AE4ASM506 Notes 2012
AE4ASM506 Notes 2012
AE4ASM506 Notes 2012
60
50
40
30
20
10
0
0 50 100 150 200
S. J. Hulshoff
HSL 0.36, Aerodynamics Group
Faculty of Aerospace Engineering, TU Delft
S.J.Hulshoff@TUDelft.NL
Version 11.1
2
Contents
1 Introduction 9
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2 Course Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3 Additional Course Information . . . . . . . . . . . . . . . . . . . 13
1.3.1 Course format: . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3.2 Prerequisite Concepts . . . . . . . . . . . . . . . . . . . . 14
1.3.3 Main References . . . . . . . . . . . . . . . . . . . . . . . 14
1.3.4 Instructor Contact . . . . . . . . . . . . . . . . . . . . . . 14
1.4 Characteristics of Aeroelastic Problems . . . . . . . . . . . . . . 15
3 Unsteady Aerodynamics 47
3.1 Vorticity Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.2 Compressibility effects . . . . . . . . . . . . . . . . . . . . . . . . 52
3.3 Viscous effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3
4 CONTENTS
Introduction
1.1 Background
The field of aeroelasticity considers phenomena in which interactions occur be-
tween aerodynamic flows and elastic structures. In this definition, the word
“interaction” is significant. To be precise, we mean phenomena in which the
aerodynamic forces are a function of the structural displacement, and the struc-
tural displacement is a function of the aerodynamic forces. This is easily demon-
strated by an example, such as the glider shown below:
9
10 CHAPTER 1. INTRODUCTION
Modern Aircraft
Maximum Efficiency/Performance
@
@
R
@
Slender Lightweight
@
@
R
@
Flexible !
over the last few decades as the performance of aircraft has increased. High-
performance aircraft tend to be light and slender, which implies reduced induced
aerodynamic drag, but also increased flexibility. In addition, high operating
speeds mean that the changes in aerodynamic force for a given amplitude of
deflection are relatively large, leading to large-amplitude aeroelastic responses.
? ?
Static Dynamic
- negligible time gradients - oscillatory
- Known as “divergence” - Known as “flutter”
∆L ∆L
∆α ∆α
t t
? ?
Static Dynamic
Determine the steady Determine the unsteady
operating condition response of an aircraft
of an aircraft taking to an applied excitation:
structural deformation
into account: - pilot input
- atmospheric turbulence
- Elevator angle to - landing loads
trim for a flexible fuse-
lage
δe
Forces:
SS A: Aerodynamic
A E: Elastic
CE I: Inertial
D DS
Problem Type:
F DR
F: Flutter
DR: Dynamic Response
D: Divergence
E I CE:
SS:
Control Effectiveness
Static Aircraft Stability
DS: Dynamic Aircraft Stability
MV MV: Mechanical Vibrations
Effects:
I A: Aerodynamic
E: Elastic
E I: Inertial
A C: Control
T: Thermodynamic
The date on figure 1.7 gives you some idea of the context. Recently it has be-
come fashionable to generalise this a bit, as shown in figure 1.8. This is because
high-speed designs may also experience thermal coupling, and there is consider-
able ambition to optimise the response of multi-physics systems using advanced
control concepts. These newer problems fall under the areas of Aeroservoelas-
ticity (Aeroelasticity + high-gain control systems), and Aerothermoelasticity
(Aeroelasticity + thermal effects).
A nice example of Aeroservoelastic engineering is the X-53, an aircraft which
makes use of active leading and trailing edge controls to achieve aeroelastically
deformed wing shapes with optimal performance (figure 1.9). This is a big
contrast to the old way of doing things, (i.e. make the wing stiff enough to
resist the moments incurred by control motion) and can lead to much lighter
structures and increased control effectiveness. The latter is demonstrated figure
18 CHAPTER 1. INTRODUCTION
1.10, which shows that for the X-53, the largest roll control authority is obtained
using leading-edge outboard devices on a lighter, flexible wing.
1.4. CHARACTERISTICS OF AEROELASTIC PROBLEMS 19
Figure 1.10: X53 control effectiveness for leading and trailing edge (LE, TE)
inboard and outboard (I,O) devices
20 CHAPTER 1. INTRODUCTION
Chapter 2
Simplified Aeroelastic
Problems
We will begin our look at aeroelastic analysis techniques using structural models
which have only one or two degrees of freedom (DOF), and using only quasi-
steady aerodynamics. This allows all the basic concepts of aeroelastic analysis
to be introduced, however, and provides a first look at some of the physical phe-
nomena that play a role in realistic systems. The models presented here may be
too crude for use in detailed design, but they can still be applied in preliminary
design stages as their simplicity allows the influence of design variables to be
clearly identified.
We will start by describing simple structural and aerodynmaic models for
a typical airfoil section. Then approaches to deriving both static and dynamic
aeroelastic boundaries will given for the problems of divergence, control reversal
and flutter. Afterwards we will briefly introduce two other simplified systems
which can be used for preliminary analysis, namely the semi-rigid wing and a
model for propeller whirl flutter.
21
22 CHAPTER 2. SIMPLIFIED AEROELASTIC PROBLEMS
Typical
@@@@
Section
@@@@ Mean
@@@@
y
Position
z
U Linear Spring
Torsional Spring
@@@@@@
@@@@@@
@@@@@@
Mean
h
@@@@@@ Position
EA
h
Kh
ab
Kθ
EA θ b b
c
EA CG L
h
θ AC
α
U EA
MAC
xθ b
ec
b b
Mass Parameters:
Aerodynamic Parameters
CG: Center of Gravity
xθ: Displacement of CG from EA U: Freestream velocity
m: Mass AC: Aerodynamic center
Sθ: Static moment related to EA (mxθb) e: eccentricity factor
rθ b: Radius of gyration about EA α: Angle of attack
L: Lift
Iθ: Mass moment of inertia related to EA MAC: Moment about AC (independent of α)
(m rθ2 b2) Dynamic pressure ( ρU /2)
2
q:
h
EA H
θ
xδ b
δ
δ: Control deflection
xδ b: Hinge axis position
H: Hinge moment
Sδ: Static moment related to hinge axis
I δ: Mass moment of inertia related
to hinge axis
α
EA
U
ḣ(t)
αLF (t) = α(t) + (Small-angle approximation) (2.3)
U
Thus:
!
