower is qD
for example below −30° divergence speed is close to zero.
When Λ > ΛCRIT we have a loss in lift (we have to go faster) BUT it is more stable.
When Λ < 0 the ratio grows very fast: we have a lot of lift at higher speed BUT it
is less stable. For large positive sweep angles, the divergence disappears but there is a
reduction of the lift effectiveness.
CL
L = q̄S C̄Lα θ̄ − φ̄ tan Λ + q̄sCLβ β = Q̄(θ̄ − φ̄ tan Λ) + Q̄ β β
C̄Lα
Cm β
MAC = q̄Sc̄Cmβ = Q̄c̄ β
C̄Lα
Cm
CL 1 + ēc̄ CL β
θ̄
(Ks − Q̄KA ) = Q̄ē β β β
φ̄ C̄Lα b̄
2ē
S C̄Lα c̄ Cmβ
L 1+ kθ̄ C Lβ cos2 Λ
= q̄
LR 1− q¯D
Imposing the numerator null will give us the Control reversal speed:
kθ̄ CLβ 1
qR = −
S C̄Lα c̄ Cmβ cos2 Λ
We can use again the PVW to get the equilibrium equation in the coupled case.
Find then the divergence pressure and impose the denominator null: the critical condi-
tion corresponds to QD = ∞.
−1 kφ̄ ē − k b̄/2
ΛCRIT = tan
b̄/2kθ̄ − kē
In this way, for a negative swept wing (i.e. forward swept), it is possible to obtain
a negative critical sweep angle if k > 0 and large enough in modulus. It means to
implement a washout effect on the forward swept wing.
More generally using composite material with appropriate direction for the deposition
of the fibers, it is possible to obtain the level of coupling required.
4.6 Summary
Positive sweep:
• Decreases controllability.
Let’s introduce a new parameter: time. Considering the typical section we have two
d.o.f: pitch θ(t) and plunge h(t) both measured at the elastic axis.
The equation of motion is:
m md ḧ kh 0 h Qh
2 + =
md md + Iθ θ̈ 0 kθ θ Qθ
Where:
Qh = −L
Qθ = eL + MAC
Or briefly:
Mq̈ + Kq̇ = Q
Re(λ) = 0
We have purely imaginary eigenvalues:
The system is stable but not asymptotically.
So:
m Sθ ḧ kh 0 0 −1 h
+ + qSCLα =0
S θ Iθ θ̈ 0 kθ 0 e θ
λi = σ + jω
From which:
ξ > 0 Asympthotically stable
ξ = 0 Stable
ξ < 0 Unstable
Re(λ)
g=
Im(λ)
But if
ξ << 1 =⇒ g ≈ −ξ
We know that each peak is going down following a damped exponential term:
x(t + T ) = x(t)e−ξωn T
x(t)
δ = ln
x(t + T )
λi = ±j − ωi ∈ C
To identify it, we have to find the dynamic pressure at which the quantity (b2 − 4ac)
gets negative.
Flutter: find qF s.t. (b2 − 4ac) < 0, where qF is the flutter dynamic pressure.
c1 qF2 + c2 qF + c3 = 0
5.2.4 Divergence
If we increment further q we get to a point where c = 0
kθ − qSeCLα = 0 =⇒ q = qD
Under the flutter condition (V ≈ 1.18) the phase shift is 0. Where are moving on a
straight line: the work done is null. We are keeping all our energy. Above the flutter
condition we have a phase shift that grows up to 180°. This means that the airflow
pumps energy into the structure that necessarily amplifies its oscillations. The air is
injecting energy into the system. We have to slow down.
ḣ
α=θ+
U
Positive in the upward direction increases the angle of attack.
m Sθ ḧ qSCLα 1 0 ḣ kh 0 0 −1 h
+ + + qSCLα =0
Sθ Iθ θ̈ U −e 0 θ̇ 0 kθ 0 e θ
We have a model that says that the flutter occurs at 1.18 and the other at 0.47. Which
one is correct?
Let’s now look at an actual flutter model, confirmed experimentally. The graph is from
Dowell’s book. (Fig:32). The actual flutter is at 0.86.
Another interesting phenomenon that we can observe in this graph is how the damping ξ
(lower part of the graph) tends to become more negative as the velocity increases. This
is dangerous behavior since it looks like it will get more stable. Suddenly one of the two
ξ increases and becomes positive. This is known as hard flutter.
This leads to a butterfly speed distribution along the chord that has an analogous effect
of a cambered profile.
We generally do not consider this effect alone. We will study the complete unsteady
model in chapter 13.
Note: to not consider a static model we need to have at least k < 0.001.
k = 0.001 −→ η ≈ 3141.5c
• Small displacement
• Geometric and structural properties are assumed to change way more regularly
span-wise than section-wise. This means that abrupt section changes or abrupt
bends are not allowed.
• From the forces point of view, the beam is represented by the resultant stresses
acting on its section.
• From the kinematic point of view, the beam is described by the behaviour of a
reference line; generalized stain measures must be defined.
Thanks to assumption 1 the displacements in the plane of the section are rigid, i.e. not
function of the position on the section
Thanks to assumption 3:
dw
φ2 = − = −w′
dx1
dv
φ3 = = v′
dx1
6.2.2 Bending
∗ ′′ ′′ ′
(EI2 w ) + mẅ = q3 + m2
4BC
2IC
∗ ′′ ′′ ′
(EI3 v ) + mv̈ = q2 + m3
4BC
2IC
Note: This is valid only for the principal reference system, placed in the centroid of the
section. In general we have three coupled equations.
′
N EA V3 −V2 u
M2 V3 EI2 −EI23 −w′′
M3 −V2 −EI23 EI3 v ′′
6.2.3 Torsion
∗ ′ ′
(GJ θ ) − I θ̈ = −qt
2BC
2IC
The simplest model to recover an “average” effect of warping is to consider that sections
do not remain perpendicular to the beam axis.
EIw′′′′ + mẅ = 0
Solve this equation by separating variables and then adding BCs and ICs.
w(x, t) = f (x)q(t)
w(x, t) = wi (x)ejωi t
Z L
mwi wj dx1 = µi δij
0
Z L
EI2 wi′′ wj′′ dx1 = ωi2 µi δij
0
Note:
• The mode shapes corresponding to distinct frequencies are orthogonal with respect
to the mass distribution.
• The mode shapes are also orthogonal with respect to the stiffness distribution.
Define the Rayleigh quotient
Z L
EI2 (wi′′ )2 dx1
ωi2 = Z0 L
m(wi )2 dx1
0
Substituting we get:
Z L
2
δqi µi (q̈i + ωi qi ) − wi q3 dx1 = 0 ∀i
0
Note that the forcing term for each equation is the work of distributed eternal forces
along each modal shape.
Each mode is energetically decoupled from the others. No energy transmission from one
to the other is possible.
Physically, this equation states that every mode behaves like a single d.o.f. oscillator of
mass µi and frequency ωi ; the generalized force is once again the work of the external
distributed force p on the mode wi .
