You are on page 1of 71

Optimization of windtunnel

[Title]
tests for flight flutter prediction
[Title]
Jawad Bessedjerari

[Student’s name (first name and family name) ]

Master thesis submitted under the supervision of


Prof. Dr. Ir. Rik Pintelon
The co-supervision of
Dr. Master thesis submitted under the supervision of
Ir. John Lataire & Dr. Ir. Julien Ertveldt
Academic year In order to be awarded the Master’s Degree in
Prof. [first name and family name]
2017-2018 Electromechanical Engineering: Aeronautics
The co-supervision of
Prof. [first name and family name]

Contents

1 Introduction 1

2 Aeroelastic flutter: a theoretical background 3


2.1 Qualitative description . . . . . . . . . . . . . . . . . . . . . . . 3
2.2 Mathematical description . . . . . . . . . . . . . . . . . . . . . 5
2.3 Grey box modelling: transfer function denominator polynomial-
and extrapolation order determination . . . . . . . . . . . . . . . 13
2.4 List of simplifications and flutter parameters . . . . . . . . . . . 15

3 Grey box modelling: experimental setup and methodology 17


3.1 Experimental setup: the AATB . . . . . . . . . . . . . . . . . . 17
3.2 System identification for the characterization of flutter behaviour 21
3.2.1 Signal design: odd multisine excitation . . . . . . . . . . 21
3.2.2 Non parametric identification: the local polynomial method 22
3.2.3 Parametric identification: the Maximum Likelihood method 26

4 Results 30
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.2 First series of experiments: setup without additional mass . . . . 31
4.2.1 Identification . . . . . . . . . . . . . . . . . . . . . . . . 31
4.2.2 Error analysis . . . . . . . . . . . . . . . . . . . . . . . . 37
4.3 Second series of experiments: increased inertial coupling through
added mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.4 Final series of experiments: further increase of inertial coupling . 44

5 Flight flutter prediction 49


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5.2 Flutter speed prediction: extrapolation of transfer function coef-
ficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5.3 Validation: evaluation of the extrapolations . . . . . . . . . . . . 53
5.4 Validation: error on the extrapolated poles . . . . . . . . . . . . 58

i
6 General conlusion 62

Appendices 66

ii
Abstract

Aeroelastic flutter is a well known instability phenomenon caused by aerodynamic,


inertial and elastic couplings in an aeroelastic system, resulting in a complete
structural failure. Hence, in the phase design of an aircraft structure, flight flut-
ter tests are indispensable. This thesis presents the classical approach to flutter
analysis performed on the Active Aeroelastic Test Bench(AATB), providing on
the one hand a descriptive manual for the conduction of flutter experiments on
the AATB and on the other hand the demonstration of the capability of the AATB
to bring forth reliable results. A theoretical description of the nature of flutter is
given and a ready-made procedure to implement a white box model into a simu-
lation if presented before presenting the experimental setup and the methodology
followed for the generation of time invariant, linear models for the estimation of
the damping behaviour as a function of the wind speed of the aeroelastic system
for different configurations. Finally, the experimental obtained damping ratios
are extrapolated towards higher velocities. In doing so, the mathematical model
obtained through a white box approach is integrated in the extrapolation pro-
cedure, yielding reliable extrapolated dampings. This leads to the prediction of
the velocity for which the damping of the aeroelastic system changes its sign,
i.e. the flutter velocity. However these extrapolations are evaluated through a
validation procedure and a method is described to determine error bounds on the
extrapolated dampings.

iii
Chapter 1

Introduction

In the past century, many studies were made to characterize flutter. This in-
stability phenomenon can occur in structures subject to aerodynamic forces like
airplane wings and bridges and is the consequence of a coupling of those forces
with the natural modes of vibration of the structure, leading to quite catastrophic
situations, like the well known Tacoma Narrows bridge disaster. As the beginning
of the twentieth century knew a fast paced progress in the field of aerodynamics,
partly due to competition in weaponry between great powers during the world
wars, the subject of flutter became more and more understood.

Flutter occurs in certain velocity ranges and for certain aerodynamic, elastic
and inertial properties. Consequently, the classical approach to the prevention
of flutter for e.g. wings begins at an adapted design of the structure to the in-
stability phenomenon based on theoretical descriptions, after which ground test
vibrations and in flight flutter tests are performed for a range of velocities within a
prescribed flight envelope of which the results are used to make models which can
produce estimations of the damping and the resonance frequencies as function
of the velocity. The obtained dampings are then extrapolated towards velocities
which are outside of the flight envelope to make predictions of where the flutter
speed could be located, without necessarily damaging any flight components.
This approach is based on system identification techniques, which produce so
called black box models by studying the output of a (mechanical)system as a
result of an input, reducing the modelling effort significantly.

In this thesis, a recently developed test bench is tested for the application
of the above approach to predict flight flutter for a cantilever wing in a wind
tunnel. This experimental setup has some advantages with respect to others like
for example that it allows precise motion of a cantilever wing using linear actu-
ators, compared to conventional ones equipped with shakers responsible for the

1
excitation. In doing so, a guide has been developed for the conduction of such
experiments on the new test bench. The results however are prone to verification
as they are based on extrapolations. This is ensured through a validation proce-
dure. Furthermore, a ready made procedure to implement a white box aeroelastic
model into a simulation is presented.

This thesis starts with a description of flutter in an attempt to give an in-


tuitive understanding of its nature followed by a more mathematical description,
serving as basis for the mathematical modelling based on vibration experiments
in a later stadium. This mathematical description provides the means to imple-
ment a white box model into a simulation. In this same chapter, a summary
is made of the simplifications and a flutter influencing parameter is put forward
and translated into an experimental setup parameter, providing the means of
investigating the effect of this parameter on the flutter behaviour. Next, an
elaborate explanation of the used identification techniques is given, providing an
understanding of the manipulation of the data to produce whether or not reliable
linear models used to extract the damping behaviour of the cantilever wing as
function of the wind velocity. The dependence of the dampings of two modes
on the structure properties is also investigated. Finally, an extrapolation is per-
formed, yielding an estimate for the velocity at which the wing becomes unstable,
i.e. the flutter velocity. It will be shown that even though the experimentally
obtained damping through system identification as function of the velocity do
not suggest any decrease of the damping for increasing velocity, a flutter speed
can still be obtained thanks to the use of the theoretically prescribed behaviour
of the aeroelastic system as function of the velocity. As this result still is based
on extrapolations based on a simplified model, it is prone to verification which
will be done through a validation procedure. Finally, a ready to be implemented
method is described in detail to determine error bounds on the extrapolations.

2
Chapter 2

Aeroelastic flutter: a theoretical


background

2.1 Qualitative description


An important type of aeroelastic problem that has been observed for a long
time in mechanical structures is flutter. This unstable oscillatory phenomenon
occurs in ranges of wind speeds which are limited by their so called critical
flutter- or flutter onset speeds. For a simple cantilever wing in a wind tunnel
setup, increasing the wind speed from initially wind-off conditions first increases
the structural damping which causes the free oscillatory motion after an initial
deflection to damp out gradually. This damping however decreases rapidly from
a certain velocity on until it reaches 0 at the flutter onset speed, at which the
oscillatory motion is self sustaining with a constant oscillation amplitude. Beyond
this limit, if the wing is deflected by even the smallest disturbance, an unstable
oscillation of great violence occurs, resulting in a complete structural failure of
the wing.
Note that the motion of such a fluttering wing has 2 harmonically oscillat-
ing components according to the degrees of freedom of the cantilever wing: a
torsional- and a flexural one. Indeed, an airfoil so constrained as to have only
a flexural degree of freedom does not flutter, whereas one with only a torsional
degree of freedom flutters only if the angle of attack is at the proximity of the
stalling angle[5].
This coupled motion is due to the structural and inertial coupling of the wing,
which is related to its flutter behaviour as will be shown experimentally along the
project(see section 2.4 and chapter 4). Now, let's take a look at the nature of
this coupling. An aircraft wing in general can be idealized as a beam, especially
in conceptual and preliminary design. Hence, the torsional and flexural deforma-

3
tions can be approximated by use of beam theory[1]. This deformation can be
described by considering a beam's elastic axis, which are the loci of the shear
centers of the cross sections along the span. The shear center of a cross section
is a point through which a shear force will produce a pure, torsion free bending.
However for a beam whose elastic axis does not coincide with its centroidal axis,
a shear force through the elastic axis does induce twist and it can be shown that
the oscillation of such a beam is always a combination of flexure and torsion[5].

As mentioned above, two types of couplings can be considered in this case:


inertial and structural couplings. Inertial coupling is directly related to the offset
between the centroidal- and elastic axis, whereas structural coupling manifests
itself as an alteration of the constitutive laws, i.e.,the relationship between cross-
sectional stress resultants and the generalized strains(see [1] and equations 2.9
and 2.13).

Now, aeroelastic oscillations are sustained by aerodynamic forces which are


induced by an airstream and the structure itself. For a given offset of the elastic
axis w.r.t. the centroidal axis, the frequency of oscillation and airstream ve-
locity can be such that the aeroelastic system becomes unstable. Equivalently,
the flutter speed depends on the structural and inertial couplings of the wing.
Note that the corresponding frequency of oscillation naturally converges to the
resonance frequency of the (mainly)torsional mode and thus it is this mode who
becomes unstable[5]. As a result, the damping corresponding to this mode is
expected to decrease from a certain velocity on. This will be shown with the
help of the extrapolations based on theoretical trends of system parameters as
function of the velocity(see section 2.3 and chapter 5). Furthermore, it follows
that flutter occurence depends on the aerodynamic properties of the wing, as
well as inertial and stifness properties so that those need to be adapted in order
to prevent this instability phenomenon for a given flight envelope in the design
phase of an aircraft wing.

Note that a cantilever wing in theory has an infinite number of degrees of


freedom, but as the elastic torsional and flexural deformations are much more
important than the e.g. in plane deformation, the deflection of the wing can
be accurately described by looking only at the flexural and torsional deflections,
hence a assuming a system with 2 degrees of freedom.

4
2.2 Mathematical description
This qualitative description can be translated into a mathematical model through
a white box approach, given a set of simplifications. The implementation of
such a model enables us to simulate resonance frequencies and dampings as
function of the airspeed, the flutter speed and its dependence on the elastic
axis location. This simulation can be compared to experimental results as such
making a comparison between white box and grey box models. The purpose of
the thesis however is to construct a grey box model, which is one that is shaped
from the fusion of vibrational experimental data and a mathematical model as
the one which will be described in this paragraph.
The purpose of this paragraph is to present a ready-made procedure to imple-
ment a white box model into a simulation. Furthermore, it provides a theoretical
aeroelastic system which can be integrated in the construction of a black box
model, as described in section 2.3.

The aeroelastic model can be divided into a structural- and an aerodynamic


part. In order for this white box model to match our windtunnel model as close
as possible, a 3D model is built using the so called Strip Theory [2]. The struc-
tural part is modelled using the so called Hancock model[2]. Strip theory breaks
the wing into spanwise small strips. One can intuitively deduce that the solution
would depend on the thickness of these strips. The instantaneous aerodynamic
lift and moment acting on each strip are given by the 2D sectional lift and mo-
ment given by the 2D airfoil aerodynamics. Furthermore, this model assumes
the absence of 3D effects at the wing tip, and an infinite aspect ratio.

Now, consider a rigid flat plate of span s, chord c and thickness t suspended
by two torsional springs, one in roll(characterized by its stiffness Kγ ) and one
in pitch(characterized by its stiffness Kθ ), located at the flexural axis of the flat
plate(at x = xf ). Note that the pitch angle is assumed to be constant along the
span, which isn't the case in practice. The reason for this is the fact that the
wing is assumed to be rigid, therefore allowing only rigid body motion(pitch and
roll) but not bending or torsion.

5
Figure 2.1: Hancock model[2]

The equations of motion are derived using energy considerations. The wing
has 2 degrees of freedom, roll(γ) and pitch(θ). The kinetic energy dT of a small
mass element dm of the wing is then given by:
1 1
dT = ż 2 dm = dm(y γ̇ + (x − xf )θ̇)2 (2.1)
2 2
Integration over the entire plate gives its total kinetic energy:
m 2 2
T = (2s γ̇ + 3s(c − 2xf )γ̇)θ̇) + 2(c2 − 3xf c + 3x2f )θ̇2 ) (2.2)
12
The total potential energy V of the wing is simply:
1 1
V = Kγ γ 2 + Kθ θ 2 (2.3)
2 2
This yields the following structural equations of motion[2]:
! ! ! ! !
Iγ Iγθ γ̈ Kγ 0 γ M1
+ = (2.4)
Iγθ Iθ θ̈ 0 Kθ θ M2

Where
Z s
M1 = − yl(y)dy (2.5)
Z s0
M2 = − mxf (y)dy (2.6)
0

6
ms2
Iγ =
3
m(c − 2xf )s
Iγθ =
4
m(c − 3xf c + 3x2f )
2
Iθ =
3
It can be noticed here that we assume the system to be inertially coupled but not
structurally coupled since the cross terms are assumed to be zero in the above
stiffness matrix. The technique of the strip theory consists of applying 2D airfoil
aerodynamics to infinitesimal strips of the Hancock model, yielding aerodynamic
lift l(y) and moment mxf distributions depending on the spanwise position of
the wing and then carrying out the integrations in expressions 2.5 and 2.6, which
are the total moments around the flexural axis and the y = 0 axis.