ḣ(t)
L(t) = qSCLα α(t) + (2.4)
U
!
ḣ(t)
MEA (t) = qSecCLα α(t) + (2.5)
U
2.1. THE TYPICAL SECTION 25
α
ULF EA
h
U
mḧ + Sθ θ̈ + Sδ δ̈ + Kh h + L = 0 (2.7)
Sθ ḧ + Iθ θ̈ + (Iδ + xδ bSδ )δ̈ + Kθ θ − MEA = 0 (2.8)
Sδ ḧ + (Iδ + xδ bSδ )θ̈ + Iδ δ̈ + Kδ δ − H = 0 (2.9)
where θ and δ are small and the aerodynamic moment about the elastic axis is
defined by:
Often simplified forms of the equations are used in order to clarify the roles
of the structural and aerodynamic properties for certain classes of problems.
For the static problem classes introduced in in chapter 1, for example, one may
ignore the terms with time derivatives.
α θ AC
MAC
Kθ@@@@
@@@@
αo
@@@@
@@@@
Zero−Lift
@@@@
EA
Line
U ec
∂M
≤0 (2.13)
∂θ
Equation (2.13) simply says that if there is a positive perturbation in θ, then a
negative (nose-down) opposing moment must be produced. Applying it to the
moment equation leads immediately to:
ecqS
1− CLα ≥ 0 (2.14)
Kθ
from which can be concluded that the configuration is unconditionally stable
if e ≤ 0 (AC downstream of EA), but only conditionally stable if e > 0. The
former is the case for weather vanes, while the latter is often true for realistic
wings. We can also re-arrange (2.14) to obtain:
Kθ
q≤ (2.15)
CLα ecS
AC @@@@
Kθ@@@@
θ @@@@
@@@@
MAC
@@@@
U EA
At equilibrium:
X
M = [(CLα θ + CLδ δ) e + CMacδ δ] cqS − Kθ θ = 0 (2.16)
Noting CLα , CLδ , S, c, Kθ > 0, and CMacδ < 0, there will be a q > 0 at which
L
L(R)
will switch from positive to negative, denoted as qreversal :
CLδ Kθ
qreversal = − (2.19)
CLα CMacδ cS
Above this speed, the loss in lift due to (nose-down) structural deflection
will be greater than the lift gained by (positive) control deflection
28 CHAPTER 2. SIMPLIFIED AEROELASTIC PROBLEMS
Flutter Point
Altitude
Decaying Growing
Oscillation Oscillation
ary
nd
ou
rB
tte
Flu
Speed
addition, one must normally consider several types of flutter modes, for which
variations with fuel, payload and aircraft configuration must be established.
The combination of these factors translates into a large number of flight cases
which must be analysed.
To keep things simple, we start our analysis with the case of “Classical
Flutter” which refers to a class of two-degree of freedom oscillations which are
characteristic of 2D airfoil sections. Our approach will be to assume a solution of
an amplified harmonic form, substitute the assumed solution into the equations
of motion, and then solve for the unknown frequencies and amplification factors.
Our task will then be to determine under which conditions the amplification
factors transition from negative (damped case) to positive (growing case). The
analysis is precise (within the bounds of the assumed simple aerodynamic and
dynamic behaviours) but does not necessarily give a phenomological answer
to the question “What is flutter?”. For that, it is useful to appeal to energy
concepts, which at the very least can account for the existence of the phenomena.
When an aircraft travels in relative motion to air, it is in fact contact with
an energy reservoir. More succinctly, certain motions of the aircraft can extract
work from the airstream. If the speed (and structural integrity) of the aircraft
is maintained, such energy extractions will ultimately show up in the aircraft’s
fuel bill. So how is energy transferred to the structure? In our analysis we
will ignore the drag force, so that the transfer of energy must occur when the
airfoil’s vertical speed vector is aligned with the lift vector. When the lift and
vertical speed have the same sign, work is done on the structure (figure 2.11).
When the lift opposes the motion, energy is extracted (this is similar to what
occurs when pushing in and out of phase on a child’s swing).
If we consider an airfoil which can plunge but not pitch (h 6= 0, θ = 0), the
low-frequency aerodynamic model predicts that the motion is always damped
(check this yourself). The key to classical flutter, therefore, is that the timing
of the motion in θ and h must be such that over one cycle, the lift is on av-
30 CHAPTER 2. SIMPLIFIED AEROELASTIC PROBLEMS
Lift
(L) L(t)
Vertical
Speed (z) z(t)
−
+ +
+ t
− −
erage
R aligned with the airfoil motion, rather than opposing it (in other words
L(t)ż(t)dt > 0). Such a situation is shown in figure 2.11.
Classical flutter with a growing amplitude can therefore occur with the
proper timing of θ and h, which implies the proper timing of aerodynamic forces
with structural motion. However, non-classical flutter modes are also possible.
For example, single-degree-of-freedom flutter can occur when unsteady aerody-
namic forces have a phase different from that applied by the steady aerodynamic
model. A famous example is single-DOF control-surface flutter, which can occur
in transonic conditions. In this case, shock movement as a function of δ is lagged
such that the net aerodynamic damping is negative. Another possibility is the
pitching airfoil at very low speeds, for which wake effects can also introduce
negative damping.