Remeber:
lim εn = 0
n→∞
Note: To be complete the set of functions Ni must satisfy the kinematic boundary
conditions (Essential or Dirichlet).
Premultiply by Ni :
< Ni (x), w(x) >= αi < Ni (x), Ni (x) >
< Ni , w >
αi =
||Ni ||2
Mq̈ + Kq = Fp
6.8.2 FEM
We can subdivide the domain and apply for each the Ritz formulation. Then we can
proceed to assemble the matrix of the system starting from all the matrices of the single
elements.
• The prediction of local effects may require the inclusion of local shape functions
who may be difficult to define
• With well defined shape functions the convergence is fast (with few dofs)
FEM
• The prediction of local effects may be obtained with local refinement of the number
of elements
Mq̈ + Kq = 0, q = 0 ∈ RN
Where:
q = ϕi ejωi t
ϕi i-th eigenvector (eigenmode)
(−Mωi2 + K)ϕi ) = 0
KΦ = MΦΛ
Φ = [ϕ1 , ...., ϕN ]
Λ = Diag(ωi2 )
Note: The eigenvector is a direction for which the elastic force is equal and opposite to
the inertia force. If we move along the eigenvector we generate a movement with zero
work.
7.2 Normalization
We can normalize a modal shape unless for a constant:
ϕTj Mϕi = 0 if i ̸= j
ϕTj Kϕi = 0 if i ̸= j
q = Φz
Rewriting:
MΦz̈ + KΦz = 0
And
ΦT (MΦz̈ + KΦz) = 0
But these are all diagonal matrices. Therefore:
ΦT MΦ = Diag(µi )
ΦT KΦ = Diag(µi ωi2 )
We get N decoupled scalar equations:
µi (z̈i + ωi2 zi ) = 0 ∀i
7.4.1 Decoupling
Just as before we can write (droppings the tilde)
ϕTj Aϕi = λi ϕTj ϕi
We define:
• Algebraic multiplicity: the number of times an eigenvalue is repeated
• Geometric multiplicity: is the dimension of the eigenspace Ev associated with an
eigenvalue λ
When the AM and GM are the same we can still say that the eigenvalues are orthogonal
and we can decouple the problem through diagonalization.
For a free-flying aircraft the K matrix is 6 times singular: 6 rigid body modes.
It is very important to note that because ωi = 0, we cannot say anything about φTj Mφi ,
therefore rigid body modes may not be orthogonal with respect to mass. The only
condition for which the modes are orthogonal is when we consider as degrees of freedom
the rotation and translation of the CG.
Exploiting the definition of modal mass and the orthogonality of the modal shapes with
respect to mass we can write:
N
X
φTk Mq = φTk M φi (Ai cos(ωi t) + Bi sin(ωi t)) = φTk Mφk (Ak sin(ωk t) + Bk cos(ωk t))
i=1
Note: The rigid body modes make the system unstable because the solution grows
indefinitely with t.
This means that the Rayleigh Quotient is a good approximation of the associated eigen-
value ωi .
Because the Rayleigh quotient is quadratic in the error, then the first derivative of the
quotient with respect to u will be null. Therefore:
dR(u) dR(u) dR(u) du
= 0 =⇒ = =0
du dα du dα
6. Restart
In this way we are able to calculate a set of eigenvalues and associated eigenvectors,
in this case, we are interested in studying all of the modes and not just evaluating the
divergence pressure value.
• Dissipation associated with the straining of the material, due to viscoelastic be-
haviour of the constitutive law (very limited in a metallic material, more relevant
in composite structures due to matrix and to micro-sliding of fibers).
• Dissipation due to friction for sliding inner joints, support connections, etc. This
is extremely important for complex structures made of several parts.
F = Fe + Fd = ku + cu̇
ue = ud = u
Solution of the equation:
If F = F̄ = const.
F̄ −k/ct F̄
u(t) = − e +
k k
If u = ū = const.
F = kū
For small amplitude oscillations, the mechanism can be represented through a linear
contribution proportional to the time derivative of the free coordinates.
Fd = Cu̇
In resonance:
x = x0 cosωn t
Substituting we get:
|F0 | |F0 |
|x0 | = =
ωn c 2kξ
Knowing ωn , we can find ξ:
|F0 |
ξ=
2k|x0 |
R = kx + cẋ
p
R = kx − cωn x0 quad(ωn t) 1 − cos2 (ωn t)
s 2
x
R = kx − cωn x0 quad(ωn t) 1 −
0
The work: I I I
Wc = R dx = kx dx + cẋ dx
Wc = πcx20 ωn
Or we consider the elastic work from x to x0 , which is a quarter cycle.
Z x0
1
Wk = kx dx = kx20
x 2
Since we can measure both the WC and the Wk experimentally we can find the value of
ξeq .
C = αM + βK
1 α
=⇒ ξi = + βωi
2 ωi
The constants α and β could be computed starting from the experimental evaluation of
modal damping factors ξi .
The two constants are sufficient to obtain a perfect modal diagonal damping only if
In many cases we can consider it being diagonal. Despite it being heuristic, it may be
sufficiently accurate.
Fd = sign(ẋ)µN
1
Wk = kx20 Wc = 4µN x0
2
2µN
=⇒ ξ =
πkx0
In order to verify if the damping ratio behaves as such, we can experimentally increase
the amplitude of the oscillation.
We can see how the damping depends on the amplitude of the oscillation x0 in this
model. This is, in practice, not true. We have to correct ”manually” this behavior by
removing this dependency.
And
h
Fc = 2mξωn ẋ = ... = ẋ
2ωn
In frequency domain:
Fc = jhx
ξ
mẍ + k(1 + jη)x = 0, η =
2
Where η is the loss factor.
or
We are dealing with a system of differential equations in matricial form. The direct
solution in the time domain is generally impossible. We have to find a way to solve our
problem.
We can rewrite:
q(s) = (s2 M + sC + K)−1 F(s)
Where we define the transfer function G(s) as:
Calling:
q
z= , F = Tu
qd
We can rewrite it as:
0 I 0
ż = z+ u(t) ⇐⇒ ż = Az + Bu(t)
−M−1 K M−1 C M−1 T
In general, applying the Laplace transform is not very useful because it is difficult to anti-
transform and go back to the time domain. We generally cannot say that the function
in the Laplace domain is a linear combination of notable solutions of which we know the
inverse.
Since:
ejα + e−jα
cos α =
2
=⇒ F (t) = F̂ ejωt + F̂ ∗ e−jωt , F̂ ∈ C
I can therefore compute the generic complex response to a generic complex input. Then
by summing the response with its complex conjugate I can retrieve the correct real
response.
Or
q̂ = (−ω 2 M + jωC + K)−1 F̂
Where we call dynamic stiffness:
In this case, the analogue of the transfer function is the dynamic compliance:
Ĥ(ω) = K−1
dyn (ω)
So:
Diag(µi (−ω 2 + jω2ξi ωi + ωi2 ))ẑ = ΦT F̂
By inverting and changing variable again back to q̂:
1
q̂ = Φẑ = Φ ΦT F̂
µi (−ω 2 + jω2ξi ωi + ωi2 )
Getting:
N
X φi φT i
Ĥ(ω) = D̂i (ω)
i=1
µi ωi2
• for terms where ω << ωi the contribute is almost static (left of graphic).