As the flow around an aircraft wing and in the windtunnel is in general


unsteady and an oscillatory motion will be imposed on the wing in the windtunnel,
Theodorsen's unsteady theory [5] will be used to describe the aerodynamics at
the wing sections[2]:

3
l = ρπU cC(k)(U θ + y γ̇ + ( c − xf )θ̇)
4
c
+ ρπb2 θ̇ + ρπb2 (yγ̈ − (xf − )θ̈) (2.7)
2

3
mxf = ρπU ec2 C(k)(U θ + y γ̇ + ( c − xf )θ̇)
4
3 c c b4
− ( c − xf )ρπb2 θ̇ + ρπb2 (xf − )(yγ̈ − (xf − )θ̈) − ρπ θ̈ (2.8)
4 2 2 8
Where
0.165 0.335
C(k) = 1 − 0.0455 −
1 − k j 1 − 0.3 k
j
ωc
k= = reduced f requency
U
ec = of f set centroidal − elastic axis
c
b=
2
Now, this model with its simplifications can be representative for a windtunnel
model, but in practice equivalent stiffnesses Kγ and Kθ need to be found for
the torsional springs, given the material properties of the wing. To determine

7
these which correspond respectively with a bending stiffness EI and a torsional
stiffness GJ, local stiffnesses at the wing tip will be defined using statics and
beam theory. Next, the torques applied at the tip needed to overcome the elastic
forces due to the derived stiffnesses in order to bring about a rolling/pitching
motion will yield the stiffnesses Kγ and Kθ .

Consider a cantilever beam with length L on which a load P is applied at a


distance L1 from the tip. Before deriving an expression which relates the load P
to the deflection δ(x) as function of the spanwise position x and the spanwise lo-
cation of the applied load L1 , notice that we assumed the Hancock model to have
structurally uncoupled torsional and bending modes, which implies that bending
due to a shear force through the flexural axis does not cause a restoring torsional
force to bring back the beam to its equilibrium position. As a consequence, the
determined stiffnesses in the following represent respectively purely bending and
torsional elastic resistances. However though, it musn’t be forgotten that, as
the beam is inertially coupled, a force through the flexural axis will require to
overcome translational inertia as well as rotational inertia as the beam will twist
as a consequence of that force.

From statics, it can readily be found that the bending moment My around
the neutral axis is:

My (x) = P (x − L1 ) L1 ≤ x ≤ L
My (x) = 0 0 ≤ x ≤ L1

(b) Bending moment as function of longi-


(a) Cantilever beam with applied load P
tudinal(spanwise) coordinate x

Figure 2.2: Bending moment of a cantilever beam resulting from applied load P

8
Let v = x − L1 , then if x ∈ [L1 , L] ⇒ v ∈ [0, L − L1 ] and My = vP .
Substituting in the found expression for My the constitutive law

My (v) d2 δ(v)
= (2.9)
EIy dv 2
and integrating twice along the v coordinate yields, after substituting v = x − L1
(x − L1 )3 P
δ(x) = +C
6EIy
For the determination of C, the boundary conditions are expressed:
(x − L1 )3 P
x = L → δ = 0 ⇒ δ(L) = 0 = +C
6EIy
(L − L1 )3 P
⇔C=−
EIy

−(L − L1 )3 + (x − L1 )3 1
⇒ δ(x) = [ ]P = P (2.10)
6EIy Kδ
Where Kδ is the local stiffness as function of x and for the load P being at
x = L1 . As mentioned earlier, the local stiffness at the wing tip for a load
applied there will be used, which is equal to:
6EIy
Kδ = (2.11)
L3
Now, consider a load P acting at the wing tip at the flexural axis trying to cause
a small rolling motion. The vertical displacement of the Hancock wing at the tip
can then be approximated by δ ≈ Lγ, yielding

⇒ Kδ δ = Kδ Lγ

Multiplying by L gives the torque resisted by the torsional spring with stiffness
Kγ :
LP = Tγ = Kδ L2 γ = Kγ γ

6EIy
⇒ Kγ = (2.12)
L
Recall that the stiffness Kδ depends on the load location. This means that, as
according to strip theory, a lift distribution is obtained, an error in the estimation

9
of the stiffness Kγ is introduced by assuming a point load at the tip. Also, the
stiffness Kδ is a function of where we are considering a deflection δ. Hence, we
can expect as a consequence of these simplification a modelling error using a
Hancock model.

(a) Cantilever beam with applied load tor- (b) Torsional moment as function of longi-
sion T tudinal(spanwise) coordinate x

Figure 2.3: Torsional moment of a cantilever beam

For the ’pitching’ stiffness Kθ an analogue treatment as above yields, now


using the constitutive law
dθ(x) T (x)
= (2.13)
dx GJ
and applying a torque T at a position x = L1 of the beam:
L−x 1
θ(x) = = T (2.14)
GJ Kθ
This expression is independent of where the torque is applied. We then get, for
the torque being applied at x = 0:
L 1
θ(0) = T =
GJ Kθ

GJ
⇒ Kθ = (2.15)
L
Note that this expression, in contrast with the bending case, the stiffness Kθ is
directly obtained. Here also, in reality the stiffness changes as function of where
the deflection θ is considered, as θ should vary along the wing span. But as the
wing is assumed to be rigid, the stiffness is considered constant along the span.

10
The value Iy can been calculated using Autodesk Inventor®, where the wind-
tunnel model can been drawned. E and G are material dependent. The material
used in the experimental setup is Styrofoam FLOORMATE™200 SL-X (see ap-
pendix). For the value of J an elliptic geometry of the cross section is assumed
as the formula is then readily available, with the major axis being equal to the
chord length and the minor axis equal to the maximum camber:
Aa2 b2
J= (2.16)
4(a2 + b2 )
Where
ab
A = Area = π
4
a = major axis
b = minor axis
At this point, we managed to build up a simplified white box model. To calculate
resonance frequencies and dampings, the so called p-k method can be used,
which is the most popular technique for obtaining aeroelastic solutions from
Theodorsen-type aeroelastic systems. The advantage of this method is that it
quickly converges to the correct eigenvalue[2]. Following explanation is based
on the descriptions found in [1] and [2]. The p-k method uses the structural
equations of motion in the standard form, i.e. without assuming a solution for
γ and θ, coupled with Theodorsen's aerodynamic forces, whereby an oscillatory
motion is assumed. This results in Time-Frequency domain equations of motion,
where the coefficients depend on the frequency. The general form can be written
as:
1
IS q̈ + KS q − ρU 2 Q(k)q = 0 (2.17)
2
Where
!
γ
q=
θ
and the subscript S stands for structural and k is the reduced frequency. Solving
a system of time-frequency domain equations is a nonlinear eigenvalue problem.
Hereby, the equations of motion are rewritten by defining:
d
p= ⇒ q̈ = p2 q (2.18)
dt
The equations of motion then become:
1
(pIS + KS − ρU 2 Q(k))q = 0 (2.19)
2
11
Which is clearly an eigenvalue problem, with p2 being an eigenvalue of the matrix
−IS −1 (KS − 12 ρU 2 Q(k)). For a nontrivial solution we have following condition:
1
|p2 I2 + IS −1 (KS − ρU 2 Q(k))| = 0 (2.20)
2
Since this is one equation with two unknowns, we need a second condition. For
this, the imaginary part of the eigenvalue is imposed to be the frequency:
U
Im(p) = k (2.21)
b
Now, the problem consists of finding an eigenvalue p and a frequency k that sat-
isfy both equations 2.20 and 2.21, which requires an iterative procedure. Thereto,
a desired airspeed is chosen, a value for k is assumed and values for p are calcu-
lated from 2.20. As starting value for k, one of the following uncoupled resonance
frequencies can be used:
s

ωγ = (2.22)

s

ωθ = (2.23)

Indeed, it can be shown that there are two modes with the most significant de-
flections of which the first is mainly flexural and the second is mainly torsional[5].
The calculated values for p are sorted in ascending order of imaginary part. The
first eigenvalue corresponds to the bending mode(here γ) and then second to
the torsional mode(here θ). Next, the imaginary part corresponding to the mode
corresponding to the chosen starting value for the uncoupled resonance frequency
is taken as new value for k. The iterations are repeated until the found imaginary
part of p equals the input value k. This is called frequency matching. At that
point, the iterative procedure is repeated, but with a starting value corresponding
to another mode. If one wants to calculate resonance frequencies for a whole
speed interval, the whole iterative procedure is simply repeated for different speed
values within the desired range. The damping ratios are found as the ratio of
the real and imaginary parts of the eigenvalues:
Re(p)
ξ= (2.24)
Im(p)
In general, the Q(k) matrix is obtained using a panel method-based aeroelastic
model, which is constructed by means of commercial packages such as NAS-
TRAN. For a chosen set of k values, the corresponding Q(k) matrices are re-
turned. The values of Q(k) at intermediate k values are obtained by interpola-
tion. In our case, a more simplified aeroelastic model has been assumed, i.e. a

12
Hancock model for the structural part and strip theory has been applied for the
aerodynamic part of the model. In this case, the Q(k) matrix can be obtained
analytically. Performing the integrations (2.5) and! (2.6), letting θ = θ0 ejωt ,
M1
γ = γ0 ejωt , decomposing the obtained vector into a matrix-vector prod-
M2
!
γ
uct M · , isolating − 21 ρU 2 from M and substituting ω = k Ub in M yields
θ
Q(k) for our Hancock model:
3 3 2 !
jkπcC(k) s3b − 2πk 2 s3 πeC(k)s2 + πcC(k)jk( 34 c − xf ) sb + πjk Ub s2 + πk 2 (xf − 2c )s2
Q(k) = 2 2
πec2 C(k)jk sb − πk 2 (xf − 2c ) πk 2 (xf − 2 2c )2 + πec2 C(k)( 34 c − xf )jk sb − 2( 34 c − xf )πjk Ub s + 2sπk 2 (xf − 2c )2 + π b4 k 2 s
(2.25)

2.3 Grey box modelling: transfer function de-


nominator polynomial- and extrapolation or-
der determination
In the sequel of this thesis, a study of the aeroelastic system in the Laplace
domain is performed. This involves constructing a transfer function in the fre-
quency domain based on in- and output experiments. If this transfer function is
obtained solely from in- and output experiments, the modelling is then a pure
black box approach. However, here we will integrate a white box model into the
construction of the transfer function, in this way obtaining a grey box model.
Note that the consequence of this fusion is that the form of the transfer
function is defined, which will be a rational for and that the order of the cor-
responding denominator polynomial will be fixed. This indicates the number of
physical poles(see also chapter 5 section 5.2) and determines the size of the trans-
fer matrix. Also, the velocity dependence of the coefficients of the denominator
polynomial will be derived.
The latter consequence will provide a theoretical basis for the extrapolation
of these coefficients to higher velocities through imposed extrapolation orders.
As the zeros of this polynomial correspond to the poles of the aeroelastic system,
corresponding to its modes, the extrapolation will yield extrapolated dampings
and resonance frequencies. This extrapolation is performed to give the possibility
to predict a flutter speed without inflicting any damage to the wing.
The extrapolated dampings and resonance frequencies towards higher veloci-
ties can then be validated by extrapolating the dampings and resonance frequen-
cies given a part of the experimentally obtained poles in a velocity range within
the one for which all the poles were obtained experimentally. The extrapolated
dampings and resonance frequencies at the velocities where also experimentally

13
obtained dampings and resonance frequencies are available are compared to these
experimentally obtained dampings and resonance frequencies by looking at the
uncertainties of the latter and evaluating to which extent the extrapolated damp-
ings and resonance frequencies are outside of those uncertainty bounds.
Finally, a method will be described in chapter 5 to estimate the uncertainties
of the extrapolated dampings given the uncertainties of the experimentally ob-
tained dampings.

In the sequel of this chapter, the aeroelastic system defined in previous para-
graph will be further
! elaborated and converted into a transfer function. For this,
M1
the vector is decomposed into the sum of matrix-vector products:
M2
! ! ! !
M1 γ̈ γ̇ γ
= IA + CA · U + KA · U 2 (2.26)
M2 θ̈ θ̇ θ

Where we will call Ia the aerodynamic inertial matrix, Ca the aerodynamic damp-
ing matrix and Ka the aerodynamic stiffness matrix. Note that the elements of
Ca are a lineair function of the airspeed and the elements of Ka a function of
the velocity squared, hence they are isolated to express the velocity dependence
of the system explicitly. Now, two moments are applied on the Hancock model
in order to cause a forced motion, which also will be the case in the experimental
setup. Indeed, an oscillating pitching moment will be applied, hence inducing a
forced pitching motion θ. The aeroelastic system 2.4 now becomes:

IS q̈ + KS q − IA q̈ − U · CA q̇ − U 2 · KA q = Me

⇔ Iq̈ + C(U )q̇ + K(U 2 )q = Me (2.27)


Where

I = IS − IA
C(U ) = −U · CA
K(U ) = KS − U 2 · KA
2
!
Mγ , e
Me = Externally applied moment
Mθ , e

Transforming the equations of motion to the Laplace domain gives:

[Is2 + C(U )s + K(U 2 )]q(s) = Me (s) (2.28)

14
Where q(s) and Me (s) are the Laplace transforms of respectively the position
vector q and the externally applied moment Me and s the Laplace variable.