Both of the single-degree-of-freedom cases above highlight the importance of
unsteady aerodynamic effects in determining the timing of the forces and overall
energy balance. Although in many cases the steady aerodynamic model we are
about to use may perform reasonably, it can be insufficient for the conditions of
most relevance for modern aircraft, the transonic regime. We will examine the
reasons behind the limits of the steady aerodynamic model in Chapters 3-6.
L
ec
AC h
EA CG
MAC θ
ab
xθ b
b b
where
m Sθ Kh 0 0 −SCLα
[M ] = [K] = [A0 ] = (2.25)
Sθ Iθ 0 Kθ 0 2SebCLα
This form is typical for the larger systems we will consider later in the course.
To find solutions to the system, we begin by assuming that the 2 DOF response
ˆ pt so that:
of the section, {x} = [h, θ]T , is of the form {x} = {x}e
or
where for the latter q, the dynamic pressure, has been separated out to indicate
its importance as a parameter. For non-trivial {x}, the determinant of the
32 CHAPTER 2. SIMPLIFIED AEROELASTIC PROBLEMS
qf1 qf2 qd
ω σ
1 4 1 3
q q
Figure 2.13: Flutter diagram for 2DOF section with Kθ = 0.05, xθ = 0.2
a4 p 4 + a2 p 2 + a0 = 0
a4 = mIθ − Sθ2
a2 = mKθ + Iθ Kh − (2meb + Sθ )qSCLα
a0 = Kh (Kθ − 2ebqSCLα )
To get an idea of how the motion changes over the operating range of a given
configuration, we can plot the components of p as a function of the dynamic
pressure. Such diagrams are known as “flutter diagrams” and are frequently
used to express the overall characteristics of a system (they can also be plotted
with speed as the ordinate). Note that the imaginary component of p corre-
sponds to the frequency of oscillation, while the real component corresponds
to the growth rate. While plotting we reject p’s with negative frequencies as
unphysical solutions.
2.1. THE TYPICAL SECTION 33
qf qd
ω σ
1 4 2 3
q q
Figure 2.14: Flutter diagram for 2DOF section with Kθ = 1.0, xθ = 0.2
qd
ω σ
1 3
q q
Figure 2.15: Flutter diagram for 2DOF section with Kθ = 1.0, xθ = −0.1
34 CHAPTER 2. SIMPLIFIED AEROELASTIC PROBLEMS
C2 Q2 + C1 Q + C0 < 0 (2.34)
with:
We can find Q at the the flutter boundary using the quadratic formula:
p
−C1 ± C12 − 4C2 C0
Q1 , Q2 = (2.38)
2C2
Requiring Q to be real leads to:
Conditions (2.40) and (2.42) are know as Pine’s conditions. They can be used
to investigate the possibility of flutter for various combinations of geometric,
stiffness, and inertial parameters.
a4 p 4 + a3 p 3 + a2 p 2 + a1 p + a0 = 0 (2.46)
(2.47)
38 CHAPTER 2. SIMPLIFIED AEROELASTIC PROBLEMS
σ
ω qf
q q
Figure 2.16: Flutter diagrams with the steady and low-frequency aerodynamic
models (∗) - steady aerodynamic model (x,+) - low-frequency aerodynamic
model
where
As before, the roots of the characteristic equation define the motion types.
Unfortunately, it is now a full-forth order equation so it can not be so easily
manipulated. We therefore resort to its numerical solution. Some example
results are shown compared to their steady aerodynamic model counterparts in
figure 2.16.
The low-frequency aerodynamic model results show the same initial behaviour
of the mode frequency branches, but these become substantially different as q is
increased. Note that there is no longer an intersection of the frequency branches.
Examining the σ plot, we observe that at small q the motion is damped, as we
might have anticipated from simple reasoning, but then for one oscillatory mode
σ grows rapidly, so that the critical dynamic pressure is more than 30% less than
that computed with the steady aerodynamic model.
2.2. THE SEMI-RIGID WING 39
Sweep is often used in high-speed aircraft, not only to raise the critical Mach
number of a wing, but also to favourably affect its aeroelastic properties. A very
simple model which can be used to illustrate this is the semi-rigid wing. This
model combines bending and torsion with wing sweep as a parameter.
7777
7777 Kγ
7777
7777 U
Λ: Sweep angle
7777
7777 θ: Twist
Kθ7777
7777 γ: Bending angle
7777
7777 Kθ : Torsional stiffness
7777 ycp
7777
7777 Λ
Kγ : Bending stiffness
7777
7777
ecp Iθ : Torsional inertia
7777
7777
Iγ : Bending inertia
7777 Iθγ : Product of inertia
7777 y S: Total wing area
γ θ b: Semi-span
x b ȳ: Coordinate along
elastic axis
x̄: Coordinate normal
to elastic axis
ȳcp , ēcp : Center of pressure
location
The lift is usually assumed to be a function of the time history of the local
effective angle of attack:
L = L(α(t)) (2.56)
Following a procedure similar to that used for the typical section (see the prob-
lems at the end of this section), the divergence speed for the semi-rigid wing
can be derived. Doing so reveals that forward-swept wings (e.g. figure 2.19)
have lower divergence speeds. If aft sweep is used, there will be a point where
the effect of bending cancels the effect of torsion and no divergence is obtained.
This is referred to as the aeroisoclinic condition.
2.3. PROPELLER WHIRL FLUTTER 41
For these cases, it is believed that the whirl flutter mode also induced flutter
in the main wing, leading to the wing’s separation from the aircraft. Photos
of a model of the aircraft before and after wind-tunnel testing are shown in
figure 2.21.