Or
∞
X
f (t) = ĉn ejΩnt , ĉ−n = ĉ∗−n
n=−∞
Where: Z T /2
1
ĉn = f (t)e−jΩnt dt
T −T /2
Note: the Fourier transform is a restriction of the Laplace transform where s = jΩ:
F[f¨(t)] = −ω 2 F (ω)
Note: We are losing every information on the initial conditions. This is not a problem
if the system is stable (supposed a priori), since the homogeneous term dies out.
This is a very important distinction from the Laplace transform. This is why we require
that f (t) ∈ L1 .
Applying it to our system we get (remember that a forced response has q0 = q̇0 = 0):
We can rewrite:
q(ω) = (−ω 2 M + ωC + K)−1 F(ω)
Where we define the frequency response function H(ω) as:
And
q(ω) = H(ω)F(ω)
We can now anti-transform the solution:
Z ∞
−1 1
q(t) = F [H(ω)] = H(ω)F(ω)ejωt dω
2π −∞
This integral is easily solved using a quadrature formula, limiting the domain up to a
significant frequency ωmax and exploiting the symmetry of the domain of integration.
1 ωmax
Z
q(t) ≃ H(ω)F(ω)ejωt dω
π 0
Remember:
T −1 1 −1 1
Φ KΦ = Diag(µi ωi2 ) =⇒ (Φ KΦ)T
= Diag =⇒ K = ΦDiag ΦT
µi ωi2 µi ωi2
Note: qqs = RF(ω) is called the quasi-static correction and takes into account the
behavior of the high-frequency modes that behave statically.
The system’s response could be approximated by accounting for the dynamic response
of low-order modes included in the bandwidth plus the static response of the others.
Consider simplifying the expression taking C = 0 (no damping) and the modal forms
normalized at unit mass (so, µi = 1∀i). Separate the DoFs into rigid and elastic DoFs.
q = qe + qR = Φe ze + ΦR zR
It is possible to compute the acceleration that is necessary to satisfy the equilibrium of
rigid body equations (pre-multiplying the structural equation for the transposed of the
rigid body modes).
MΦR z̈R + MΦe z̈e + KΦe ze = F
No elastic energy is associated with rigid body modes.
ΦTR MΦR z̈R = ΦTR F
Unit mass normalization ΦTR MΦR = 1. Therefore we obtain the following important
result:
z̈R = ΦTR F
Substituting this in the original equation:
MΦe z̈e + KΦe ze = F − MΦR z̈R
MΦe z̈e + KΦe ze = F − MΦR ΦTR F
MΦe z̈e + KΦe ze = I − MΦR ΦTR F = PT F
2. The work of P T F for the elastic modes is the same as F for the same modes.
ΦTe PT F = ΦTe F =⇒ ΦTe PT = ΦTe =⇒ PΦe = Φe
How does this help us? We are now able to constrain the system to the ground with
dummy constraints without affecting the elastic modes.
m
!
−1
X φi φTi
RF = K − 2 1 F(ω) = q0e + qi
µω
i=1 i i
Since the system is in equilibrium as a rigid body it is possible to add dummy constraints
to the ground to remove rigid body modes. They are not working anyway and their
reaction on ground is always null.
K̃ = K + ΦR Kg ΦR
Now K̃ is not singular so ∃K̃−1 It can be demonstrated that it is valid ∀K̃
At this point we can find:
q0e = K̃−1 F = K̃−1 PT F
The term q0e corresponds to the static elastic modes which appear in the quasi-static
correction approach.
q0e = PK̃−1 PT F = PK−1 PT F
9.7 Anti-resonance
Consider an undamped system ξi = 0. The transfer matrix is going to be:
m
X φi φT i
Ĥ ≃ D̂i (ω) + R
i=1
µi ωi2
With:
1
D̂i (ω) = 2
1 + j2ξi ωωi − ω
ωi
Of such transfer function we want to study the collocated transfer function which relates
the displacement of node k with the force applied to the same node. Therefore of the
vector φi we just consider the k th component.
m
X φ2ik 1
Hkk = 2 + Rkk
µ ω2
i=1 i i
ω
1− ωi
Where:
h(t) = F −1 [H(ω)]
So the function h(t) is the response of the system to an impulse in time t=0.
This is:
(s2 m + sc + k)x = f0 =⇒ (s2 m + sc + k)x − f0 = 0
Remember:
L[f¨(t)] = s2 F (s) − sf (0) − f˙(0)
So we can also think of this system as a free response with a non-null initial condition:
mẍ + cẋ + kx = 0
x0 = 0, ẋ0 = f0
m
Integrating by parts:
Z t
f
q(t) = a(t)f (0) + a(t − τ ) (τ ) dτ
0 dτ
σ = Cε
d
ε= u = Du ≃ DNq = Bq
dt
σ ≃ CBq
The internal loads are the integral of stresses. As such they are proportional to the
elastic forces in the structure K.q
Or
Kq = (Kq)S + (Kq)F
If the bandwidth is limited then we could think of neglecting the fast modes. Kq ≈
(Kq)S . We need to assess the validity of such an approximation.
Remember that the following is the solution to the homogeneous problem (7.3):
KΦ = MΦDiag(ωi2 )
This shows that with the direct recovery, we cannot use the slow modes for the compu-
tation of the loads as the fast modes behave statically and this equation shows it.
The only particular case where the fast modes can be used is when the excitation F is
bandwidth-limited and F is spatially distributed in a regular way (as in lift loads). In
this way, the overall work done by the loads is going to be null.
Kq = F − Mq̈
ω2
2 2 2 1 T
q̈ = ω q = ω HF = ω ΦDiag 2 Φ F = ΦDiag ΦT F
ωi − ω 2 ωi2 − ω 2
Therefore:
ω2
Kq = F + MΦDiag ΦT F
ωi2 − ω 2
ω2
T ω
Kq = F + MΦS Diag 2 − ω 2 ΦS F + MΦF Diag ω 2 ΦTF F
ωiS iF
ω2
≈ F + MΦS Diag 2 − ω2 ΦTS F
ωiS
This method allows for recovery of the static deflection (and so internal loads) of fast
modes because it is equivalent to applying back to the full stiffness matrix the instanta-
neous loads as static loads. What is missing is only inertia (and damping) forces due to
fast modes.
There are cases where the excitation signaling terms of value assumed in space and time
are random and can only be defined through a probability of occurrence. In this case,
the excitation is called non-deterministic, and it is defined through a stochastic pro-
cess. The physical characteristics of a stochastic process are described by its statistical
properties.
An attempt to predict its value at a certain position and time can only be performed in
a statistical sense. An observed set of records cannot precisely be repeated, but it will
follow a certain pattern that may only be mathematically represented by statistics.