The inverse of the 2 × 2 matrix on the left hand side results in the Transfer
matrix, transferring input forces into output responses of the resulting motions:
adj(Is2 + C(U )s + K(U 2 ))
T(s, U ) = (2.29)
det(Is2 + C(U )s + K(U 2 ))
Note that the denominator of this expression is a scalar expression, hence being
a common denominator for all elements of the obtained matrix. As mentioned
above, our aim is to find the dependence of the poles of the obtained transfer
functions on the airspeed, which correpond to the zeros of the common denom-
inator. Those zeros depend on the velocity through the denominator polynomial
coefficients as can be easily proven by elaborating the above determinant:

det(Is2 + C(U )s + K(U 2 )) =


1 2 3 2
α0,0 s4 + α1,p U p s3 + α2,p U p s2 + α3,p U p s + α4,p U p (2.30)
X X X X

p=0 p=0 p=0 p=0

As stated above, the common denominator polynomial coefficients are polyno-


mials in U having each specific polynomial orders. This information can be very
useful as when conducting wind tunnel experiments, one doesn’t necessarily want
to increase the wind speed until it causes flutter to occur in order to predict the
flutter speed. In this case, the flutter speed can be predicted by means of ex-
trapolation, using these relations between the coefficients and the wind speed as
theoretical basis leading to reliable extrapolation results. This is dealt with in
chapter 5 of this thesis.

2.4 List of simplifications and flutter parameters


In this last section we will dwell on a bit on the parameter that influences flutter
behaviour. This parameter, which is the flexural axis location, has already been
mentioned in the previous paragraph. Here a link will be made with the experi-
mental setup in a way that allows us to modify the location of the flexural axis
experimentally.
Then, a summary of the simplifications is given, regrouping the origins of
modelling errors which would explain the difference between results of a white
box and a grey box model. Furthermore, these modelling errors quantify the
accuracy of the grey model with respect to reality.

15
The location of the flexural axis can be modified in the simulation described
in section 2.2 directly, whereas if we want to modify this location experimen-
tally, we need to make an appeal on the knowledge of the relation between the
elastic/flexural axis location and inertial properties which can be changed in the
experimental setup. In equation 2.4 it can be seen that Iθ and Iγθ are related to
the flexural axis location xf . Increasing Iθ without increasing Iγθ doesn't result
in an increase in inertial coupling, as this coupling manifests itself in the cross
terms of the rotational inertia matrix. To affect the inertial coupling is thus
necessary to introduce assymetry, which will be done in the experimental setup
by installing an extra mass at one extremity of the wooden beam(see 4.3).
In the following a list is made of the assumptions/simplifications in the con-
struction of the aeroelastic models:

• Absence of 3D effects at the wing tip

• Infinite aspect ratio

• Constant pitch angle along the span

• Only inertial and aerodynamic coupling

• Torsional constant J assumed for elliptic beam

16
Chapter 3

Grey box modelling: experimental


setup and methodology

3.1 Experimental setup: the AATB


The experimental data needed for the development of models through system
identification is provided by the Active Aeroelastic Test Bench(AATB), which
was designed for the study of low subsonic unsteady aerodynamics and aeroe-
lasticity[4]. It is equipped with 2 linear motors allowing forced motion of a wing
in pitch and plunge degrees of freedom through a pitch beam. Support in the
4 remaining degrees of freedom is provided. The setup is equipped with load
cells, glass scale encoders embedded in the linear actuators and accelerometers
mounted on the wing, allowing to respectively record the forces acting on it, the
position of the linear actuators and the acceleration that the wing undergoes,
quantifying the aeroelastic response. Furthermore, a wooden beam was fixed to
the wing in order to be able to attach a weight to it to shift the flexural axis which
on his turn lowers the flutter onset speed(see chapter 2, section 2.4). Photo’s of
the actuating system and the wing installed in the wind tunnel can be visualized
in figures 3.1 and 3.2.
Moreover, the wind tunnel is equipped with a blower located upstream of
the setup which can be set according to a desired wind speed. The maximum
achievable wind speed is 20m/s.

17
Figure 3.1: Wing with accelerometers installed at wing tip

Figure 3.2: Linear actuators

Applying identical signals to both actuators leads to plunge movement whereas

18
applying different signals leads to the rotation of the pitch beam, providing pitch
motion of the wing. The pitch angle is obtained from the position of the actua-
tors using equation 3.1 which is based on figure 3.3. Note that this linear relation
is an approximation assuming small pitch angles as the actuator displacements
are not vertical as modelled in figure 3.3 on the right but instead the actuators
pivot(figure 3.3 left) as otherwise the pitch beam is required to become longer(or
shorter).
∆X1 − ∆X2
θP itch ≈ (3.1)
l

Figure 3.3: Rotating actuator(left) and simplified representation of the actuator


displacements(right)[4]

Where l is the distance between the actuators, amounting to 250mm and


∆X1 and ∆X2 are the respective actuator positions obtained by the encoders.
In the context of this project, only the response due to an applied pitch motion to
the wing is studied. This is sufficient as the pitch and plunge modes are inertially
coupled as assumed in chapter 2. Hence, excitation in the pitch direction brings
forth a motion in the plunge direction and vice versa.

The pitch angle obtained with equation 3.1 from the position readouts of
the linear actuators represents the input, which deviates from the real input as it
includes nonlinearities due to the kinematics of the AATB. As a matter of fact,
the relation between the pitch angle and the commanded actuator position is
non linear whereas the above approximation implies that the relation is linear.
As a consequence, a distorted sine will enter the aeroelastic system when anti-
symmetric pure sines are applied at the input of the actuators. Antisymmetric
here means that the signals are identical except for their sign, which are opposite
with respect to each other. Hence, these nonlinearities will be present in the
output. The derivation of the linear as well as the non linear relation between
the pitch angle and commanded actuator positions can be found in[4].

19
One would expect a flat input spectrum as is the case when working with
multisine excitations but here the input spectrum, as can be seen in figure 3.4.
is a decreasing function of the frequency. The reason for this is that in practice
an airplane wing oscillates at high amplitudes and low frequencies and at low
amplitudes and high frequencies. Furthermore, the evolution of the amplitude
spectrum of the actuator position is such that the amplitude is inversely propor-
tional to the frequency squared, as this results in a constant acceleration of the
actuators as function of the frequency. Indeed, deriving twice the position in the
time domain with respect to the time variable yields in the frequency domain
the spectrum of the position times the frequency squared, resulting in a constant
amlitude spectrum for the acceleration:

F {f 0 (t)} = iωF (ω) (3.2)

Figure 3.4: Input spectrum: inversely proportional to the frequency. The distor-
tion in comparison with the corresponding smooth function is due to the dynamics
of the actuators

A constant acceleration level as function of the frequency is preferred as we


would like to study the vibrational behaviour of the cantilever wing and vibrations
are expressed in terms of acceleration.

The setup is equipped with 2 accelerometers from which the measured sig-
nals(which represent an oscillating acceleration) represent the output. 1 is
mounted at the trailing edge side extremity of the wooden beam in order to
record a maximal deflection in the pitch direction and 1 is mounted at the level
of the pitch axis in order to record a pure plunge deflection. As both modes are
interconnected, i.e. measuring at the trailing edge gives also information about
the flexural mode, the information provided by the accelerometer at the trailing

20
edge should be sufficient to describe their velocity dependent behaviour. How-
ever processing both output signals allows more accurate results as it provides
additional information about the system under study. Indeed, the identifica-
tion performed further in this paper assumes a common denominator for the
modeled elements of the F RM . Hence, for the same denominator, 2 datasets
are available, which lowers the variability of the coefficients of the denominator
polynomial coefficients(see also section 4.2). Furthermore, the wooden beam is
installed at the tip of the wing, for a maximal deflection. Hence, a single input
multiple output system is defined, which is illustrated in Figure 3.5.

Figure 3.5: Schematic of how the commanded position is transferred into mea-
sured acceleration

3.2 System identification for the characteriza-


tion of flutter behaviour
3.2.1 Signal design: odd multisine excitation
The signal that is used as input is an odd multisine with random odd frequencies
not being excited to be able to estimate nonlinearities. Indeed, leaving random
odd frequency lines unexcited allows to make an estimation of odd nonlinearities
given only 1 realization as that is the only contribution that should be present
at the odd frequency lines(note that this is not true as we have nonlinearities
in the input, which is dealt with in the following section). The even frequency
lines of the output only contain even nonlinearities. Hence, a distinction could be
made between odd- and even nonlinearities given only 1 realisation. Furthermore,
the reason why the unexcited frequency lines were chosen to be random is to
make sure that we do not blind ourselves for nonlinearities by e.g. systematically
choosing certain odd frequency lines to be excited where the nonlinearities could
occur and where they would not occur in the non excited odd lines, leading to
the false conclusion that there are no odd nonlinearities.

21
3.2.2 Non parametric identification: the local polynomial
method
In this section, the local polynomial method is described briefly and is based
on the description given in [8]. This method describes a non-parametric iden-
tification method which estimates an F RF along with its covariance matrix.
Its advantage over classical non-parametric methods is that it suppresses noise
leakage(transients) when present resulting in a smaller variance of the estimated
F RF [8]. The non-parametric identification step is beneficial as it allows the
evaluation of the quality of the measurements and the complexity of the prob-
lem can be revealed. Furthermore, the variance can be used in the parametric
identification step as explained in the next section.

In the local polynomial method, the energy at the non excited DFT lines
kP + m, k = 0, 1, ..., N/2 − 1 and m = 1, 2, ..., P − 1 of a random phase
multisine of P periods is used for the estimation of the noise covariances and noise
transients at the excited DFT lines by approximating the transient contribution
in the neighbouring non excited frequencies by performing a Taylor expansion at
the DF T frequencies:
R
tr (k)mr (3.3)
X
THZ (ΩkP +m ) = THZ (ΩkP ) +
r=1

Here, k is the bin number, P the number of periods, m represents the location
of the non excited DFT frequency in the neighbourhood of the excited DFT
frequency kP and N is the number of sample points per period. k extends until
N/2 − 1(N/2 corresponding to the Nyquist frequency), hence considering the
first half of the discrete Fourier spectrum. Ω is the frequency corresponding to
bin kP +m and kP . Note that we neglected the remainder as also done in [8]. Z
is the vector that is obtained by putting the input-output signals on top of each
other into a 2x1 vector Z(k) = [Y (k)U (k)]T , T representing the transposing
operator and is in relation with the transient contribution as follows:

Z(kP + m) = V (kP + m) + THZ (ΩkP +m ) (3.4)

Where V (kP + m) is the noise contribution.


For each excited frequency, a system of 2n equations of order R is then
constructed with the coefficients of the Taylor expansion and the transient term
at the excited frequency being the R + 1 unknown parameters to be determined,
hence linking the neighbouring(or local) non excited frequencies through these
coefficients. n denotes the number of non excited DFT lines belonging to the
set N\{kP |k ∈ N}.

22
Afterwards, the estimated transient contribution at each non excited fre-
quency(in the neighbourhood of the excited frequency) are substracted from the
spectral data, resulting in the so called residuals of the least squares fit V̂n , yield-
ing an estimation for the noise Vn at those non excited frequencies. Finally the
noise covariance matrix at the excited frequencies is then obtained as:
1
ĈV (kP ) = V̂n V̂nH (3.5)
q
Where q = 2n − (R + 1) and H is the hermitian(complex conjugate) trans-
posing operator.

The transient contribution at the excited frequencies are also estimated when
creating the local polynomial and thus can be substracted from the energy on
the excited frequencies yielding the input-output sample means over the periods:

Ẑpoly (kP ) = Z(kP ) − T̂HZ (ΩkP ) (3.6)


Now, a distinction is made between the fast local polynomial method and the
robust local polynomial method. The fast version handles one realization of a
random phase multisine, i.e. it analyzes the noise as described above and it
estimates noise and the nonlinear distortions using the level of non excited odd
harmonics. In the following, the estimation of the level of the nonlinearities by
the fast LP M is explained.