Figure 2.21: L188 model in the NASA Langley Transonic Dynamics Tunnel
Κψ
Κθ
θ y
x ψ Ω
FA
b
x’
z z
y y
Ω Ω
x x
θ Ω
x U
U θ
r −Ωr
x’
p2.4.1 Derive and expression for the qdivergence of the semi-rigid wing (hint:
first use the equations of motion and angle of attack equation to derive an
explicit expression for αd ). Plot the variation of the divergence speed with wing
sweep. Derive an expression for the aeroisoclinic condition.
2.4. PRACTICE PROBLEMS 45
c) Assume an aerodynamic balance (see figure above, right) has been added
to the control surface, such that the net aerodynamic hinge moment is
zero (CHδ = 0). Draw flutter diagrams for the configuration for both
positive and negative values of Sδ (assume Kh 6= 0 and Sδ2 < mIδ ).
p2.4.3 For the typical section with two degrees of freedom (θ, h), under what
conditions is flutter possible if the AC lies ahead of the EA and the EA lies
ahead of the CG?
46 CHAPTER 2. SIMPLIFIED AEROELASTIC PROBLEMS
p2.4.4
z
Kθ
x
θ ψ
y
Kψ
Iy θ̈ + Ix Ωψ̇ + Kθ θ − a b qθ = 0
Iz ψ̈ − Ix Ωθ̇ + Kψ ψ − a b qψ = 0
where a is the lift-curve slope of the engine nacelle, q is the dynamic pressure,
and Ω is the engine rotation speed.
Unsteady Aerodynamics
In the previous chapter, it was demonstrated how energy can be added to the
structural system if we have the correct phase between structural motion and
aerodynamic forces. But what if the aerodynamic forces are themselves not in
phase with the instantaneous position of the aerodynamic surfaces? This is pre-
cisely what occurs when the effects of unsteady aerodynamics are significant. In
fact, not only may the relative phases be large, but the magnitudes of unsteady
forces can also differ considerably from their quasi-steady counterparts. As a
result, neglecting unsteady aerodynamic effects can result in substantial errors
in predictions of aeroelastic behaviour.
There are several physical processes which contribute to unsteady aerody-
namic effects. For airfoils, important examples include the influence of vortical
convection, compressible effects such as time delays due to the finite speed of
sound and shock movement, and viscous effects including flow separation and
boundary-layer dynamics. We will begin this chapter with some qualitative
descriptions of these phenomena. Then we will examine a hierarchy of model
equations which can be used to predict all or some of their behaviour. We will
finish by listing some of the analytic solutions which have been derived for these
equations to date.
47
48 CHAPTER 3. UNSTEADY AERODYNAMICS
Figure 3.1: A flow with compressible effects (e.g. shocks), viscous effects
(e.g. shock-boundary layer interaction, separation), and vorticity dynamics (e.g.
wake and turbulent regions).
Now consider the impulsive start of an airfoil from rest. Initially, the flow
pattern will indeed resemble figure 3.2. Zooming into the trailing edge, however,
reveals some basic problems for a real viscous flow (figure 3.3). At the trailing
edge, the sharp geometry will lead to velocities which tend towards infinity,
implying very low pressures. Meanwhile, at the stagnation point on the rear
3.1. VORTICITY DYNAMICS 49
+
+
−
−
− −
upper surface, the pressure reaches its highest possible value. Therefore the flow
which tracks around the trailing edge towards the stagnation point experiences a
severe adverse (positive) pressure gradient. From boundary layer theory, adverse
pressure gradients in wall-bounded flows tend to induce separation. This can
be seen by considering the boundary-layer streamwise momentum equation (
see equation (3.7) later in this chapter) for a constant viscosity evaluated at the
airfoil surface, where the velocities are zero due to the no-slip condition:
∂ 2 ui ∂pe
µ = (3.1)
∂x21 ∂xi
Where x1 is the wall-normal direction. Thus in a positive pressure gradient,
the second derivative of the profile in the normal direction also tends to be
positive at the wall. This is characteristic of profiles near separation, as shown
in figure 3.4.
Thus there is separation from the sharp trailing edge, leading to the forma-
tion of a vortex, known as the “starting vortex” above the airfoil (figure 3.5).
This is in turn carried downstream with the flow (figure 3.6). Kelvin’s theorem
states that the total vorticity within a closed contour which convects with the
flow remains constant. If we draw this contour around the airfoil alone we re-
cover the circulation of the bound vortex. If we draw it large enough to enclose
both the bound and starting vortices, we see that the total vorticity must be
zero (its value before the impulsive start) and that the starting vortex is equal
and opposite in strength to the bound vortex.
While the starting vortex is in the immediate vicinity of the airfoil, its in-
duced velocity field tends to lower the local angle of attack. The result is a lift
value which is lower that of the final steady-state value (in an incompressible
inviscid flow, the initial value of airfoil lift is roughly half that of its final value).
As the starting vortex convects downstream, its influence wanes so that the final
steady lift value is approached asymptotically.
50 CHAPTER 3. UNSTEADY AERODYNAMICS
x1
u( x1 )
∂ 2 ui
Figure 3.4: A profile with ∂x21
> 0 near the wall
The wake vorticity also induces a velocity field which influences the shape
of the wake itself (remember vorticity is carried with the local velocity). This
leads to complex wake shapes as time evolves, a factor which can considerably
complicate numerical methods which rely on wake tracking.
3.1. VORTICITY DYNAMICS 51
Γ
−Γ
(u+c)local (u−c)local
the airfoil is moved suddenly, the three points shown in the figure will immedi-
ately begin to adjust their local states to the new boundary conditions. Their
transients won’t be complete, however, until the information reaching them from
all other points on the airfoil remains fixed. The speed at which the information
travels depends on the wave direction (±c) and on the local velocity produced
by the flow u. If u is of a comparable size to c, then the point in the middle
will hear from the downstream point at a much later time, resulting in a pro-
longed transient, and a complex local behaviour in time. Note that information
travelling from the airfoil’s wake can experience significant delays, resulting in
significant response time lags.