If the excitation random process is the cause of another process, this will also be a
stochastic process. A stochastic input will provide a stochastic output.
FX (x, t) = P(X(t) ≤ x)
2
σXX (t) = E[(X(t) − µX (t))2 ]
Cross correlation
RXY (t1 , t2 ) = E[X(t1 ), Y (t2 )]
µX (t) = µX
It is possible to start by taking the time history dividing it in several portions of size T
and use them as realizations.
So: Z
RX (τ ) = E[X(t + τ )X(t)] = − x(t + τ )x(t), dt
Cross-covariance
kXY (τ ) = RXY (τ ) − µX µY
Define the coherence:
kXY (0)
ρXY =
σx σY
We have that:
−1 ≤ ρXY ≤ 1
Note: If X and Y are Uncorrelated then ρXY = 0.
The more rapidly the auto-covariance decays with τ the less the shifted time histories
are similar.
So more rapidly the auto-covariance decays and more casual or chaotic is the signal.
RY Y (τ ) = RX X(τ ) + 2RN X (τ ) + RN N (τ )
RN N → 0 RN X → 0
So:
RY Y ≈ RXX (τ )
10.7.1 Properties
ΦXX ∈ R if kXX is EV EN
ΦXX ≥ 0
Z +∞
1
kXX = F −1 [ΦXX (ω)] = ΦXX (ω)ejωt dω
2π −∞
Since KXX (0) = σ 2 :
Z +∞
2 1
σXX = ΦXX (ω) dω
π 0
Kmq = mF
So:
mq = K−1 mF
Note: To compute the variance of the output we need the impulse response and the
auto-covariance of the input.
Defining as before:
By transposing the initial first equation and by applying the same procedure we obtain:
σx2ẋ = σxx
2
AT + σxu
2
BT (2)
2 2 T
σxu = σux
2 1
σxu = BW
2
2 T 2 1
=⇒ σxu = σux = WBT
2
Substituting in (3) we get:
2 2
Aσxx + σxx AT = BWBT (4)
This is the Lyapunov equation and it is a matrix equation that allows us to compute
the cross-variance if the states are subject to a white noise input.
2 we can compute the variance of the output:
Once found σxx
∆y = C∆x + D∆u
2 2
σyy = Cσxx CT + Cσxu
2
DT + Dσux
2
CT + Dσuu
2
DT
If the input is a white noise W, then:
2 2 T 1 T 1
σyy = Cσxx C + C BW D + D BW CT + DWDT δ(0)
2 2
Note: since the white noise has an infinite amount of energy (remember it’s only the-
oretical) we cannot have a direct connection from input to output. Basically we need
that D = 0. The relation becomes:
2 2
σyy = Cσxx CT
Given an input vector u with its assigned PSD Φuu , find the transfer matrix of the
state-space filter Hf that returns the same PSD subjected to a white noise input w.
The transfer function Hf can be determined in the following way. The transfer function
will be governed by the following state space system:
(
ẋf = Af xf + Bf uf
u = Cf xf
By adjusting the coefficients we can create the filter. Combining in a single state-space
state where we have an additional state which is the filter state xf . The input of the
system u is equal to yf = Cf xf .
ẋ A BCf x 0
ẋ = 0 A + n
xf B
f f
x
y = [C DCf ] x
f
Apply the Lyapunov equation to this system to obtain the variance of the output.
Note: if the PSD is not a polynomial ratio we CANNOT COMPUTE the needed
shape filter. We must use the direct method.
In this way we can determine the likelihood of an event and evaluate its consequences.
In order to compute the damage, we have to count the number of times the quantity of
interest crosses a certain threshold b.
The derivative of H will be different from zero every time H is discontinuous. We are
interested in the rate of threshold crossing is independent of the direction of crossing.
Therefore:
|ẋ|δ(x(t) − b)
The number of threshold crossing is equal to the integral of:
Z t2
N (b, t2 , t2 ) = |ẋ|δ(x(t) − b) dt
tq
We can compute the expected value for the number of threshold crossings per unit time.
Z +∞ Z +∞
N̄ = E[N (b, t)] = |ẋ|δ(x(t) − b)p(x, ẋ, t) = dx dẋ
−∞ −∞
T n
P (x < b, t ∈ [0, T ]) = lim 1 − N̄+ (b) = e−N̄+ (b)T
N →∞ n
1 − e−N̄+ (b)T
The n exponent comes from the fact that if the events are independent then the combined
probability is the product of all probabilities for every event.
Clearly we are making the strong hypothesis that what happens at a certain time is
independent of what happens at other time instances.
12.1 Interface
To create the loop between the structural dynamics and aerodynamics we need to first
translate the information of the displacement of the structural nodes into the displace-
ment of the aerodynamic nodes to compute the aerodynamic loads.
Once computed the displacements we need to convert back the information of the loads
into concentrated loads applied to the structural nodes.
N
Sy
Z
Sz
= IT P (x, t)A dS
T
S
My
Mz
The use of CFD is generally not the best option as it is very computationally expen-
sive. Furthermore, during the early design phase, there is no detailed information on
the geometry therefore the method is excessive. Another important drawback is that
the method depends on time marching integration therefore to pass into the frequency
domain we need to apply heavy post-processing. The use of CFD is therefore limited
only to viscous phenomena such as aileron buzz, buffeting, stall, etc.
In all other cases, we can rely on simplified models that are going to use simplified ge-
ometries (from full 3D geometry to 2D lifting surface) and simplified equations (from
NS to Euler or potential equations).
Incompressible RANS
Compressible RANS
NS → (
Incompressible Euler
Euler equations →
Full potential flow → Linearized potential
In aerodynamics instead the local effects are extremely relevant. As an example, the
local pressure distribution may have major consequences on the rest of the airfoil (BL
stability).
This implies that in the aeroelastic models when we approximate the aerodynamic ge-
ometry as a flat plate and therefore lose the local effects, we need to be careful that we
work within the hypothesis of the approximation.
df ∂f X ∂f ∂xi ∂f
= + = + u · ∇f
dt ∂t ∂xi ∂t ∂t
i
Note that df Df
dt = Dt is true only in a Lagrangian setting. Otherwise, the material deriva-
tive is not strictly a derivative.
d u2
1
= − u · ∇P
dt 2 ρ
For inviscid flows, the rate of change of kinetic energy is equal to the work done by
pressure forces (there is no dissipation through viscosity).
12.8 Entropy
Recall the 1st principle of thermodynamics:
1
de = T ds − P
ρ
We use the first principle of thermodynamics inside the equation of energy balance
(Euler).
Ds
= 0 =⇒ s = s0 = const
Dt
We find that in an inviscid fluid flow, the total time derivative of entropy in the flow is
null. This means that if there is no initial gradient of entropy in the fluid then the flow
will remain isoentropic.