First, it must be noticed that, because of nonlinearities present in the in-


put(energy present on the non excited lines), the linear contribution response of
the system on these non excited lines need to be suppressed in order to be able
to make an estimation of the nonlinearities due to the aeroelastic system. This
can be done by estimating this level of the linear contribution of the response
of the system on the excitation at the non excited frequencies, after which they
are substracted from the energy on the non excited frequencies of the response,
yielding only the nonlinear contribution of the system’s response on the excited
input.
The response Ytot (fnexc ) at the non excited frequencies fnexc can be divided
into a contribution of the linear response on the non excited input YBLA (fnexc ),
the nonlinear response on the excited input YSexc (fnexc ) and the nonlinear re-
sponse on the non excited input YSfnexc (fnexc ):

Ytot (fnexc ) = YBLA (fnexc ) + YSexc (fnexc ) + YSfnexc (fnexc ) (3.7)

In the following, the contribution YBLA (fnexc ) will be estimated. Assuming that
the response divided by the input at the excited frequencies equals the best

23
linear approximation(indeed, as the output also includes a significant non linear
contribution of the response on the input at the other excited DFT lines yet a non
significant nonlinear contribution of the response on the non excited contribution
which can be neglected):

YBLA (fexc ) + YSexc (fexc ) + YSnexc (fexc )


≈ GBLA (fexc ) (3.8)
U (fexc )
and interpolating the results of those divisions yield an estimation of the BLA
at the non excited frequencies:

Interpolate(GBLA (fexc )) −→ GBLA (fnexc ) (3.9)

The obtained function can then be multiplied with the energy at the non excited
DFT lines of the input, yielding the linear contribution of the response at these
non excited frequencies:

YBLA (fnexc ) = GBLA (fnexc ) · U (fnexc ) (3.10)

Next, this linear contribution is substracted from the energy at the non excited
frequencies(equation 3.7). As the non linear contribution of the response at the
non excited frequencies, on the non excited frequencies of the input, can be
neglected, the result of the substraction can be considered as the level of the
stochastic non linearities.

Ytot (fnexc ) − YBLA (fnexc ) = YSexc (fnexc ) + YSfnexc (fnexc ) ≈ YSexc (fnexc ) (3.11)

Now that we have a pure nonlinear contribution of the system’s response on


the excited input YSexc (fnexc ), this information can be used as intended to do
when designing the odd multisine with random non excited odd lines(see subsec-
tion 3.2.1). On the odd lines, only odd nonlinearities are present. As we already
have the level of the nonlinearities on the non excited odd frequency lines, it
remains to determine the level of the nonlinearities at the excited odd frequency
lines. The idea to this aim is to use the information on the non excited odd fre-
quency lines by interpolating this energy(square of the amplitude of the output)
to the excited frequency lines. It can be shown that this interpolation asymp-
totically(for the number of excited frequency approaching infinity) matches in
expected value the variance of the BLA, ˆ which is obtained by dividing output
by input for this realization[3]. Subsequently, this variance is an estimation for
the level of stochastic odd nonlinearities at these odd frequency lines.

The robust version starts from at least two realizations, then analyses the
noise as described above for each realization after which the obtained noise

24
(co-)variances are averaged over the realizations, yielding an improved esti-
mate of the noise (co-)variance. Then, the sum of the noise (co-)variance and
the (co-)variance of the nonlinear distortions(the total variance) is estimated
by averaging over the realizations and neighbouring excited frequencies. This
[m] [m]
means that each ’squared’ error rX (k + ki ) · rL (k + ki ) corresponding to re-
alisation m is averaged over neighbouring excited frequencies k + ki , where
L, X ∈ {U (Input), Y (Output)}. This leads to following algorithm for the cal-
culation of the total variance:
M nR [m] [m]
rX (k + ki )rL (k + ki )
2
(3.12)
X X
σ̂X R LR ,poly
(kP ) =
m=1 i=−nR (M − 1)(2nR + 1)

[m] [m]
rX (k) = X̂R,poly (kP ) − X̂R,poly (kP ) (3.13)

M
1 X [m]
X̂R,poly (kP ) = X̂R,poly (kP ) (3.14)
M m=1
[m]
Where X̂R,poly (kP ) is a sample at frequency bin kP from the in- or output from
realisation m, k + nR is the bin number of the endmost neighbouring excited
frequency and the R stands for ’Robust’.

The variance of the nonlinearities is then obtained by substracting the noise


(co-)variance from the total (co-)variance calculated above.

The latter approach has the advantage over the former one, in addition to
producing a better estimate of the noise (co-)variance, that the resulting (co-
)variance of the nonlinear distortions are a better estimate(have a smaller vari-
ance) than the (co-)variance estimated in the fast method[3]. Furthermore,
according to the square root law, the uncertainty on the obtained BLA, which
is a result of an averaging over the realizations and neighbouring frequencies, is
divided by the number of realizations compared to the uncertainty on the ob-
tained BLA in the fast method. This means that the variance of the obtained
BLA is the result of the above calculated total variance divided by the number
of realizations.

Finally, the described method can be applied on data from SISO experi-
ments, whereas we are dealing with a multivariate case here. The robust method
then requires a reference signal, which was available in our case(see section 3.1).
This method is similar to the one above and the differences are explained more
in detail in[10].

25
3.2.3 Parametric identification: the Maximum Likelihood
method
The second step in the identification procedure consists of a parametric estima-
tion of the system. The next explanation is based on [7].

The Errors-in-Variables framework is used, in which the stochastic char-


acter is expressed of both the input and the output through a cost function
VF (θ, Z)(where F is the number of excited frequency lines and θ a parameter
vector(containing Up and Yp as well which are the parametrized in- and output
as their true values are unknown)) which is to be minimized[7]:
F
!H " #−1 !
1 X Y (k) − Yp (k) σY2 σY2 U Y (k) − Yp (k)
VF (θ, Z) = (3.15)
F k=1 U (k) − Up (k) σU2 Y σU2 U (k) − Up (k)

implying that a direct division of the measured spectra, which can lead to a situa-
tion where in general the variance doesn’t exist[7], is avoided. This direct division
is what is performed in the intuitive approach to the parametric estimation, which
burdens the user to verify whether or not he can use it(for the formula express-
ing this approach see [7]). As using the Errors-in-Variables framework however
avoids an eventual division by zero, it leads to a robust identification method.
Note that the above expression implies that the in- and output are unknown,
which makes them also parameters to be estimated.

The specific method used to this aim is the Maximum Likelihood Method,
which incorporates the knowledge about the probability density function(pdf ) of
the noise NZ , satisfying:
Z = Z0 + NZ (3.16)
Where Z, Z0 and NZ are respectively the DF T of the data vector, the true data
vector and the noise vector, which are obtained by stacking the corresponding
input and the output into a 2F x1 vector as follows(applies also to Z0 , NZ and
later Zp(,0) ):
Z T = [Z T (1)Z T (2)...Z T (F )] (3.17)
Where
Z T (k) = [Y (k) U (k)] (3.18)
and F is the number of excited frequencies. Indeed, the method maximizes the
function which expresses how likely a certain set of parameters will generate a
given measurement Z. Then, let a set of parameters θ0 , Zp,0 be the true param-
eters. The likeliness that those parameters will be responsible for transferring

26
an input U into an output Y depends on the in- and output respective noise
distributions fnX , where X ∈ {U, Y }. Hence, we can write:

f (Z|θ0 , Zp,0 ) = fnZ (Z − Zp,0 ) = fNZ (NZ ) (3.19)

Note that then Zp,0 = Z0 , which is assumed to be deterministic. This function is


fully determined by the covariance matrix and mean of NZ . Indeed, if the noise is
Gaussian-(or normal-) and circular complex distributed and the true values Z0 (or
Zp,0 ) are deterministic, fnZ satisfies[7]:
1 H Γ−1 N
−NZ
fNZ (NZ ) = exp N Z
Z
(3.20)
π F det(ΓNZ )

Where ΓNZ = E[(NZ − µ) · (NZ − µ)H ] is the covariance matrix, µ being


the noise expected value and E the expected value operator. Circular complex
distributed implies that the noise NZ is zero mean(µ = 0) and CNZ = 0, where
CNZ = E[(NZ − µ) · (NZ − µ)T ] is the pseudo-covariance matrix. Uncorrelated
over the frequency implies that ΓNZ is a diagonal matrix. This follows directly
from the definitions of ΓNZ and CNZ . After some mathematical operations,
following form is readily obtained for the covariance matrix:

ΓNZ = diag(Cov(NZ (1)), Cov(NZ (2)), ..., Cov(NZ (F ))) (3.21)

Where
" #
σ 2 (k) σY2 U (k)
H
Cov(NZ (k)) = E[NZ (k) · NZ (k) ] = 2Y (3.22)
σU Y (k) σU2 (k)

Note that σU2 Y (k) = σ 2Y U (k).


Now, let Z be a given measurement and consider θ and Zp to be variable
parameters. The Maximum Likelihood function, which expresses the likeliness
that certain parameters θ, Zp will generate a measurement Z, becomes:
1 −(Z−Zp )H Γ−1
NZ (Z−Zp )
f (Z|θ, Zp ) = fnZ (Z − Zp ) = F
exp (3.23)
π det(ΓNZ )

Note that if ΓNZ is singular, then Γ−1


NZ and det(ΓNZ ) are replaced by respectively
ΓNZ (Moore-Penrose pseudoinverse) and the product of the non-zero eigenvalues
+

of ΓNZ [7]. In the sequel of this section, the use of this method is explained.

From the covariance matrices obtained from the the LP M method described
previously, the pdf of the frequency domain noise is constructed. Let that pdf
be nZ (k), where k is the bin number(k = 1, 2, ..., F ) and Z the corresponding

27
(by the LP M ) estimated in-and output. Let Zp be the parametrized in- and
output and θ the vector of model parameters(without the parametrized in- and
output), which are related with the equation errors defined as[7]:

e(Ωk , θ, Zp (k)) = A(Ωk , θ)Y (k) − B(Ωk , θ)U (k) (3.24)

This equation error is defined as the difference between the left- and righthand
sides of following assumed transfer function model:

Y (k) = G(Ωk , θ)U (k) (3.25)

Where U (k) is the DF T spectrum of a band limited non noisy periodic signal with
harmonically related frequencies hf0 , h ∈ H ⊂ N, and period T0 = f10 = NFpperS
,
where FS is the sampling frequency and Npper the number of sample points
per period of the signal. Y (k) is then the DF T spectrum of the steady state
response that is obtained when observed during an integer number of periods
M T0 with M ∈ N at the excited frequency lines k = M h. The corresponding
model transfer function is then assumed to have a rational form[7]:
nb
b r Ωr
P
B(Ω, θ)
G(Ω, θ) = = Pnr=0 r
(3.26)
A(Ω, θ) r=0 ar Ω
a

Where Ω = s, the complex Laplace variable. Note that because of this definition,
orders have to be chosen for the numerator and denominator polynomials. The
procedure followed for this choice is explained in detail in section 4.2.1.
Under the assumption that nZ (k) circular complex normally distributed and
the true values Z0 (or Zp,0 ) are deterministic, the Maximum Likelihood cost func-
tion becomes, where the negative natural logarithm of the likelihood function
has been considered, the Lagrange multiplier method has been used and the
parametrized Zp has been eliminated[7]:
F
|e(Ωk , θ, Z(k))|2
(3.27)
X
VM L (θ, Z) =
k=1 σe2 (Ωk , θ)

Where σe2 (Ωk , θ) is the variance of the equation error[7]:

σe2 (Ωk , θ) = |A(Ωk , θ)|2 σY2 (k) − |B(Ωk , θ)|2 σU2 (k)
− 2Re(σY2 U (k)A(Ωk , θ)B(Ωk , θ)) (3.28)

Note that the negative logarithm implies that in order to maximize the likeli-
hood function, the obtained negative logarithm of it has to be minimized. The

28
Maximum Likelihood estimate θ̂M L (Z) is then the minimizer of 3.27 and the
estimated output and input by the Maximum Likelihood method is then[7]:

ŶM L (k) = Y (k) − (σY2 (k)A(Ωk , θ̂M L )−


e(Ωk , θ̂M L , Z(k))
σY2 U (k)B(Ωk , θ̂M L )) (3.29)
σe2 (Ωk , θ̂M L )

ÛM L (k) = U (k) − (σY2 U (k)A(Ωk , θ̂M L )−


e(Ωk , θ̂M L , Z(k))
σU2 (k)B(Ωk , θ̂M L )) (3.30)
σe2 (Ωk , θ̂M L )

Note that the obtained cost function is highly nonlinear in the parameters(present
in the denominator and squared in the numerator). Therefore, a numerical op-
timization algorithm is required to find the minimizer θ̂M L (Z), of which the
convergence strongly depend on the starting values for θ. In order to generate
staring values, the cost function is then modified such that advanced numeri-
cal techniques, such as singular value decomposition, can be used to find easily
its global minimum, which is used as starting value. The cost function into
which the one associated with the Maximum Likelihood is changed in our case,
is the one associated with the Generalized Total Least Squares method. The
found global minimum of this cost function is then used as starting value for the
cost function associated with the Bootstrapped Total Least Squares method, of
which the global minimum is then used as starting value for the Maximum Like-
lihood method. The reader is referred to [7] for a description of those methods.
The minimum of the final cost function is found using the Newton-Gauss or
Levenberg-Marquardt iteration schemes.

For the uncertainty bounds the reader is referred to[7]. Note that these uncer-
tainty bounds on the parameters have been passed on to the estimated damping
ratios and the resonance frequencies.

Note that the obtained errors for the dampings and the resonance frequen-
cies which are based on the covariances of the nonparametric estimate don’t take
into account the errors introduced between experimental sessions(see chapter 4:
change of inertial coupling), e.g. due to removing and reinstalling the wing in the
windtunnel and the accelerometers on the wing, and temperature and humidity
changes.