As the Mach number is increased into the transonic region, shocks will begin
to appear. A typical transonic pressure flow topology, and its associated pressure
distribution is shown in figure 9. It is clear from the pressure distribution that
the position of the shock has a strong influence on the the forces and moments,
and thus the aeroelastic response. At very high excitation frequencies it is
possible to show that the amplitude of shock movement is small, albeit with
significant phase lag. At very low excitation frequencies one can expect a quasi-
static transonic response. In between this range, where the majority of cases
lie, shock dynamics are hard to predict. This is because they are determined
by the time history of the acoustic wave pattern, which is even more complex
when the flow contains embedded regions of supersonic flow. Figure 3.10 shows
the time it takes for a signal travelling forward from the trailing edge of an
airfoil to reach a given streamwise position. In the transonic case (right), the
delay is significantly increased by the presence of the shock, as the information
3.2. COMPRESSIBILITY EFFECTS 53
−Cp
x/c
Shock Wave
Sonic Line
Figure 3.9: A typical transonic flow topology (below) and its associated pressure
distribution (above)
must first travel around the supersonic region before reaching the points ahead
of the shock. The time required for this is a strong function of the shock
geometry, which is a function of the airfoil geometry (e.g. compare supercritical
profiles to conventional ones). If the shock is moving, the transient becomes even
more complex. The difficulty of predicting responses in the transonic regime is
further highlighted by figures 11 and 12. which compare linearised theory to
experiment for an oscillating flap as a function of the flow Mach number. The
′
mean measured pressure coefficient, Cp , along with the in-phase Cp and out-of-
′′
phase Cp measured and computed unsteady pressure perturbations are shown.
Clearly the linearized theory does fairly well until transonic effects begin to
appear M = 0.85, after which it becomes increasingly inaccurate.
54 CHAPTER 3. UNSTEADY AERODYNAMICS
Figure 3.10: Time delays due to the presence of an airfoil (From Tijdeman, NLR
TR 77090U)
Figure 3.11: Pressure distributions for a NACA 64006 with an oscillating flap.
Steady (upper) and real and imaginary components (lower) comparison of exper-
iment (symbols) with linear theory (lines). (From Tijdeman, NLR TR 77090U)
steady flows, the upper surface boundary layer tends to thicken more rapidly
than lower surface boundary layer as the angle of attack is increased. This
results in a reduction in the effective camber of the section. (figure 3.14). This
effect is further enhanced if a control surface is deflected. In unsteady flows, the
decambering effect becomes a function of the boundary-layer dynamics, which
is itself different from the main flow dynamics due to the relatively low-inertia
near-wall flow.
If separation occurs, the forces and moments are much more directly af-
fected. Conventional separation patterns include the leading and trailing edge
separations shown in the left part of figure 3.15, and the transonic ones shown
at right. The latter occur due to the strong adverse gradients encountered when
passing through the shock wave, and can manifest themselves as either bubbles
or complete separated regions. What makes it interesting is that the shock po-
sition is also a function of the separation pattern. These types of problems pose
difficult challenges to even the most advanced computational techniques.
Under certain circumstances, it is possible to have a self-sustained oscillation
which is driven by cycles of separation and re-attachment. A prominent example
56 CHAPTER 3. UNSTEADY AERODYNAMICS
Figure 3.12: Pressure distributions for a NACA 64006 with an oscillating flap.
Steady (upper) and real and imaginary components (lower) comparison of exper-
iment (symbols) with linear theory (lines). (From Tijdeman, NLR TR 77090U)
of this is dynamic stall, where leading edge separation, vortex convection and re-
attachment can provide significant pitching moment excitation. If the structure
is flexible, its motion can in turn drive the dynamic stall cycle. The basic
pattern is illustrated in figures 16 and 17. (Note that the letter indications do
not correspond between the two figures).
Counter vortex
@@@@@
@@@@@
@@@@@
@@@@@@@
@@@@@
@@@@@@@
@@@@@
@@@@@@@
@@@@@@@
δ∗U
δ∗L
Shock
Separation bubble
@@@@@
Leading edge separation
@@@@@
Separated wake
@@@@@@@@@@@@@@@@@@@
@@@@@@@@@@@@@@@@@@@
@@@@@@@@@@@@@@@@@@@
Trailing edge separation
@@@@@@@@@@@@@@@@@@@
@@@@@@@@@@@@@@@@@@@
Figure 3.15: Conventional and transonic separation patterns
58 CHAPTER 3. UNSTEADY AERODYNAMICS
Figure 3.16: Dynamic Stall (1) From McCorskey, Journal of Fluid Engineering
March 1977, 8-39
d
Z Z
W dV + (F~ i − F~ v − ẋW ) · ~n dS = 0 (3.2)
dt
V S
ρ~u 0
ρ
= =
W = ρ~u , F~ i = ρ~u~u + p I , F~ v = τ
=
ρE ρ~uH τ ·~u − ~q
Here W is known as the state vector, and contains the unknown conservative
variables density (ρ), momentum (ρ~u), and total specific energy (ρE), where ~u
is the local fluid velocity and E = ǫ + 21 |~u|2 , the sum of internal and kinetic
energies. The total enthalpy H, is given by H = E + p/ρ, where p is the local
value of pressure.