Now let’s differentiate Γ and exploit the relation between acceleration and vorticity and
using the momentum balance equation we get:
∇P ∇P
Z Z Z I
dΓ d
= ∇ × u · n̂ dS = ∇ × a · n̂ dS = − ∇ × · n̂ dS = · dl
dt dt S S S ρ ∂S ρ
Therefore:
∇P
I I
dΓ dP
= · dl =
dt ∂S ρ ∂S ρ
This means that an isoentropic flow that is also barotropic. In fact, P = P (ρ, s0 ) = P (ρ).
Then the flow will be irrotation, given an initial null vorticity.
I
dP dΓ dω dΓ
= 0 =⇒ = 0 =⇒ = lim =0
∂S ρ dt dt S→0 dt
dΓlm
=0
dt
Kelvin’s theorem has a major implication in the phenomenon of the sudden start of a 2D
airfoil. Indeed for Kelvin’s theorem to be valid, the overall vorticity inside the boundary
∂S must be null therefore it must generate a vorticity that is equal and opposite to the
one created by the 2D airfoil, Γw .
d ∂Φ |∇Φ|2
dP dP dh
dh = =⇒ =ρ =ρ +
ρ dt dt dt ∂t 2
Recalling again the Bernoulli equation we can define an as:
∂Φ |∇Φ|2 − U∞ 2
2
a (Φ) = −(γ − 1) + + a2∞
∂t 2
Therefore by combining the two together, we get the full velocity potential equation:
ρ d ∂Φ |∇Φ|2
2
ρ∇ Φ = +
a2 (Φ) dt ∂t
2
2 − U2
2 ∂Φ |∇Φ| ∞
+ a2∞
a (Φ) = −(γ − 1)
+
∂t 2
In order to solve the problem we need to ass the boundary conditions, by defining the
lifting surface B(x, t) = 0, we then impose that the velocity of the fluid normal to the
body boundary is equal to the velocity of the boundary:
lim u = lim ∇Φ = U∞
x→∞ x→∞
Note: This equation is very complicated to solve and it results even more difficult than
resolving directly the Euler equations
The Kutta condition is a condition that says that the flow exits the trailing edge regularly
without curving around it as it would be non-physical. This same condition can be seen
also as a condition on the regularity of pressure as the pressure at the TE is equal
between the top and bottom and it can also be seen as a condition that the vorticity at
the TE is null.
The Kutta condition allows us to impose a vorticity on the airfoil such that the trailing
edge becomes a stagnation point.
a = a∞ + δa
P = P∞ + δP
ρ = ρ∞ + δρ
u = U∞ + δu
Therefore:
Φ = U∞ x + φ
This directly implies that:
∇2 Φ = ∇2 φ
∂Φ ∂φ
=
∂t ∂t
Furthermore, we can exploit the following two approximations:
|u|2 U2 ∂φ
≈ ∞ + U∞
2 2 ∂x
d ∂ ∂
= + U∞
dt ∂t ∂x
Therefore using these approximations inside the potential equation, we get the linearized
potential equation:
2
∂2φ 2
2 1 ∂ φ 2 ∂ φ
∇ φ= 2 + 2U∞ + U∞ 2
a∞ ∂t2 ∂t∂x ∂x
This equation is not valid only in the presence of strong shockwaves which are going to
break the irrotationality hypothesis.
∇2 φ = 0
We can write:
bu = N (x)q(t)
Therefore the linearized boundary condition becomes:
∂N ∂φd
N (x)q̇(t) + U∞ q(t) = =W
∂x ∂z
Note: We can even subdivide the deformation potential into sub-components which are
related to the individual shape functions.
Note: By dividing the boundary conditions by U∞ we can see that the first term relates
to the change in AoA due to kinematics while the second relates to a local change in
AoA due to the geometric deformation.
N (x)q̇(t) ∂N W
+ q(t) =
U∞ ∂x U∞
1 −∆x
w= γ
2π ∆x2 + ∆z 2
Since there is a continuous array of vortices, the sum of all the effects is:
Z xT E
1 z
u(x, z) = γ(ξ) dξ
2π xLE (x − ξ)2 + z 2
Z xT E
1 x−ξ
w(x, z) = γ(ξ) dξ
2π xLE (x − ξ)2 + z 2
The induced speed along the vortical line z = 0 is:
u(x, 0) = 0
Z xT E
1 γ(ξ)
w(x, 0) = dξ
2π xLE x − ξ
13.1 Hypothesis
• 2D irrotational, incompressible flow undergoing unsteady motion in uniform flow
• Airfoil chord: 2b
r s
b
b−x b + ξ wA (ξ, t) − λ(ξ, t)
Z
2
γB (x, t) = − dξ
π b+x −b b−ξ x−ξ
This, though, is not a proper solution as we still don’t know how to express λ which is
a function of γB . Therefore this is an implicit equation in γB .
x = b cos θ
Write:
∞
X ∞
X
λ(θ, t) = λn cos nθ wa (θ, t) = wan cos nθ
n=0 n=0
Where: Z π Z π
1 1
λn = λ(θ, t) cos nθ dθ wan = wa (θ, t) cos nθ dθ
π 0 π 0
Plugging these equations in (1) we can get:
∞
X
γb = 2 (wan − λn )fn (θ)
n=0
Where:
tan θ
n=1
fn (θ) = 2
sin(nθ), n ≥ 1
γb = γbC + γbN C
In particular, γbC is responsible for Γ (0 and 1 terms) but with zero induced speed. So
that when we integrate it to find Γ:
Z b
Γ= γbC dξ
−b
Z b
1 γbC (ξ, t)
wb = dξ = 0 ∀x ∈ [−b, b]
2π −b x − ξ
2 1 1
γbC = wa0 + wa1 + λ0 + λ1
sin θ 2 2
On the other hand, for the non-circulatory component, it is exactly the opposite: it will
have no effect on the circulation but it will create an induced speed.
Z b
γbN C dξ = 0
−b
Z b
1 γbC (ξ, t)
wb = dξ = (wa − λ) ∀x ∈ [−b, b]
2π −b x−ξ
∞
2 1 X
γb N C = − (wa0 − λ0 ) cos θ + (wa1 − λ1 ) cos 2θ + 2 (wa n − λn ) sin nθ
sin θ 2
n=2
Note: The non-circulatory component depends on all components of w and λ with respect
to the circulatory component that depends only on the first two components.
Where ∆φ is the jump of potential across the airfoil, due to the distributed vorticity.
Remember that:
γb = ∆u(x) = u(x, 0+ ) − u(x, 0− )
Then, we know that:
∂φ
u=
∂x
So:
∂∆φ
γb =
∂x
But: Z x Z x
∂∆φ
Γx = γb dξ = dξ = ∆φ =⇒ Γx = ∆φ
−b −b ∂ξ
This also implies that:
∂∆φ ∂Γx
=
∂t ∂t
Substituting in (1):
∂Γx
∆p(x, t) = −ρ U∞ γb +
∂t
And Z x
∂
−∆p(x, t) = ρ U∞ γb + γb (ξ) dξ
∂t −b
We have the expression of γb :
∞
X
γb = 2 (wan − λn )fn (θ)
n=0
So:
∞
X
−∆p = pn fn (θ)
n=0
p0 = 2ρU∞ (wa0 − λ0 )
13.5.1 Loads
Now that we have the pressure along the chord we can calculate everything.