29
Chapter 4

Results

4.1 Introduction
Time invariant input- and output measurements were performed on the AATB
at wind speeds varying between approximately 10 m/s and 20 m/s and with dif-
ferent inertial couplings through the addition of masses of approximately 100g.
The increments of the velocities were performed in steps of 1 m/s, choosing a
step of 2m/s for the two first velocity increments for the obtained setup after
the last adding of a mass at the extremity of the wooden beam installed at the
tip of the wing. Note that the wind speed is a part of the system so that at each
velocity a new system is obtained with new properties and for which a new time
invariant identification is performed.

The additional mass was intended to bring the aeroelastic system closer to
instability as without additional mass, no decrease of the damping corresponding
to the torsional mode could be observed. This mass increases the asymmetry of
the wing with respect to the flexural axis, which increases the coupling between
flexural and torsional deflection as described in chapter 2. This should hence
result in the flutter speed becoming lower. As flutter occurence happens in a
quite agressive way[5], the velocity increments were first tuned as to avoid sudden
unstable behaviour of the wing, but after noticing visually that still no signifi-
cant change in torsional deflection showed up, the velocity increments could be
increased until 2 m/s from 10 m/s on. Then, from 14 m/s on the increments
were lowered so as to not take any risk. This was done in the last experiment
to find a balance between the time available to perform the experiment at once
and choosing a high frequency resolution for maximal accuracy of the obtained
linear system. The agressive behaviour of flutter was also the reason why the
adding of the mass happened in two steps. Because the experiments following

30
the first adaptation of the experimental setup didn’t lead to satisfactory results
as in the first set of experiments, which are a beginning decrease of the damping
corresponding to the torsional mode, a second mass was added.

The acquired data were processed in a second step in order to be able to


identify the evolution at different wind speeds of the damping of the system.

4.2 First series of experiments: setup without


additional mass
4.2.1 Identification
A random phase multisine of 8 consecutive periods sampled at a frequency of
3000 Hz and exciting only the odd frequencies from which, in each block of 6
subsequent odd harmonics, 2 random lines were not excited to serve as detection
lines of odd harmonic nonlinear contributions between approximately 0Hz and
50Hz was used as excitation signal. The RMS value of the excitation amounted
to 0.2◦ (recall that the input is a multisinusoidal pitch angle, which is expressed
in ◦ ). The frequency resolution was here chosen to be 0.067Hz, which resulted
in a measurement time of 15s per period. In order to reduce the uncertainty on
the BLA(as well as on the variance on the noise and nonlinearities), multiple
random phase realizations were performed. The number of those realizations
varied for the different wind speeds.
As the response of the system as a result of a forced pitch motion was stud-
ied, the 2 actuators got the same but opposite reference signals. Subsequently,
the position of the actuators determined the pitch angle as described in the para-
graph describing the experimental setup(see 3.1).

Important to mention is that the acquisition was capable of sampling at up to


1000Hz, which divided the number of data points by 3 and thus also the number
of data points per period. Hence, the frequency resolution stayed the same as
also the sampling frequency was divided by 3.
FSamp
FRes = (4.1)
Npper

Where Npper is the number of datapoints per period recorded by the acquisition.

Afterwards, the measurement data was provided for the identification which
consisted of a parametric- and a non-parametric step.

31
In the first one, the different realizations are passed to the ’RobustLocalPolyAnal’
MATLAB function which can be found in the frequency domain identification
toolbox written by Prof. Rik Pintelon. This step consists of making an F RF (or
F RM ) estimate by linking supposed transient errors on the non excited frequen-
cies in the neighbourhood of the excited frequencies of the input- and output
spectra and substracting them from the measured input- and output spectra.
Furthermore, the noise covariance is calculated based on the obtained transient
errors and the total variances is calculated by averaging over the realizations
and the neighbouring excited frequencies. This is explained in more detail in
chapter 3 section 3.2.2.
This identification step is necessary as it allows to assess the plant and mea-
surement quality which can be taken into account in the parametric identifica-
tion and can be used for the model validation. Moreover, qualitative information
about the complexity of the problem can be obtained, allowing to make an es-
timation of the model order in the parametric identification, which brings us to
the following step in the identification procedure[7].

The parametric identification step consisted of applying the Maximum Likeli-


hood estimator to the estimated F RM along with its covariance matrix. To this
aim, the order of the numerator and denominator of the elements of the to be
modelled F RM had to be assumed, as explained in section 3.2.3. This can be
done by visualizing the estimated F RM . When we look at one of the elements of
the estimated F RM , we find peaks, each of which corresponds to a mode of the
aeroelastic system. Indeed, a mechanical structure can be modelled by a system
of differential equations of which, after transformation to the Laplace domain,
the obtained characteristic polynomial after some mathematical operations rep-
resents the common denominator of the obtained F RM orTransfer Matrix (see
also chapter 2 section 2.3). Then, as for mechanical systems the damping can
be assumed to be small, we find that the zeros of the characteristic polynomial
of which the order is determined by the number of degrees of freedom that is
assumed for that system, are a set of DOF (=number of degrees of freedom)
complex conjugate pairs[6]. Hence, the characteristic polynomial has order 2
for a 1 DOF -mechanical system, order 4 for a 2 DOF -mechanical system, etc.
Those pole pairs manifest themselves as peaks in the elements of the estimated
F RM and as they each correspond to a mode of the system, the peaks also do.
As a consequence, counting the number of peaks in elements of the estimated
F RM yield an estimation for the number of modes, degrees of freedom and the
number of complex conjugate pole pairs which on its turn yields the order of
the characteristic polynomial to be chosen for the to be modelled F RM . As
for the numerator, they are polynomials which are the elements of the adjoint
matrix of the matrix corresponding to the system of differential equations. The

32
zeros of those polynomials correspond to downward peaks in the elements of
the estimated F RM . However for the sake of convenience, the order of the
numerator is chosen equal to the order of the denominator, as a wrong order for
the numerator doesn’t influence the dampings and resonance frequencies of the
system.
Note that when a linear time invariant system of multiple degrees of free-
dom is modelled, the size of the obtained F RM or Transfer Matrix through a
white box approach is DOF xDOF . As our system can be theoretically mod-
elled as a 2 DOF (bending and torsional mode) system, the resulting transfer
matrix is a 2x2 matrix(see also chapter 2 section 2.3). The estimated F RM
here has 2 elements because it is based on 1 input, 2 outputs measurements.
As the input consists in the excitation of the torsional mode and the outputs
are placed at the trailing edge and the flexural axis, the parametric identification
actually produces the first column of the F RM obtained in chapter 2 section 2.3.

As an illustration, the identification procedure at a wind velocity of 10m/s


is shown. The number of peaks in the first element of the estimated F RM
amounts to 4 so a denominator- and numerator order of 8 is chosen for the
construction of the transfer function. From here on, the orders of the numerator
and denominator will always be chosen to be equal when performing the para-
metric identification and we will speak about model order instead of numerator
and/or denominator order. Figure 4.1 illustrates the first element of the esti-
mated F RM , the corresponding standard deviation(which is the square root of
the total variance), the fitted model and the residuals.

Figure 4.1: F RF estimate, parametric model(8th order), total standard deviation


and residuals for system at 10m/s. The residuals are systematically higher than
the total standard deviations.

As can be noticed, the residuals are systematically higher than the stan-

33
darddeviations, which suggests to increase the model order as there has been
undermodelled. Indeed, if the residuals are scattered around the total standard-
deviations, it can be concluded that the model is a good fit. Also, the variability
has to be kept low. Hence, the best model order is the one for which a balance
is found between the mean(over the frequencies) difference between residual and
the standard deviations being the nearest to zero and for which the errors on the
dampings and resonance frequencies being the lowest. This is the criterion that
is used in this thesis. Furthermore, the first peak of the estimated F RF is not
considered by the parametric modelling to be a pole of the system, and therefore
does not fit the F RF there. As a consequence, no damping nor a resonance
frequency is defined at those frequencies. Note that the importance of having
a good fit lies in the fact that we need to be able to see the evolution of the
modes of the system at the different wind speeds.
Applying the above criterion for the above conditions of the windtunnel(without
added weight, wind speed of 10m/s), we get following fitted model of the twelfth
order corresponding to the first element of the estimated F RM :

Figure 4.2: F RF estimate, parametric model(12th order), total standarddevia-


tion and residuals for system at 10m/s. Now the residuals are scattered around
the standard deviations.

The residuals now clearly are scattered around the total standard deviations
and also the first peak of the estimated F RF is considered to be a pole of the
system and therefore the peak is also fitted by the model.

This identification procedure has been performed for each velocity. Hence
the evolution as function of the velocity of the damping ratios and resonance
frequencies corresponding to the torsional(first) and bending(second) modes(first
2 peaks) can be visualized in figures 4.3 and 4.4(again, only the first element
of the estimated F RM is illustrated). The two first peaks are attributed to the

34
flexural and torsional modes because these are the first two modes with the most
significant deformation amplitude compared to other modes in this frequency
band([5]).

Figure 4.3: Velocity dependence of the estimated F RF ’s corresponding to setup


without additional mass

(a) Velocity dependence of first mode (b) Velocity dependence of second mode

Figure 4.4: Velocity dependence of the first 2 peaks of the estimated F RF ’s


and corresponding modelled damping ratios and resonance frequencies.

When we take a look at the evolution of the second mode of the system at
different wind speeds, we notice that the damping ratios increase as the velocity
increases. Also, the resonance frequencies decrease. The increase in damping
however should theoretically be followed by a decrease[5] but since we were re-
stricted to a wind speed of 20m/s, this could in this case not be verified.

As is the case for the second mode, the first mode also increases as the ve-
locity increases whereas its resonance frequency increases with the exception of

35
the increase from 12m/s to 13m/s, at which the resonance frequency decreases.

For the sake of clarity, following figures show the evolution of the damping
ratios and the resonance frequencies graphically. In addition, a table summarizing
these values along with their uncertainty bounds is also shown in figure 4.1.

(a) velocity dependence of damping: flexu- (b) velocity dependence of damping: tor-
ral mode sional mode

(c) velocity dependence of resonance fre- (d) velocity dependence of resonance fre-
quency: flexural mode quency: torsional mode

Figure 4.5: Graphical representation of damping ratios and resonance frequencies


corresponding to torsional- and bending modes as function of wind speed

36
Mode Bending Torsion
Wind speed(m/s) fn (Hz) ∆fn (Hz) ξ(%) ∆ξ(%) fn (Hz) ∆fn (Hz) ξ(%) ∆ξ(%)
10 3.0871 0.0157 12.71 0.45 15.5380 0.0005 1.54 0.0031
11 3.8366 0.0136 12.87 0.38 15.5094 0.0006 1.60 0.0037
12 3.8659 0.0120 14.61 0.33 15.4941 0.0006 1.60 0.0039
13 3.8612 0.0116 16.01 0.32 15.4844 0.0008 1.62 0.01
14 3.8789 0.0111 17.75 0.29 15.4738 0.0010 1.64 0.01
15 3.9010 0.0125 21.33 0.32 15.4750 0.0010 1.56 0.01
16 3.9209 0.0143 22.12 0.36 15.4697 0.0017 1.61 0.01
17 3.9312 0.0151 23.37 0.37 15.4438 0.0022 1.67 0.01

Table 4.1: Summary of velocity dependence of torsional and bending modes

It should be noticed that the errors on the estimated resonances and damp-
ings are quite small, which are, converted to relative errors, in the order of 0.5%
to 5%. These can be verified in some way by repeating the measurements within
the same experimental session, re-estimating the dampings and resonances and
finally verifying if the values lie within the estimated uncertainty bounds.

We expect flutter to occur when the damping ratio corresponding to the


torsional mode becomes negative. The attempt here is to produce data which
show us that the damping corresponding to the torsional mode begins to decrease,
as this would lead to a more “data based” extrapolation(see chapter 5). As we
can notice, no decreasing trend can be found for the damping ratio of the torsional
mode as the flutter onset speed for this wing design is too high with respect to
the wind speed that can be emulated. Hence, the structural properties of the
wing had to be changed in order to lower the flutter onset speed. This has been
done by adding a mass at the trailing edge side extremity of the wooden beam,
following the procedure described in chapter 2 for decreasing the flutter onset
speed.

4.2.2 Error analysis


In this section, we will take a closer look at the variances of the estimated F RF ’s.
As mentioned before, different random phase multisine realizations were per-
formed in order to be able to make a distinction between nonlinearities and noise
on the one hand and the underlying linear system on the other hand. A further
distinction between noise and non linearities allow to make an estimation of the
nonlinear behaviour of the setup and the system. Furthermore, the contribu-
tions to the nonlinearities can be divided into systematic- and stochastic ones.
The stochastic contributions are the ones which vary as function of the random

37
phases of the input and they are zero mean, whereas this is not the case for the
systematic ones. Hence, by averaging the F RF over a number of experiments
with different realizations of the random multisines, the stochastic contribution
is averaged towards 0[7][8].
Note that one could have the tendency to average out the nonlinearities over
the periods as can be done for the noise. However as the input doesn’t change
over the periods, the nonlinear response also won't change over the periods and
thus it makes no sense to average over periods. Over different realisations in
contrast it makes sense to average out as the input changes over the realisations
and thus also the output.
Important to mention also is that the assumption is made that the system is
a P ISP O(Period In Same Period Out) system, which means that the nonlinear
contributions are only present at integer multiples of the excited frequencies, ex-
cluding phenomena such as chaos and bifurcations. It is for such systems, which
can be approximated by a Volterra series(this is a class of nonlinear systems),
only that the concept of its BLA applies, implying among other things that the
stochastic nonlinearities can be averaged out over different realisations[7][8].