The inviscid flux vector, F~ i , in combination with the term ẋW , expresses
the transport of the conserved variables across the control surface. It also in-
cludes the effect of the pressure acting on the control surface on the balance of
momentum within the control volume. The viscous flux vector, F~ v , contains
the effects of viscous stresses. These are described by the viscous stress tensor,
60 CHAPTER 3. UNSTEADY AERODYNAMICS
Navier−Stokes Equations
Small−Disturbance Potential
Equation
=
τ , which has components given by:
∂ui ∂uj ∂uk
τij = µ + + λδij (3.3)
∂xj ∂xi ∂xk
where µ and λ are the viscosity coefficients. Normally λ is taken to be 2µ/3. The
viscosity coefficients are functions of temperature, and for most circumstances
their variation is determined experimentally.
To complete the system, a thermodynamic relation is required which relates
the pressure to the remaining flow variables. For many aerodynamic flows, an
ideal gas model may be used. In some specialized cases (e.g. re-entry vehicles)
this assumption is inappropriate, and more elaborate thermodynamic relations
must be employed.
Although a few specialized analytic solutions to the Navier-Stokes equations
exist, numerical techniques must usually be used to solve flow problems of en-
gineering interest. The computational cost of numerical simulations, however,
is strongly affected by the enormous range of length scales present at the high
Reynolds numbers typical of aerodynamic flows. This arises from the need to
represent the effects of turbulence, for which relevant length scales can be several
3.4. MODEL EQUATIONS FOR UNSTEADY FLOWS 61
n
V
V : Control volume
S: Surface of control volume
x
n: Local surface normal vector
x: Local surface velocity vector
dS
orders of magnitude smaller than the global scales of interest for an aeroelastic
problem (airfoil chord, wing span etc.) Various approaches for dealing with this
difficulty are summarized below:
DNS
DNS stands for direct numerical simulation, and refers broadly to computations
which attempt to resolve all the turbulent lengths scales of a flow. These com-
putations must be performed in three spatial dimensions in order to properly
capture the dynamics of turbulence. Although feasible for low-Reynolds-number
flows, the expense of DNS increases rapidly with Reynolds number. For the
foreseeable future, DNS techniques will be impractical for direct application to
aeroelastic problems. They will continue to play an important indirect role,
however, in the the formulation of models of turbulent behaviour, which are
required to exploit the techniques described in the next two sections.
LES
Large Eddy Simulation (LES) techniques make use of a sub-grid scale model
(SGS) to eliminate the smallest turbulence length scales from consideration,
while fully simulating the dynamics of the larger structures. This technique
works well for free turbulent flows, where the dynamics of the smallest scales can
be sufficiently well represented using simplified models. For wall-bounded flows,
however, it can be difficult to model smaller scales accurately, due to the highly
anisotropic nature of turbulent boundary layers. Although the development of
accurate near-wall SGS models is an area of active research, presently DNS-
like resolution is required near the wall in order to obtain accurate results.
Consequently, LES computations remain too expensive to be routinely used for
aeroelastic computations.
F/RANS
The classical approach to dealing with turbulence is based on the concept of
averaging, as introduced by Reynolds in 1895. In summary, one may decompose
62 CHAPTER 3. UNSTEADY AERODYNAMICS
the quantities of interest in the flow, for example the pressure, p(x, t), into mean
and fluctuating components:
′
p(x, t) = p̄(x, t) + p (x, t) (3.4)
where the mean value is given by averaging over a time scale, T :
1 t+T
Z
p̄(x, t) = p(x, t)dt (3.5)
T t
where T must be chosen to be large compared to the time scale of turbulent
fluctuations. Substitution of the mean and fluctuating components of each
variable into the Navier-Stokes equations and averaging over time leads to the
Reynolds-averaged Navier-Stokes equations (RANS) for incompressible flows, or
the Favre-averaged Navier-Stokes equations for compressible flows. These equa-
tions contain additional terms arising from the correlation of the fluctuating
solution components. The correlations represent new unknowns, the behavior
of which must be described by an additional turbulence model.
There are numerous approaches to turbulence modelling, each of which has
different ranges of applicability. Although reasonable results can be obtained for
certain classes of flows, the development of turbulence models remains an active
research area, as obtaining accurate results for arbitrary flow conditions can be
difficult. When applied to unsteady flows, an additional problem arises in the
choice of the averaging period, T . In practice, it may be difficult to separate
the time scales of slowest turbulent fluctuations from those of the aeroelastic-
induced flow unsteadiness. In such cases, the use of the time-averaging approach
becomes questionable, and other modelling procedures must be applied.
In spite of the difficulties mentioned above, the application of the Reynolds
or Favre-averaged Navier-Stokes equations to aeroelastic problems can produce
useful results. However, for realistic three-dimentional aircraft geometries sim-
ulation costs remain relatively high, limiting their repeated use for aeroelastic
design procedures.
where x1 is the direction normal to the surface, and pe is the pressure of the
flow at the edge of the boundary layer, which is that of the external inviscid
flow. The resulting system, which includes the continuity and energy equations,
is solved only within the thin viscous layer, with boundary conditions given by
the no-slip condition at the wall, and the external flow velocities and pressures
on the boundary layer edge. In many cases, the solution to the boundary layer
equations may be found using marching techniques which exploit the parabolic
character of the equations. This makes such techniques very efficient.