Z b ∞
X Z π
L(x, t) = − ∆p(ξ, t) dξ =⇒ L(x, t) = pn fn (θ)dθ
−b n=0 0
Note: Only the components multiplied by f0 and f1 have integrals over the chord different
from zero!
Note: It is possible to verify that only the non-circulatory components of bound vorticity
led to contributions to terms depending on derivatives with respect to time.
∂∆φ
γb =
∂x
And
Γx = ∆φ
The jump of potential along the wake is the jump of potential in the trailing edge
computed with a delay in time: when the wake element was at the TE.
13.6.2 Solution
Z ∞
1 1
λ(x̃, t) = ∆φw (ξ, t) dξ
2bπ 0 x̃ − ξ
∆φw = Γe−jk(x̃−1)
Take now:
∂∆φ ∂∆φ
∆p = −ρ U∞ + (1)
∂x ∂t
Transform it:
−ρU∞ ∂∆φ
∆p = + jk∆φ
b ∂ x̃
The lift is:
LC = LQ + Lw
Where:
LC = ρU∞ G(k)Γ(k)
This is a generalization to the frequency domain of the Kutta-Joukowsky formula.
• For higher k the oscillation has a hysteretic behavior, as seen before. The CL lags
after the instantaneous angle of attack. This results in energy being pumped into
the system: we have the flutter.
• For k → ∞ we have a purely real C(k). The ellipsis collapses into a line again
with half the slope.
Figure 58: Effects on the normalized CL for a pure pitch oscillation around the LE
The non-circulatory part starts from zero amplitude and then increases almost quadrat-
ically (∝ ω 2 ) and the phase shift is always +90° because acceleration is 90° in advance
of speed. The anticipation of lift with respect to wing movement determines if we are
injecting or absorbing energy inside the system therefore crucial for flutter identification.
Cicala, Küssner, and Schwarz models can be used to obtain the response for an arbitrary
movement.
With:
Cp0 = −wa1 + C(k)(wa0 − wa1 n = 0
Cpn = wan + jk (wan+1 − wan−1
2n
2 jk
L = 2πρU∞ (wa0 − wa1 )C(k) + (wa0 − wa2 )
2
2 jk
M0 (k) = πρU∞ (wa0 C(k) − wa1 (1 − C(k)) − (wa1 − wa3 ) − wa2
4
This approach could be used if the airfoil changes shape. For instance, a change of
camber is the one that is generated when a movable surface is rotated by an angle β.
Response to a step change in α. Define the indicial function (or Wagner function) ϕ:
Z +∞
2 C(k) jkτ 2
L(τ ) = ρbU∞ α e τ = 2πρbU∞ ϕ(τ )α
−∞ jk
To get the overall lift and sum the circulatory and non-circulatory components, we
exploit the Duhamel convolution for the circulatory part.
Z τ
2 dα
L(τ ) = LN C (τ ) + 2πρbU α0 ϕ(τ ) + ϕ(τ − σ) dσ
0 dσ
To do so, at every instant in time we are going to find the speed at the collocation point
by imposing the boundary condition we impose:
Γi
|vi | =
2πd
We then compute lift as:
Z b
∂Γ
L= −∆P dx = ρ U∞ Γ(t) + 2b
−b ∂t
The airfoil can be split into portions called panels and the same procedure can be applied
to each panel. Therefore we are going to impose as many boundary conditions on the
collocation points as is the number of panels (and therefore of lumped vortices).
Another thing we can do is to compute the induced velocity in the location of every
lumped vortex in the stream and therefore consider also the motion of the vortex. This
is called the ”free wake vortical method”.
vG = vG (ξ)
The aircraft, initially at a distance x0 from the gust front, travels at U∞ . Call x the
position of a point of the aircraft on the local reference frame.
ξA = U∞ t − (x − x0 )
Consequently, the gust speed in the local reference system of the a/c is:
So: Z +∞ ξ+x+x0 1
vG (x, ω) = vG (ξ))e−jω U∞ dt
−∞ U∞
Defining now x̃ = x/b and ξ¯ = ξ/b:
Z +∞
−jk(x̃+x̃0 ) b ¯ −jkξ̄ dξ¯ =
vG (x̃, ω) = e vG (ξ))e
U∞ −∞
13.11.2 Gust
The normal velocity of the body minus the gust should be equal to the velocity of the
flow in z;
vbz − vG = wa
Since the problem is linear we can say:
wa = −vG
We can express the first term as a Fourier series through a Bessel Function of I kind:
∞
X
ejz cos θ = J0 (z) + 2 j n Jn (k) cos nθ
n=1
We can represent the gust speed where wn is composed of different Fourier components
due to the gust speed.
∞
!
X
vG (θ, k) = −wa (θ, k) = −W (k)e−jkx̃0 J0 (k) + 2 (−j)n Jn (k) cos nθ
n=1
b
M0 (k) = L(k)
2
The function ψ (Sears function) is defined as:
It starts as C(k) and then is transformed into a more difficult expression due to the
exponential term of delay. Combining the movement in the circle of the exponential and
the C(k) function we get the spiral movement in the figure.
We get that the response to the plunge and the gust differ due to the ex-
ponential term. Physically what’s changing is the so-called penetration effect: the
gust speed perceived at the TE is different than the gust speed in the other points of
the profile.
When k is small the exponential term is not very important and we can consider every-
thing without the exponential: the gust response is equivalent to the plunge response.
At high k the perceived effect is as if the TE is plunging in a different way than the LE
edge.
• The advantage is to lead to a very simple integral expression that can be easily
solved using a numerical approach
• Several authors worked on this formulation. However, the most complete work was
developed by Prof. Morino who applied it to subsonic and supersonic flows. This
is why here it will be nominated as “Morino’s method”
• This method represents a prototype for all possible panel methods in 2D/3D com-
pressible/incompressible subsonic/supersonic flows
∇2 G = δ(|x − x0 |)
The solution in 3D is
1
G=−
4π|x − x0 |
Where
0
if x0 ∈
/ V,
E(x0 ) = 1 if x0 ∈ V and x0 ∈
/ ∂V,
1
2 if x0 ∈ ∂V.
Using this in the Green’s identity we get
Z Z
∂φ ∂G
E(x0 )φ(x0 ) = −G +φ dS + φ∇2 G dV
∂V ∂n ∂n V
This expression can be used to compute the potential φ when the solution G is known.
∇2 φ = 0
∂V = Σ ∪ B ∪ W
Or
∂φ
uW · nW = 0 −→ =0
∂n W
We get Z Z Z
∂G ∂φ ∂G
E(x0 )φ(x0 ) = φ − G dS + φ dS
B ∂n B ∂n W ∂n
∆Cp |W = 0
Through Bernoulli
2 ∂∆φ ∂∆φ
∆Cp |W =− 2 U∞ + =0
U∞ ∂x ∂t W
The jump of potential on the wake at the TE is equal to the circulation at the span
station y as the Kutta condition says
For x0 ∈ V and x ∈ R3
Z Z Z
1 1 ∂ 1 1 1 ∂φ 1 ∂ 1
φ(x0 ) = − φ + dS − ∆φ dS
2 4π B ∂n r 4π B r ∂n 4π W ∂n r
The obtained integral expression can be used to compute the potential on every point in
the flow field when we know the potential on B or to compute the potential on B itself.