However we must not lose sight of the fact that also the noise is contributing
to the variance calculated over the different realizations. Indeed, by averag-
ing over different random phase realizations we get the total variance(see sec-
tion 3.2.2). If we want to know the level of the stochastic nonlinearities, it is
necessary to substract the noise variance from the total variance. Recall that
the noise variance is estimated with the localpolynomial method as described in
section 3.2.2. Figure 4.6 illustrates standard deviations of the nonlinearities and
of the noise and also shows the total standard deviation for a wind velocity of
10m/s. The estimated F RF as well as the model are also shown.

38
Figure 4.6: F RF estimate, total-, stochastic nonlinearities- and noise standard
deviation for aeroelastic system without added weight and at wind speed of
10m/s

Note that in the above figure the total variance is the one with respect to
the obtained BLA, which is the one obtained after dividing the total variance
estimated by the robust LP M by the number of realizations.

The level of the nonlinearities with respect to the estimated BLA can be
expressed in terms of signal to distortion ratio SDR defined here as the ratio of
the power of YBLA and YS , where YS is the vector of standarddeviations of the
odd nonlinear distortions :
RM S(GBLA )
SDR = (4.2)
RM S(GS )

This number can be used to check the validity of the linear approximation by
comparing it with a given threshold SDRthreshold . If the obtained SDR is then
lower than SDRthreshold , a nonlinear system has to be modelled.

4.3 Second series of experiments: increased in-


ertial coupling through added mass
In this case a random phase multisine of 6 consecutive periods sampled at the
same sampling frequency and exciting in the same frequency band was used as
excitation signal. However the frequency resolution was doubled with respect to
the previous case. This resulted in a measurement time of 180 s for each real-
ization. Again multiple realizations were performed of which the number varied
as function of the different wind speeds.

39
The same identification procedure as before has been performed. Again, the
evolution as function of the velocity of the estimated F RF ’s and the correspond-
ing modelled damping ratios and resonance frequencies corresponding to the first
2 peaks of the estimated F RF ’s can be visualized in figures 4.7 and 4.8. As in
previous paragraph, we look at the 1st element of the estimated F RM .

Figure 4.7: velocity dependence of the estimated F RF ’s corresponding to setup


with additional mass

(a) velocity dependence of first mode (b) velocity dependence of second mode

Figure 4.8: velocity dependence of the first 2 peaks of the estimated F RF ’s and
corresponding modelled damping ratios and resonance frequencies

As is the case for the system without the added weight, the damping ratio
corresponding to the second mode increases as the wind velocity increases. As
for the resonance frequencies, the values decrease with increasing velocities. The
evolution as function of the velocity of the first mode takes place towards higher
damping ratios also with the exception of the decrease from 11m/s to 12m/s.

40
As for the resonance frequencies, they increase with increasing velocity with the
exception of the decrease from 12m/s to 13m/s.

For the sake of illustration, the comparison is made as in section 4.2 between
a model for which the order is chosen according to the peaks of the estimated
F RF and a model for which the residuals are scattered around the total standard
deviations.

(a) Eighth order model (b) sixteenth order model

Figure 4.9: Comparison between eighth and sixteenth order models for system
at 16m/s wind speed: sixteenth order model captures first mode of the system
and residuals are scattered around the total standard deviations

Figure 4.9 illustrates the comparison between an eighth- and a sixteenth or-
der model for the aeroelastic system at 16m/s. Again the good tracking of the
system by the model can be testified by all the peaks of the F RF being fitted
and the residuals being in the proximity of the total standard deviations, whereas
the eighth order model does not consider the system to have a pole at approx-
imately 4Hz and the corresponding residuals are systematically higher than the
total standard deviations.

As a summary, the following table and charts with the damping and resonance
frequencies as function of the velocity are shown along with their corresponding
uncertainty bounds.

41
(a) velocity dependence of damping: flexu- (b) velocity dependence of damping: tor-
ral mode sional mode

(c) velocity dependence of resonance fre- (d) velocity dependence of resonance fre-
quency: flexural mode quency: torsional mode

Figure 4.10: Graphical representation of damping ratios and resonance frequen-


cies corresponding to torsional- and flexural modes as function of wind speed:
with added weight

42
Mode Bending Torsion
Wind speed(m/s) fn (Hz) ∆fn (Hz) ξ(%) ∆ξ(%) fn (Hz) ∆fn (Hz) ξ(%) ∆ξ(%)
10 2.9082 0.0136 8.81 0.50 8.6378 0.0025 3.21 0.02
11 3.0220 0.0363 12.74 1.35 8.5927 0.0013 3.52 0.02
12 3.1072 0.0261 8.36 0.82 8.5871 0.0012 3.87 0.01
13 2.9897 0.0213 9.96 0.65 8.4718 0.0019 4.36 0.02
14 3.0143 0.0133 12.68 0.40 8.4488 0.0019 4.75 0.02
15 3.0197 0.0133 13.47 0.41 8.3289 0.0025 5.24 0.03
16 3.0784 0.0098 15.09 0.30 8.2901 0.0026 5.75 0.03
17 3.1258 0.0105 16.27 0.29 8.1926 0.0033 6.18 0.04
18 3.1550 0.0086 18.63 0.24 8.0949 0.0032 7.01 0.04
19 3.2228 0.0102 18.91 0.28 7.9780 0.0043 7.25 0.05

Table 4.2: Summary of velocity dependence of torsional and bending modes:


with added weight

Now if one compares the velocity dependence between the system with- and
without added weight in figure 4.11, one can observe that the velocity dependence
of the first modes are quite the same. The difference however is in the overall
damping values which are lower for the system with added weight.

(a) velocity dependence of damping corre- (b) velocity dependence of damping cor-
sponding to flexural mode: without added responding to flexural mode: with added
weight weight

Figure 4.11: Comparison of damping corresponding to flexural mode between


system with- and without added weight

The velocity dependence however of the second mode is more pronounced


for the system with added weight with respect to the system without added
weight, i.e. without added weight the system's damping corresponding to the

43
torsional(second) mode stays constant with velocity. However, the overall val-
ues of the damping ratios are higher, whereas one would expect a decrease in
overall damping as the inertial coupling has been increased, which should bring
the system closer to instability(see chapter 2, section 2.4). This is shown in
figure 4.12.

(a) velocity dependence of damping corre- (b) velocity dependence of damping corre-
sponding to torsional mode: without added sponding to torsional mode: with added
weight weight

Figure 4.12: Comparison of damping corresponding to torsional mode between


system with- and without added weight

Moreover the torsional mode is expected to decrease. This however has


still not been attained so more weight had to be attached to the wing in order
to change the structural properties such that the flutter onset speed would be
lowered some more.

4.4 Final series of experiments: further increase


of inertial coupling
Here, a random phase multisine of 6 consecutive periods sampled at the same
sampling frequency and exciting in the same frequency band was used as exci-
tation signal. However the frequency resolution was now increased with a factor
3
2
compared to the previous case. This resulted in a measurement time of 245s
for each realization. Again multiple realization were performed but this number
stayed constant for the different wind speeds.

Figures 4.13 and 4.14 show respectively the velocity dependence of the esti-
mated F RF ’s and specifically the velocity dependence of the second mode with

44
its corresponding resonance frequencies and damping ratios. The first mode is
not shown here as no meaningful trend can be deduced graphically as can be
noticed in figure 4.13. The velocity dependence of the damping and resonance
frequencies of this mode though do have a meaningful trend as can be seen in
table 4.3. This table and figure 4.15 summarize the damping ratios and reso-
nance frequencies at all the different wind speeds for the two modes along with
their corresponding standard deviation.

Figure 4.13: velocity dependence of the estimated F RF ’s corresponding to setup


with final configuration

Figure 4.14: velocity dependence of the second peak of the estimated F RF ’s


and corresponding modelled damping ratios and resonance frequencies

The aeroelastic system still behaves in approximately the same way as in the
two previous cases: an increasing damping with the velocity for both modes. The
resonance frequency corresponding to the first mode decreases with the velocity
whereas the one corresponding to the second mode increases with the velocity.
Some exceptions as before can also be noticed as the first mode having downward

45
jumps for the damping and resonance frequencies at 12m/s. One can notice that
at 14m/s, the error on the damping ratios and resonance frequencies are up to
one hundred times larger compared to the estimations at other velocities. This is
because at this velocity, the order of the model had to be pushed up as otherwise
the fit didn’t capture the first peak of the estimated F RF.

Mode Torsion Bending


Wind speed(m/s) fn (Hz) ∆fn (Hz) ξ(%) ∆ξ(%) fn (Hz) ∆fn (Hz) ξ(%) ∆ξ(%)
0 7.5120 0.001 1.33 0.05 2.4240 0.001 1.48 0.06
10 7.3995 0.0014 4.52 0.02 2.5539 0.0054 7.61 0.22
12 7.3064 0.0023 5.55 0.03 2.6472 0.0122 10.41 0.50
14 7.2194 0.4309 6.44 1.65 2.6179 0.1176 7.69 0.99
15 7.1233 0.0038 7.18 0.05 2.6396 0.0131 8.7 0.46
16 7.0300 0.0043 7.88 0.06 2.6762 0.0111 9.06 0.39
17 6.9355 0.0036 7.88 0.06 2.7168 0.0081 9.77 0.28
18 6.8317 0.0040 8.64 0.06 2.7480 0.0078 11.29 0.27
19 6.7362 0.0045 9.63 0.07 2.7466 0.0081 13.12 0.27

Table 4.3: Summary of velocity dependence of torsional and bending modes:


with added weight

46
(a) velocity dependence of damping: flexu- (b) velocity dependence of damping: tor-
ral mode sional mode

(c) velocity dependence of resonance fre- (d) velocity dependence of resonance fre-
quency: flexural mode quency: torsional mode

Figure 4.15: Graphical representation of damping ratios and resonance frequen-


cies corresponding to torsional- and flexural modes as function of wind speed:
final configuration

Comparing the velocity dependence of the damping with the previous case,
the same conclusions as previously can be drawn: an overall increase in damping
for the torsional mode and an overall decrease in damping for the flexural mode.
This shows that still, the decrease in damping corresponding to the torsional
mode has still not been attained. However we might get to this decrease by
extrapolating the transfer function coefficients, which brings us to the following
chapter. This decrease has been attained in [4][12] with this kind of extrapolation
using also experimental data where only an increasing torsional damping could
be observed. Following figures compare the dampings corresponding to the first
and second modes between the two last configurations.

47
(a) velocity dependence of damping corre- (b) velocity dependence of damping corre-
sponding to flexural mode: second config- sponding to flexural mode: final configura-
uration tion

Figure 4.16: Comparison of damping corresponding to flexural mode between


system with second and final configuration

(a) velocity dependence of damping corre- (b) velocity dependence of damping corre-
sponding to torsional mode: second config- sponding to flexural mode: final configura-
uration tion

Figure 4.17: Comparison of damping corresponding to torsional mode between


system with second and final configuration

48
Chapter 5

Flight flutter prediction

5.1 Introduction
In this chapter, an attempt is made to extrapolate the damping of the aeroelastic
system towards higher speeds.

One could think about straightforwardly extrapolate the damping ratio as


function of speed based on the experimental couples. This simple approach
however has a major drawback: as we don’t have a straightforward theoretical
relation between the damping ratio and the wind speed, the extrapolation cannot
be adjusted to the theoretical evolution of the damping as function of the wind
speed. Besides, as in the experiments no decreasing damping could be observed,
the extrapolation, even if it would succeed in finding a speed for which the damp-
ing becomes zero, this would not have any experimental basis as the results do
not suggest a decreasing damping ratio for higher speeds. Worse still, even if the
experiments would indicate a decreasing damping ratio, this still wouldn’t be an
argument to suggest that the found flutter speed has accurately been determined
using extrapolation because it cannot be confirmed that from that speed on the
damping will continue to decrease.

A more reliable approach would be to extrapolate the coefficients of the


denominator polynomial corresponding to the modelled aeroelastic system(see
equation 2.30). Indeed, as this polynomial describes theoretically the depen-
dence of its coefficients on the speed, we now have a ’physical’ basis which can
be utilized to adjust the extrapolation to the theoretical evolution of those co-
efficients. This is done specifically by fitting a polynomial of the same order
as the ones prescribed by the(theoretical) model in equation 2.30 through the
experimentally determined denominator coefficients. Subsequently, the damp-

49
ing ratios and resonance frequencies of the aeroelastic system at those higher
speeds can be calculated given the extrapolated denominator coefficients by the
route of calculating the roots(extrapolated poles) of the corresponding polyno-
mial, leading to an estimate of the flutter speed at the point where the damping
corresponding to the torsional mode changes its sign. This procedure is elabo-
rated more in detail in the paragraph to follow. Note that the used data for the
extrapolation is the data corresponding to the last series of experiments. Then,
the extrapolations will be evaluated through a validation procedure. Finally, the
error on the extrapolations is derived.