As the size of the control volume is arbitrary, the complete volume integral
(including the first term of (3.2)) can only be true if it is true for all points
64 CHAPTER 3. UNSTEADY AERODYNAMICS
within the volume, allowing us to write the differential form of the equations as:
∂W
+ ∇ · F~ i = 0 . (3.9)
∂t
Consider for a moment the continuity and momentum equations:
∂ρ
+ ∇ · (ρ~u) = 0 (3.10)
∂t
∂ρ~u
+ ∇ · (ρ~u ⊗ ~u) + ∇p = 0 (3.11)
∂t
By subtracting the continuity equation from the momentum equation, and using
the identity:
1
~u · ∇~u = ∇|~u|2 + (ω × ~u) (3.12)
2
where ω is the local vorticity vector, ∇ × ~u, we can also also express the mo-
mentum equation as:
∂~u 1 1
+ ∇|~u|2 + (ω × ~u) + ∇p = 0 (3.13)
∂t 2 ρ
Recalling the second law of thermodynamics, which can be expressed in terms
of the enthalpy, h = ǫ + p/ρ, as:
1
dh = T ds + dp (3.14)
ρ
we then can express the momentum in terms of the total enthalpy, H, as:
∂~u
+ ∇H = −(ω × ~u) + T ∇s = 0 (3.15)
∂t
where T is the local temperature, and s is the local entropy. Equation (3.15) is
known as Crocco’s relation and is still exact within the framework of the Euler
equations. Now we assume irrotational, isentropic flow, allowing us to replace
the representation of the velocity field with the gradient of a scalar potential
function, ~u = ∇Φ. Equation (3.15) then becomes:
∂∇Φ
+ ∇H = 0 (3.16)
∂t
or
∂Φ
+ H = const ≡ H∞ (3.17)
∂t
We can use this equation to construct an explicit relationship between density
and the potential function by using the isentropic formula for a perfect gas:
1 1
∂Φ |∇Φ|2
γ−1
ρ h γ−1 1
= = H∞ − − (3.18)
ρ∞ h∞ h∞ ∂t 2
3.4. MODEL EQUATIONS FOR UNSTEADY FLOWS 65
√
Noting that h = Cp T and a = γRT , this can be written as:
1
γ−1
ρ γ−1 2 ∂Φ 2
= 1+ U ∞ − 2 − |∇Φ| (3.19)
ρ∞ 2a2∞ ∂t
Finally, we can also relate the potential function to the density via the continuity
equation:
∂ρ
+ ∇ · (ρ∇Φ) = 0 (3.20)
∂t
Equations (3.19) and (3.20) form a coupled non-linear system for only two un-
knowns, even for three-dimensional problems. These are the full potential equa-
tions, and represent a considerable simplification of the original Euler system.
As a result, numerical schemes for the full potential equations are relatively
inexpensive.
The assumptions used in deriving the full potential equations have two con-
sequences. Firstly, the assumption of isentropic flow limits consideration to
weak shocks, i. e. those with Mach numbers less than 1.3. Secondly, assuming
the flow to be irrotational requires that vortical wakes be explicitly modelled,
as opposed to being “captured” as part of the solution. This can be difficult to
do in three dimensions, especially for situations which involve strong wake-wake
interactions.
Φ = Φ∞ + φ = U∞ x + φ (3.21)
During the derivation, however, we note that that small terms which accompany
(1 − M 2 ) must be retained, as (1 − M 2 ) is itself small in the transonic regime.
This leads to the following non-linear equation for the disturbance potential:
2
∂2φ ∂2φ ∂2φ ∂ 2φ
2 2 ∂φ ∂ φ 1
1 − M∞ − (γ + 1)M∞ b + 2 + 2 = 2 2U∞ + 2 (3.22)
∂x ∂x2 ∂y ∂z a∞ ∂x∂t ∂t
can be done by considering the momentum equation written using (3.21), and
eliminating higher-order terms:
∂φ ∂φ
p − p∞ = −ρ∞ + U∞ (3.23)
∂t ∂x
Note that the equation for pressure is decoupled from (3.22), resulting in further
reductions in computational work. The pressure may be computed as a post-
processing step.
Equation (3.22), also known as the transonic small-disturbance (TSD) equa-
tion, has been widely used in aerodynamic analysis for transonic aeroelasticity,
due to its relatively low computational cost.
Steady flow
Removing the time derivatives, we obtain:
Incompressible flow
Taking the limit of a∞ → ∞ gives us Laplace’s equation:
∇2 φ = 0 (3.26)
In this case, the time derivatives are lost from the equation for the perturbation
potential. The problem remains unsteady, however, as the boundary conditions
on the airfoil surface will be changing in time. Equation (3.26) is linear, which
allows us to use superposition techniques in its solution. If the changing wake
shape is not specified, however, determining its position may still require a
non-linear iteration process.
3.5. ANALYTIC SOLUTIONS 67
• Tijdeman (1977) Local analysis at shock for frozen-Mach flows. Can use
to investigate effects of shock movement
In the chapters following this one we will examine selected analytic solutions
for the simplest of the model equations. These have been chosen in order to
consider the most important of the unsteady effects separately. Each solution
can be used to provide reasonable predictions under a restricted set of conditions,
but their chief advantage is that they help to develop insight into the relevant
physical phenomena. The discussion of prediction methods for combined or
more complex phenomena is deferred to later in the notes, where we will examine
numerical solution methods for unsteady flows.
68 CHAPTER 3. UNSTEADY AERODYNAMICS
• Basic assumptions
• Governing equations
• Geometric simplifications
• Boundary conditions
• Outline of the solution procedure
• Interpretation of results
p3.6.5 Compare the advantages and disadvantages of applying the Euler equa-
tions versus the linear potential equation to the computation of whirl flutter in
the propeller-wing-nacelle configuration shown below. Assume the configuration
is operating a with a freestream Mach number of 0.6.
M = 0.6
φ
z
∂~u ∇p
+ (~u · ∇)~u + =0
∂t ρ
derive the linearised expression for pressure:
∂φ ∂φ
p = p ∞ − ρ∞ + U∞
∂t ∂x
71
72CHAPTER 4. ANALYTIC SOLUTIONS FOR INCOMPRESSIBLE FLOWS
= + +
x
∇2 φ = 0 (4.1)
∂φ ∂φ
p − p∞ = −ρ∞ + U∞ (4.2)
∂t ∂x
Where φ is the perturbation potential and the ∞ denotes freestream conditions.