14.3.2 Discretization
1. Divide the wing surface in N = Nx × Ny panels and the wake in Ny semi-infinite
strips.
2. Define the type of approximation of φ over each panel (i.e., 0th-order: φ constant,
1st-order: φ linear, . . . ).
3. Decide how and where to impose the BCs on each panel (i.e., for the 0th-order
panel, the best collocation point (the point where the boundary condition is ap-
plied) is at the center of the panel.
N Z N Z Ny Z
X ∂ 1 X ∂φk 1 X ∂ 1
2πφi = − φk ds + ds − ∆φT Ej ds
Bk ∂nk rik ∂nk Bk rik Wj ∂nj rij
k=1 k=1 j=1
Ny Z N Z
X ∂ 1 X ∂ 1
∆φT Ej ds = hk φ k ds
Wj ∂nj rij Wj ∂nk rik
j=1 k=1
0
k is not a TE panel
hk = 1 k is a TE upper panel
−1 k is a TE lower panel
Note: It is not necessary to have potential values associated with the wake panels. The
jump of the potential of the wake panel is equal to the difference between the potential
of the corresponding TE panels (imposition of the Kutta condition).
14.3.3 System
We can write everything in a linear system
Define
∂ φ̃ ∂Nφ
= = Pφ
∂x ∂x
Remember the BC on the body (steady)
∂φ ∂η
= U∞
∂n B ∂x
So
∂η
Yφ = ZU∞
∂x
Then calling A = Z−1 Y
Aφ = U∞ α
Where
∂η
α=
∂x
Then
φ = A−1 U∞ α
Substituting in CP
2 ∂ φ̃ 2 2
CP = − =− Pφ = − PA−1 U∞ α
U∞ ∂x U∞ U∞
The pressure distribution is
Cp = −2PA−1 α
∆Cp |W = 0
Through Bernoulli
2 ∂∆φ ∂∆φ
∆Cp |W =− 2 U∞ + =0
U∞ ∂x ∂t W
Substitute: x̃ = x/b
∆φ = Γ(y)e−k(x̃−1)
The jump of potential along the wake panels is not constant anymore: it is equal to
the product of the jump of potential at the trailing edge multiplied by a delay operator
(the exponential term) that expresses the fact that the vorticity travels in the wake at
the asymptotic speed.
14.4.2 Discretization
The system is
∂φ ∂η jω
Yφ = Z = ZU∞ + η
∂n ∂x U∞
14.4.3 Loads
2 jω
Cp = − Pφ + φ
U∞ U∞
∇2 φ̂ = 0
Problem formulation
We have to define the radius:
rβ2
r̂2 = (x̂ − x̂0 )2 + (ŷ − ŷ0 )2 + (ẑ − ẑ0 )2 =
β2
The sign of rβ2 depends on the nature of the flow: positive if subsonic, negative if super-
sonic.
Z Z Z
∂ 1 1 ∂φ ∂ 1
2πφ(x0 ) = φ ds − ds + ∆φT E ds
B ∂n rβ r
B β ∂n W ∂n rβ
Z Z Z
∂ 1 ∂ 1 1
Yik = 2πδik + ds + hk ds Zik = ds
Bk ∂nk rβik Wk ∂nk rβik Bk rβik
In frequency domain
ω2
∇2 G + G = δ(|x − x0 |)δ(t)
a2∞
The general solution:
1 jωτ
G=− e
4πrβ
Where τ is the time required by the perturbation to travel from x0 to x:
Explanation of τ
1 rβ − M∞ ∆x D + M∞
τ = τ1 = =
a∞ β2 U∞
So
1 −j aω D+
G=− e ∞
4πrβ
We get:
+
Z +
Z
jω D ∂ 1 D
−jk(x̃−1) jω a∞ ∂ 1
Yik = 2πδij + e a
∞ dS + hj e e dS
Bj ∂n̂ j rij Wj ∂n̂ j rij
Z
D+ 1
Zik = ejω a∞ dS
Bj rij
From a panelization point of view, the perturbation will not completely affect all of the
panels. Indeed the cone of influence will affect partially some panels therefore in order
to discretize correctly we are going to need small enough panels.
If M >> 1 2γ
p γ−1 w γ−1
= 1+
p∞ 2 a∞
w
If a∞ >> 1 then with a first order expansion
1 2 2 w
p − p∞ = ρ∞ a∞ w = U∞
2 M ∞ M∞
2
Cp = α
M∞
Note that we have a negative altitude. It could happen that we are flying in a higher-
density air. This can be represented in the standard atmosphere by defining negative
altitude. We have to prove that within the green envelope, we have no flutter.
Where does it typically have flutter? In the high EAS low altitude (high q)
ηa = Hηs
So it will result: Z
δWa = q δ(H∗ηs )T · na Cp (xa ) dSa
As
Z
δWa = qδqT NT ηsT · na Cp (xa ) dSa
As
Or:
XZ
T
δWa = qδq NT ηsT · na Cp (xai ) dSa
i A si
Cp = P(k)φ
∂φ
φ = Y(k, M )−1 Z(M )
∂n
But:
∂φ ∂ηa jω ∂ηs jω jk
= + ηa = H + ηs = H N/x + N q
∂n ∂x U∞ ∂x U∞ b
So:
∂φ
= (A + jkB)q
∂n
Substituting:
δWa = qδqT DP(k)Y−1 (k, M )Z(M )(A + jkB)q
=⇒ Fa (k) = Ham (k, M )q
The other technique is to apply the work in the frequency domain using the Laplace
transformation:
(Ms2 + sC + K − qHam (p, M ))q = 0
sb σb jωb
Where p = U∞ = U∞ + U∞
In general we don’t have the full transfer function Ham (p, M ) but we have only available
the frequency response i.e. Ham (k, M ).
We need to find methods to compute the correct solution to the eigenvalue problem
having this partial information.
∂H s2 ∂ 2 H
H(s) = H|s=0 + s + + ...
∂s s=0 2 ∂s2 s=0
We get
s2 ∂ 2 H
∂H
Fa (s) = H|s=0 + s + + ... q(s)
∂s s=0 2 ∂s2 s=0
Using the inverse Laplace transformation:
∂H(0) −1 1 ∂ 2 H(0) −1 2
L−1 → Fa (t) = H(0)L−1 [q(s)] + L [sq(s)] + L [s q(s)]
∂s 2 ∂s2
∂H(0) 1 ∂ 2 H(0)
Fa (t) = H(0)q(t) + q̇(t) + q̈(t)
∂s 2 ∂s2
We know that
∂H(0) 1 ∂ 2 H(0)
H(0) = Ka = CA = Ma
∂p 2 ∂p2
Since H = H(p, M ) we have
∂H ∂H ∂p ∂H b
= =
∂s ∂p ∂s ∂s U∞
We get that
∂H ∂Im(H) ∂2H ∂ 2 Re(H)
= = −
∂p ∂ω ∂p2 ∂ω 2
We can now plug these results in the equation above and find the flutter equation.