5.2 Flutter speed prediction: extrapolation of


transfer function coefficients
Recall the relations between each of the denominator coefficients and the velocity:
j
ri,p U p (5.1)
X
αi =
p=0

Where i is the coefficient number(from 0 up to and including 4, for decreasing


order of the Laplace variable s) and j the corresponding order of the polynomial
in U , following the sequence [0, 1, 2, 3, 2] for increasing i. This implies that the
first coefficient α0 = 1.

Firstly, it needs to be realized that the attempt is to extrapolate linear time


invariant systems. Hence they should be consistent in terms of model order, as
otherwise the coefficients i cannot be related with each other as function of the
wind speed. However, from the previous chapter, it appears that for different
velocities, different model orders were obtained, whereas the aeroelastic system
does theoretically not suggest that a speed variation would give rise to a chang-
ing polynomial order in s. The reason for this difference in order between the
theoretical, white box model and the grey box model obtained through system
identification is that the latter is obtained by forcing a parametric rational func-
tion through a fluctuating, experimentally obtained F RF , by pushing up the
order of the corresponding denominator polynomial order(see chapter 4), giving
rise to mathematical poles in addition to the physical poles. As these physi-
cal poles can readily be recognized by looking at their corresponding resonance
frequencies, the mathematical ones can be thrown away. As appears from the
theory and the experiments(from the theory we know where to expect these poles
in the frequency domain and from the experiments we find that they are the first
ones appearing in the graph corresponding to the parametric model), the phys-

50
ical poles are two complex conjugate pairs which are located at the first two
resonance frequencies in the obtained parametric model. Note that the obtained
resonance frequencies and damping ratios presented in chapter 4 correspond to
those physical poles. Hence, the physical poles of the system can be extracted
for each obtained linear time invariant system and bring forth denominator poly-
nomials(using the function ’poly()’ in MATLAB) for the different velocities which
are consistent in terms of polynomial order.

(a) Extrapolation for α1 (b) Extrapolation for α2

(c) Extrapolation for α3 (d) Extrapolation for α4

Figure 5.1: Polynomial fits and extrapolation for the different denominator coef-
ficients, each having a polynomial order prescribed by equation 5.1(blue) along
with the corresponding uncertainties(red) and experimentally obtained coeffi-
cients(yellow stars)

The corresponding(experimentally obtained) coefficients αi at the different


wind speeds are then coupled with each other by fitting a polynomial of order
corresponding to the respective ith coefficient through the experimentally ob-
tained coefficients, using the function ’polyfit()’ in MATLAB which allows to

51
pass on the order of the fitting polynomial. Another good thing about this func-
tion is also that it outputs a(n)(optional) variable which can be used as input
when evaluating the fitted polynomial at a certain(higher) velocity to output the
corresponding standard deviation. The coefficients could then be extrapolated
towards higher speeds using ’polyval()’ at the desired velocity. The additional
interval ranged from 20m/s to 30m/s. The polynomial fits for each coefficient
αi along with their uncertainty bounds are given in figure 5.1, from which can
be noticed that the experimental coefficients follow the trend that is expected
theoretically quite well, especially the second experimentally obtained coefficient
as function of the wind speed(figure 5.1a).

In a next step, the range of velocities in which the damping ratio correspond-
ing to the torsional mode changes its sign can be tracked using this extrapolation,
by calculating for each extrapolated set of coefficients αi,Extra at higher velocities
the roots with the help from the MATLAB function ’roots()’. From the roots
the damping ratio as well as the resonance frequency can be calculated in the
same way that has been done in the previous chapter: for a given pole p∗ , the
damping ratio and resonance frequency are respectively:
Re(p∗ ) Im(p∗ )
ξ∗ = f ∗
n = (5.2)
Im(p∗ ) 2π
From the resonance frequencies it can then be deduced which damping corre-
sponds to which mode, by looking at the extrapolated resonance frequencies of
the four poles and observing which one of them lies in the continuity of the
experimental resonance frequencies found for the torsional- and flexural mode.
The following table gives the extended part of table 4.3 up to 25m/s, in steps
of 1m/s:

Mode Torsion Bending


Wind speed(m/s) fn (Hz) ξ(%) fn (Hz) ξ(%)
20 6.66 8.69 2.77 16.53
21 6.43 7.95 2.80 20.73
22 6.28 6.53 2.83 26.49
23 6.36 -2.57 2.82 34.36
24 6.32 -5.68 2.77 44.98
25 6.31 -9.01 2.65 59.13

Table 5.1: Extrapolated damping ratios and resonance frequencies towards higher
speeds for torsional and flexural modes

Firstly, it can be noticed that the damping ratios corresponding to the flexural
mode keep on increasing with velocity. Also, all the values follow a trend for both

52
modes that is in the extension of the experimental values at the lower veloci-
ties(see table 4.3). Then, it can be concluded that the flutter speed should be
somewhere between 22m/s and 23m/s as the damping of the system changes its
sign in this speed interval. Therefore, a closer look in this interval allows to look
for a more ’accurate’ value for the flutter speed. Figure 5.2 shows the evolution
as function of the velocity of the damping ratio and the approximate location of
the flutter speed which is at 22.3m/s.

Figure 5.2: Estimated evolution of damping ratio as function of wind speed and
approximate location of flutter speed

5.3 Validation: evaluation of the extrapolations


Note that as an error exists on the experimental coefficients, these errors prop-
agate through the extrapolation of these coefficients in addition to the extrap-
olation error and hence an error exists on the calculation of the extrapolated
resonance frequencies, damping ratios and thus also on the predicted flutter
speed. In order for these extrapolations to be relevant, they should be evalu-
ated through a so called validation, which consists of starting from a fragment
of the experimentally obtained coefficients corresponding to a fragment of the
velocity range, extrapolating these coefficients and comparing the calculated ex-
trapolated dampings and resonance frequencies to the experimentally obtained
dampings and resonance frequencies.

We will start from the five first sets of coefficients, corresponding to the ve-
locity interval [0, 14]m/s. Extrapolation towards higher velocities yield damping
ratios and resonance frequencies corresponding to the torsional mode which can
be visualized in figure 5.3.

53
(a) Damping ratio ξ (b) Resonance frequency fn

Figure 5.3: Extrapolation validation of ξ and fn starting at 14m/s

If the extrapolation starting from 14m/s goes through the experimental data,
we can say that the extrapolations are reliable. Unfortunately this is not the case.
It can be seen that, the further away we look in the extrapolation, the larger the
difference becomes with the experimental dampings. Therefore, we will check
the extrapolation starting from higher velocities. This is shown in figure 5.4 and
figure 5.5.

54
(a) Final velocity at 15m/s (b) Final velocity at 16m/s

(c) Final velocity at 17m/s (d) Final velocity at 18m/s

Figure 5.4: Extrapolation validation of ξ starting at 15m/s, 16m/s, 17m/s, and


18m/s

We can see that the extrapolations are more or less reliable when we extend
the velocity range with say 1m/s to 2m/s. Beyond this extended range the
extrapolated dampings fall far outside the uncertainties of the corresponding
experimental dampings.

55
(a) Final velocity at 15m/s (b) Final velocity at 16m/s

(c) Final velocity at 17m/s (d) Final velocity at 18m/s

Figure 5.5: Extrapolation validation of fn starting at 15m/s, 16m/s, 17m/s,


and 18m/s

It can be concluded that the extrapolations of the resonance frequencies have


a more stable behaviour compared to the extrapolated dampings, i.e. they are
more or less not far from the uncertainties of the corresponding experimental
resonance frequencies.

Now, a comparison is made between the experimental dampings and reso-


nance frequencies on the one hand and the extrapolated ones on the other hand.
In this way, it will be observed to which extent the extrapolated dampings and
resonance frequencies fall outside(if they do) of the uncertainty bounds of the
ones experimentally obtained. Tables 5.2 and 5.3 give respectively the differ-
ences between the experimental and extrapolated dampings and the differences
between the experimental and extrapolated resonance frequencies, both with the
corresponding experimental standarddeviations.
From the tables, we can conclude that for the dampings, the overall ratio
between the extrapolation errors and the standarddeviations is of the order of

56
UF inal 14m/s 15m/s 16m/s 17m/s 18m/s |(∆ξ)exp |
(∆ξ)14m/s 0.778% 1.65%
(∆ξ)15m/s 2.702% 0.434% 0.05%
(∆ξ)16m/s 5.75% 1.273% 0.119% 0.06%
(∆ξ)17m/s 10.38% 3.3% 1.007% 2.97% 0.06%
(∆ξ)18m/s 14.65% 5.1% 1.277% 4.68% 1.014% 0.06%
(∆ξ)19m/s 18.58% 7.21% 1.49% 6.54% 0.86% 0.07%

Table 5.2: Differences between extrapolated dampings and experimental damp-


ings for each starting set of experimentally obtained dampings and comparison
with the corresponding standarddeviations

UF inal 14m/s 15m/s 16m/s 17m/s 18m/s |(∆fn )exp |


(∆fn )14m/s -0.0244% 0.4309%
(∆fn )15m/s 0.0237% 0.0093% 0.0038%
(∆fn )16m/s 0.094% 0.015% 0.0070% 0.0043%
(∆fn )17m/s 0.2165% 0.0405% 0.0025% -0.0095% 0.0036%
(∆fn )18m/s 0.4143% 0.0833% 0.0043% -0.0137% 0.0197% 0.0040%
(∆fn )19m/s 0.6698% 0.1338% 0.0118% -0.0152% 0.0232% 0.0045%

Table 5.3: Differences between extrapolated resonance frequencies and exper-


imental resonance frequencies for each starting set of experimentally obtained
resonance frequencies and comparison with the corresponding standarddeviations

57
100. For an extended velocity of 1m/s an averaged ratio of 50 is obtained. For
the resonance frequencies, those ratios are much smaller, namely an overall ratio
of 10.

5.4 Validation: error on the extrapolated poles


In this section, a method will be presented to determine the covariance on the
extrapolated poles given the covariance on the experimentally obtained poles.

First of all, note that the error on the poles propagated through the calcula-
tion of coefficients αi,experimental from these poles, the calculation of extrapolation
polynomial coefficients for each αi,experimental , and the calculation of extrapolated
poles from extrapolated coefficients. For each of these steps the propagated error
will be derived. We shall begin with deriving a function relating the poles to the
coefficients αi,experimental .

Given are the roots p1,exp , p2,exp , p3,exp and p4,exp corresponding to the fol-
lowing polynomial(where α0,exp = 1):
P (s) = s4 + α1,exp s3 + α2,exp s2 + α3,exp s + α4,exp
This equation can be written in terms of its roots:
P (s) = (s − p1,exp )(s − p2,exp )(s − p3,exp )(s − p4,exp )
Elaborating this equation, regrouping the terms with a common order of s and
identifying with the previous equation yields a system of nonlinear equations in
the roots p~exp and solution vector α
~ exp = [α1,exp α2,exp α3,exp α4,exp ]T :
α1,exp = −(p1,exp + p2,exp + p3,exp + p4,exp )
α2,exp = p1,exp p2,exp + p1,exp p3,exp + p1,exp p4,exp + p2,exp p3,exp + p2,exp p4,exp + p3,exp p4,exp
α3,exp = −(p1,exp p2,exp p3,exp + p1,exp p2,exp p4,exp + p1,exp p3,exp p4,exp + p2,exp p3,exp p4,exp )
α4,exp = p1,exp p2,exp p3,exp p4,exp

Which can be compactly rewritten as:

α1,exp = f1 (~pexp )
α2,exp = f2 (~pexp )
α3,exp = f3 (~pexp )
α4,exp = f4 (~pexp )

58
and further as:
~ exp = f~(~pexp )
α (5.3)
With f~ : C4 → C4 being a nonlinear complex valued vector function. From [11]
we find, for two sets of complex valued variables which are not linearly related,
that the relation between their covariances is
αexp ) = J · cov(~pexp ) · JH
cov(~ (5.4)
Where the operator H denotes the hermitian or conjugate transpose and J the
Jacobi matrix:  
∂f1 ∂f1 ∂f1 ∂f1
 ∂p 1 ∂p2 ∂p3 ∂p4 
 ∂f2 ∂f2 ∂f2 ∂f2 
(5.5)
 ∂p1 ∂p2 ∂p3 ∂p4 
J=  ∂f3 ∂f3 ∂f3 ∂f3 
 
 ∂p1 ∂p2 ∂p3 ∂p4 
∂f4 ∂f4 ∂f4 ∂f4
∂p1 ∂p2 ∂p3 ∂p4
Note that we replaced pi,exp with pi to make the writing easier. Furthermore,
cov(~pexp ) satisfies following expression, assuming the poles are uncorrelated:
 2
σp1 0 0 0

 0 2
σp2 0 0 
cov(~pexp ) = (5.6)
 
 0 2
0 σp3 0 
 

0 0 0 σp24
Where
σp2i = σRe(p
2
i)
2
+ σIm(pi)
(5.7)
Now that we have the means to determine cov(~ αexp ), we will derive how this
uncertainty manifests itself in the extrapolation polynomial coefficients for each
coefficient αi . These extrapolation polynomial coefficients will be denoted ri,n ,
where the index i refers to the particular coefficient αi which is extrapolated at
higher velocities and the index n refers to the order of the corresponding variable
velocity. The order of extrapolation corresponding to the order imposed by the
theoretical form of the coefficients αi (see equation 5.1) will be denoted Rn .
Note that the above covariances are determined for each coefficient αi,exp but
also for each velocity for which the poles p~exp were experimentally determined.
Since we have nine velocities at which the poles were experimentally determined,
we will also have a set of nine coefficients αUi j ,exp for each i. Where the index
j refers to the corresponding velocity Uj at which the coefficient αi,exp was
experimentally obtained(or compactly written as αUi j ,exp ). Denote α ~ i such a set,
which can be written as follows:
1 U1 U12 · · · U1Rn
 i    
α U1 ri,0
 i  R 
αU2  1 U2 U22 · · · U2 n   ri,1 
 
~i =  ¯ i ri (5.8)
α  ..  =  .. .. ..   ..   = Ū ~
   
 .  . . .  . 
αUi 9 1 U9 U92 · · · U9Rn ri,Rn

59
Remark that we didn't write α i
~ exp as the extrapolations on the right hand side
of equation 5.8 forces a polynomial through the experimental coefficients in a
least squares sense and thus at the 'experimental' velocities the coefficients at
the left hand side will not have the same value as the experimentally obtained
coefficients. We will assume however that the covariances of these coefficients at
the left hand side will be equal to the covariances of the experimentally obtained
coefficients α i
~ exp .