Therefore:
y = (CA−1 B + D)u − CA−2 B u̇
The QS approximation of order n is equivalent to saying that the derivative (n+1) in
time is negligible.
In practice, it is more convenient to compute all the (desired) eigenvalues for a range of
EAS velocities (or dynamic pressures) starting from zero and evaluate their real part.
Flutter is detected as the lowest airstream speed at which the real part of at least
one eigenvalue crosses the zero axis from negative to positive. The knowledge of all
eigenvalues from zero to the desired airstream velocity not only allows the analyst to
check for flutter clearance but also to determine how the eigenvalues evolve with the
EAS velocity allowing an experimental verification of the aircraft aeroelastic stability.
See Figure 68 below.
Another problem is that in this way we are not investigating any other condition than
the flutter one. This leads to a validation problem. Normative imposes to asses the
model by using q < qF (that are measurable too). The certification process consists
of comparing the damping obtained by the model (at q << qf ) and the one measured.
Using the direct approach we don’t have access to these results. Hence it cannot be used
to validate the mode.
Since we are not interested in modes with high damping, associated with big Real part
Re(λ) ̸= 0 (they don’t have an effect on the flutter speed) we can think about linearizing
with Taylor’s series around s = jk in this way we are only restricting the real part around
zero. But this is exactly what we are looking for: low damping modes are associated
with Re(λ) ≈ 0.
p = jk
Expanding through a Taylor series
∂Ham 1 ∂ 2 Ham
Ham (p) ≈ Ham (k) + (p − k) + (p − k)2 + . . .
∂p 2 ∂p2
Im(Ham )(k)
Ham (p) ≈ Ham (k) = Re(Ham )(k) + jk
k
Im(Ham )(k)
Ham (p) ≈ Re(Ham )(k) + p
k
On this formula, we can build an algorithm to calculate the flutter speed
1. Compute a preliminary guess for one eigenvalue λ0i . i.e. we can use a structural
eigenvalue for small speeds
Note 2: The smaller the damping, the better the approximation. When we are at flutter
speed then the approximation is exact. This is the reason why a zero-order is sufficient
to have a good approximation.
Note 3: This method is robust as the error becomes smaller and smaller the closer we
get to the flutter condition.
f (s)q = 0 =⇒ F(s, q) = 0
Consider q = qi + ∆q and s = si + ∆s
∂f ∆s + ∂f ∆q = −f (si , qi )
∂s ∂q
∗
2qi ∆q = 1 − q∗ q
Iterate until ||∆q|| < ε and |∆| < ε to find the solution.
We start from a very low speed (close to zero) and as an estimate of the eigenvalues
we use the structural dynamics ones. To obtain an initial guess at a new speed it is
possible to use the value at the previous speed or try to estimate the eigenvalues and
eigenvectors at the new speed.
• Gusts
EASA CS-25 says “If deflection under load would significantly change the distribution
of internal loads, this re-distribution must be kept into account”.
a) Retaining a rigid aircraft model and modifying the aerodynamics to keep flexibility
into account.
q̈E ≈ 0, q̇E ≈ 0
So from the first equation
q∞
MRR q̈R + MRE q̈E − (CaRR q̇R + CaRE q̇E ) − q∞ (KaRR qR + KaRE qE ) = FeR
U∞
We get
q∞
MRR q̈R + − CaRR q̇R − q∞ (KaRR qR + KaRE qE ) = FeR
U∞
From the second equation
q∞
MER q̈R +MEE q̈E +CEE q̇E +KEE qE − (CaER q̇R +CaEE q̇E )−q∞ (KaER qR +KaEE qE ) = FeE
U∞
As before
q∞
MER q̈R + KEE qE − CaER q̇R − q∞ (KaER qR + KaEE qE ) = FeE
U∞
Define as usual KAE = q∞ KaEE .
KAE is the aeroelastic stiffness (that is singular at the divergence dynamic pressure qD)
We can consider this effect by adding a new term in the rigid equation as below
−1
q∞ −1
MRR + q∞ KaRE KAE MER q̈R + − CaRR − q∞ KaRE KAE CaER q̇R +
U∞
The equation has the same format as the rigid ones, but the matrices contain new
aerodynamic derivative terms (green).
If MER = 0 then there are no aerodynamic stability derivatives that are proportional to
rigid accelerations (as usually in Flight Mechanics).
Note: if the proper orthogonal modes of a free-free structure are used, MER = 0 by
default.
It is essential to model the behavior of a control surface and its actuation behavior.
Consider the case where a servo-system is applied. We will consider the approach
based on Laplace transform only for easiness of the exposition, but the same steps could
be followed in any approach.
Specific shape functions used to represent the rigid rotation of the movable surface will be
added to the usual shape functions. In this way, a specific degree of freedom associated
with the rotation of the movable surface will be defined
The second equation could be used to identify the load Fδ that must be applied by the
actuator/pilot to keep the position δ. This load is in general, a function of time (so a
control system is required to adapt it to keep the delta constant.
2 q Feq
(s M + xC + K − qHam (p, M )) =
δ Feδ
Write
Zqq Zqδ q Feq
=
Zδq Zδδ δ Feδ
Keep into account that this model is still ideal, because also, in this case, the force Fδ
must change instantaneously to adapt to changes of v.
δ = Hservo (s)δc
v = G(δc − δ)
δ̇ = P v
In order to be more accurate we should include the deformability of the control chain.
So:
PG
δ̇ = P Gδc − P Gδ −→ δ = deltac
s + PG
More general format
PG
δ= δC
(s + P G)(s + 2ξω0 s + ω02
2
The dynamics of the control system may become coupled with the aeroelastic dynamics
and we need to assess the stability of the overall systems. Indeed an elastic deformation
of the wing may be seen as a rolling action by the sensor therefore the control input may
worsen the situation. The sensor must filter out the aeroelastic dynamics. The analysis
of the modal shapes may help the decision on where to locate the sensors.
In the case where there is no servo-actuator too (Hservo = 1), it is possible to have δ
and δc different because the control chain is, in reality, an elastic system.
For design purposes, atmospheric change of wind speed is divided into two ideal cate-
gories
1. discrete gusts, where the gust velocity varies in a deterministic manner, usually
in the form of a ‘1-cosine’ shape
2. continuous turbulence, where the gust velocity is assumed to vary in a random
manner
1 − e−N̄+ (b)T
Where 2
1 σẋẋ − 21 σb2
N̄+ (b) = e xx
2π σxx
17.2.1 Cp in Morino
2 ∂∆φ ∂∆φ
∆Cp |W =− 2 U∞ + =0
U∞ ∂x ∂t W