Now think away the index i, as following reasoning applies for each set of
coefficients α i
~ exp . If we want to find the covariances of the extrapolation polyno-
mial coefficients rn , we first need to express the inverse relation between α~ and
~r:
¯ +α
~rˆ = Ū ~ (5.9)
Where the operator + denotes Moore Penrose pseudo inverse as we do not have
a square matrix. We now have an expression like 5.3. Again, we find from
[11] that, for two sets of complex valued variables which are (this time) linearly
related, the relation between their covariances is:
¯ + cov(~
cov(~rˆ) = Ū ¯ +H
α)Ū (5.10)
As a consequence of the above remark, we shall equate cov(~
α) to cov(~ i
αexp ).

We have now the means to determine the covariances of the extrapolation


polynomial coefficients ~rˆi for each coefficient αi . With these, we can calculate
the covariances of the extrapolated coefficients α i
~ extra at each velocity outside of
the range corresponding to the experimental data. For this, we shall first estab-
lish a relation between the extrapolated coefficients α i
~ extra and the extrapolation
ˆ
polynomial coefficients ~r . This relation is almost exactly the same as relation
i

5.8, but now we have on the left hand side a vector of extrapolated coefficients
αUi j ,extra of size equal to the number of velocities for which extrapolated co-
efficients are desired outside of the experimental velocity range, which will be
denoted m. On the right hand side, we will have the same vector ~ri but the
velocity matrix will have this time m rows and Rn + 1 columns:
U12 · · · U1Rn
 i    
αextra,U1 1 U1 ri,0
 i 2 Rn  
i
 αextra,U2 

1 U2

U2 · · · U2   ri,1  ¯ i ri (5.11)
..   ..  = ŪExt~

α
~ extra =  ..
= .. .. 
   
. . . .  . 
 
 
i 2 Rn
αextra,U m
1 Um Um · · · Um ri,Rn
Where the subscript Ext stands for extended as this matrix includes the velocities
outside of the range of velocities at which coefficients were obtained experimen-
tally. We can then, as above, derive the expression for the covariance of the

60
extrapolated coefficients:

cov(~ i
αextra ¯ i cov(~r )Ū
) = Ū ¯ iH (5.12)
Ext i Ext

As we have determined only the covariances of ~rˆi we can replace cov(~ri ) with
cov(~rˆi ).

At this point we have the covariances of the extrapolated coefficients αextra,U


1
j
,
αextra,Uj , αextra,Uj , and αextra,Uj for each extended velocity Uj . Now, let's re-
2 3 4

group for each velocity Uj those coefficients into a vector:


 T
α 1
~ extra (Uj ) = αextra,U j
2
αextra,U j
3
αextra,U j
4
αextra,U j
(5.13)

The idea is now to produce an expression like 5.3 and to find the inverse relation.
This will yield the extrapolated poles p~extra (Uj ) at velocity Uj as function of the
extrapolated coefficients α~ extra (Uj ) of which, in the same way as for 5.3, the
relation between their covariances can be derived:

αextra ) · IH
αextra (Uj )) ⇒ cov(~pextra ) = I · cov(~
p~extra (Uj ) = ~g (~ (5.14)

Where ~g is the nonlinear complex valued vector function relating the extrapolated
poles to the extrapolated coefficients and I the Jacobi matrix associated to ~g (note
that pi,extra has been replaced by pi to make the writing easier):
 
∂g1 ∂g1 ∂g1 ∂g1
 ∂p 1 ∂p2 ∂p3 ∂p4 
 ∂g2 ∂g2 ∂g2 ∂g2 
I=
 ∂p1
 ∂g3 ∂p2
∂g3
∂p3
∂g3
∂p4 
∂g3  (5.15)
 
 ∂p1 ∂p2 ∂p3 ∂p4 
∂g4 ∂g4 ∂g4 ∂g4
∂p1 ∂p2 ∂p3 ∂p4

The MATLAB function CovRoots from the toolbox of Prof. Rik Pintelon provides
the implementation of this last step, given the extrapolated coefficients and their
covariances. The description of this function can be found in [9].

61
Chapter 6

General conlusion

In this thesis, the damping behaviour of an aeroelastic system has been investi-
gated as function of the velocity and wing inertial properties, along with an error
analysis to asses the reliability of the obtained linear models.

The results show in general that the AATB in combination with the used
identification techniques succeeded in producing mathematical models which are
in accordance with what would be expected on the basis of what the theory
prescribes, which is a decrease in damping at a certain velocity. However, the
increase of the overall damping corresponding to the torsional mode for an in-
creased inertial coupling was found to be in contradiction with what intuitively
would be expected, which is that coupling in general of the modes of the torsional
and flexural vibrations result in a decreased stability of the system[2]. Yet this
evolution reveals the dependence of flutter behaviour on inertial properties.

Furthermore, the theoretical development in chapter 2 lead to the possibility


of making an extrapolation which was forced to result in an evolution as function
of the wind velocity as prescribed theoretically. The extrapolation results went
through a validation procedure, which lead to the conclusion that reliability of
the extrapolated resonance frequencies is ensured for a larger extension of the
experimental velocity interval than for the damping ratios.

As this has not been performed yet on this test bench, future works using
time varying identification techniques allowing to measure at once the speed de-
pendence of the aeroelastic system, by modelling the system as time varying,
therefore outperforming in terms of measurement time classical flutter analysis
as has been performed in this thesis. Moreover, multiple time invariant systems
have been constructed, resulting in a lot of parameters to be estimated compared
to one time varying system, implying a higher variability of the predictions. More-

62
over, a comparison can be made between grey box and white box approximations
after implementing the mathematical model described in chapter 2 section 2.2.

63
Bibliography

[1] G. Alvin Pierce Dewey H. Hodges. Introduction to Structural Dynamics


and Aeroelasticity. Cambridge university Press, 2011.
[2] G. Dimitriadis. Aeroelasticity and Experimental Aerodynamics.
[3] R. Pintelon J. Lataire J. Schoukens Gerd Vandersteen Tadeusz Dobrowiecki.
“Robustness Issues of the Best Linear Approximation of a Nonlinear Sys-
tem”. In: IEEE TRANSACTIONS ON INSTRUMENTATION AND MEA-
SUREMENT 58 (2011), pp. 1737–1745.
[4] Julien Ertveldt. “Application of frequency domain system identification to
wind tunnel experiments on the Active Aeroelastic Tet Bench”. PhD thesis.
VUB, Faculty of Engineering, 2017.
[5] Y. C. Fung. An introduction to the theory of aeroelasticity. Dover publi-
cations, 1955.
[6] Patrick Guillaume. Modal Analysis.
[7] Johan Schoukens Rik Pintelon. System Identification: A frequency domain
approach. IEEE Press, 2001.
[8] R. Pintelon K. Barbé G. Vandersteen J. Schoukens. “Improved (non-
)parametric identification of dynamic systems excited by periodic signals”.
In: Mechanical Systems and Signal Processing 25 (2011), pp. 2683–2704.
[9] R. Pintelon P. Guillaume J. Schoukens. “Uncertainty calculation in (op-
erational) modal analysis”. In: Mechanical Systems and Signal Processing
21 (2007), 2359â–2373.
[10] R. Pintelon Y. Rolain G. Vandersteen J. Schoukens. “Improved (non-
)parametric identification of dynamic systems excited by periodic signals-
The multivariate case”. In: Mechanical Systems and Signal Processing 25
(2011), pp. 2892–2922.
[11] Joel Tellinghuisen. “Statistical Error Propagation”. In: J. Phys. Chem 105
(2001), pp. 3917–3921.

64
[12] J. Ertveldt J. Lataire R. Pintelon S. Vanlanduit. “Frequency-domain iden-
tification of time varying systems for analysis and prediction of aeroelastic
flutter”. In: Mechanical Systems and Signal Processing 47 (2014), pp. 225–
242.

65
Appendices

66
TECHNISCHE
EIGENSCHAPPEN
FLOORMATE™ 200 SL-X
CE markering: T1-CS(10\Y)200-DS(TH)
Product eigenschappen 1) norm eenheid CE Code FLOORMATE 200 SL-X
Thermische eigenschappen
Warmtegeleidingscoëffciënt λD
Dikte (dN) 60 - 100mm EN 13164 W/(m•K) 0,029
Druksterkte kPa 200
EN 826 CS(10\Y)200
bij 10% vervorming/breuk (90d) N/mm² 0,20
kPa 8000
Elasticiteismodulus EN 826 -
N/mm² 8
Lange termijn druksterkte kPa 60
EN 1606 -
(2% vervorming, 50 jaar) Toelaatbare ontwerp druksterkte N/mm² 0,06
Dimensionele stabiliteit / Maatvastheid
bij temperatuur en vochtigheid onder belasting bij EN 1604 % DS(TH) ≤5
temperatuur en vochtigheid EN 1605 % - -
Wateropname
- bij langdurige onderdompeling (hele plaat) EN 12087 vol% - ≤0,4
- diffusie dN = 50 mm -
- diffusie dN = 100 mm EN 12088 vol% WD(V)3 -
- diffusie dN = 200 mm -
(tussenliggende waarden mogen geïnterpoleerd worden)
- na vries/dooi EN 12091 vol% - -
Brandgedrag
Brandgedrag is afhankelijk per toepassing en EN 13501-1 - - Euroclass E
methode van aanbrengen
Diffusieweerstand (μ-waarde) EN 12086 MU i 160 - 120
Afmetingen
dikte EN 823 mm - 60, 70, 80, 100
breedte x lengte (dikte met * op aanvraag) EN 822 mm x mm T1 600 x 1200
Afwerking
oppervlak - glad
randafwerking - sponning

Lineare uitzettingscoëfficiënt mm/m/K - 0,07


Temperatuursbestendigheid °C - -50°C / +75°C
Volumieke massa (min.) kg/m³ - 25
Toepassing - vloeren
Omkeerdak - -
Certificatie / Attesten Vloeren - KOMO IKB1816
Kelderwand - -
1)
waarden alleen voor schuim (zonder afwerking)

Warmteweerstand RD:
Dikte [mm] 60 70 80 100
RD [m².K/W] 2,10 2,45 2,80 3,50

Opmerking
De informatie en gegevens opgenomen in deze brochure kunnen niet worden opgevat
als verkoopspecificaties. Speciale kenmerken van de genoemde producten kunnen Dow Benelux B.V.
afwijken. De in deze brochure opgenomen informatie wordt in goed vertrouwen verstrekt, Oplossingen voor de Bouw
maar impliceert echter geen enkele aansprakelijkheid, garantie of verzekering voor wat
betreft de productprestatie. De koper is ervoor verantwoordelijk zeker te stellen dat deze Postbus 48
Dow-producten geschikt zijn voor de gewenste toepassing en te waarborgen dat de 4530 AA Terneuzen
plaats van het werk en de toepassingmethode in overeenstemming zijn met de huidige Nederland
wetgeving. Het onderhavige vormt geen licentieverlening voor het gebruik van patenten
of andere industriele of intellectuele eigendomsrechten. Wij adviseren om bij de aankoop E-Mail : styrofoam-nl@dow.com
van Dow-producten de meest recente aanwijzingen en aanbevelingen te volgen. Internet : www.styrofoam.nl
®™ Handelsmerk van The Dow Chemical Company (“Dow”) of van een tot de Dow-groep behorende vennootschap. 291-22536-0512

You might also like