To be consistent with the linearization of the potential, we will consider only
small angle excursions, and a linearized airfoil geometry.
If we linearize the boundary conditions on the airfoil, the solution of an un-
steady thin-airfoil problem can be broken up into its thickness (zt ), camber(zc ),
and flat-plate (za ) components, as shown in figure 4.2. As thickness and camber
are not usually considered to be functions of time, their contributions can be
obtained from standard steady thin-airfoil theory. Therefore, for the remainder
of this analysis, we will focus on the solution for a moving flat plate. In doing
so we will solve (4.1) and (4.2) subject to the following conditions:
1. Zero perturbation velocity in the far-field
2. Flow tangency on the airfoil surface
3. No pressure jump across the wake
4. Kutta condition using Kelvin’s Theorem
The first of these conditions is simple enough, we require that the perturbation
vanish when considering points far from the airfoil. The second is the standard
no-flow-through-the-airfoil condition familiar from steady aerodynamic theory.
The third is a new condition which arises from having to consider the wake in
the problem, which of course cannot maintain a difference in pressure across
it. The final condition represents an interpretation of the Kutta condition for
unsteady flow, that the vorticity shed at the trailing edge must be equal and
opposite in strength to the change in bound vorticity of the airfoil (see the
previous chapter for a discussion).
In order to solve the problem we will represent the airfoil and its wake by
a sheet of vortices of unknown strength, γ(x, t), lying on the z = 0 plane, as
4.2. FAR-FIELD PERTURBATION 73
x
b
−b
γa(x,t) γw(x,t)
The use of a vortex sheet exploits the linear nature of the governing equations
to simplify the problem. As each of the infinitesimal point vortices in the vortex
sheet satisfies the governing equations, then the entire vortex sheet satisfies the
governing equations, even for arbitrary local strengths. To find the solution,
it thus remains only to find the distribution of vortex strengths which satisfies
the boundary conditions 1-4 above. We will derive mathematical expressions of
these conditions in the following sections.
∞
γ(ξ) x−ξ
Z
w(x, z) = φz (x, z) = − dξ (4.5)
−b 2π (x − ξ)2 + z 2
Both of these components tend to zero in the far-field, as one would expect
considering the behaviour of a single potential-flow vortex.
of the vortex sheet. Taking the limit of the expressions for velocity (4.4) and
(4.5) as z → 0, we have:
Z ∞
γ(ξ) z
u(x, 0) = lim dξ (4.6)
z→0 −b 2π (x − ξ)2 + z 2
and
∞
1 γ(ξ)
Z
w(x, 0) = − dξ (4.7)
2π −b (x − ξ)
Note that the integral for u(x, 0) tends to vanish except where x ≈ ξ.
In order to find an expression for u(x, 0), we restrict the range of integration
to small region near x of width 2ǫ (figure 4.4). Introducing the intermediate
variable ξ ′ = ξ − x:
Z +ǫ
γ(ξ ′ )
z
u(x, 0) = Limz→0 ′ )2 + z 2
dξ ′ (4.8)
−ǫ 2π (ξ
A Taylor series approximation for γ(ξ ′ ) is:
γ(x − ǫ → x + ǫ) = γ(x) + O(ǫ) (4.9)
Integrating and eliminating higher-order terms:
γ(x) + O(ǫ) −1 ǫ
−1 −ǫ
u(x, 0) = Limǫ→0 Limz→0 tan − tan (4.10)
2π z z
We assume ǫ is small, but is still large compared with z → 0. This leads to the
following expression for the velocity above (+) and below (−) the vortex sheet:
γ(x)
u(x, ±0) = ± (4.11)
2
Alternatively we can express the jump across the vortex sheet directly as:
∆u(x, 0) = γ(x) (4.12)
4.4. FLOW TANGENCY 75
n dz(x,t)
dt
n
U
|Vn| − dz(x,t)
motion
|Vn| − U dz(x,t) dt
freestream
dx
we also note that for small angles w ≈ Vn , the total vertical velocity due to
freestream and motion, wF M , may be expressed:
dza (x, t) dza (x, t)
wF M (x, t) ≈ −U∞ − (4.14)
dx dt
This allows us to express the flow tangency condition as:
wF M (x, t) + wγ (x, t) = 0 (4.15)
or, by substituting our previous expressions, as:
Z b Z ∞
dz(x, t) dz(x, t) 1 γa (ξ, t) 1 γw (ξ, t)
−U∞ − − dξ − dξ = 0 (4.16)
dx dt 2π −b x − ξ 2π b x−ξ
where za (x, t) is the specified motion of the airfoil. Note that we have separated
the integration of the velocity induced by the vortex sheet into components
associated with the airfoil and wake.
of the wake in the vertical direction are so small that they can be neglected while
enforcing the zero-pressure-jump condition. This assumption is quite reasonable
when considering the relatively high forward velocities characteristic of most
aircraft wings.
We can construct an expression for the pressure jump across the wake using
(4.2):
∂∆φ ∂∆φ
pu − pl = −ρ∞ + U∞ (4.17)
∂t ∂x
(4.18)
where ∆φ = φu − φl . This equation can be re-expressed in terms of γ if we note
that:
∂∆φ(x)
= ∆u(x) = γ(x) (4.19)
∂x
Also, for x < −b, ∆φ = 0, so that for x > −b:
Z x Z x
∂∆φ(ξ)
∆φ(x) = dξ = γ(ξ)dξ (4.20)
−b ∂x −b
γw(b,t)
CCCCCC
CCCCCC
CCCCCC
CCCCCC
dx = U dt
Figure 4.6:
iω
γ̂w (b) = − Γ̂ (4.27)
U∞
iω iω
γ̂w (x) = − Γ̂e U∞ (b−x) (4.28)
U∞
Now we substitute these expressions into the wake pressure condition pu −pl = 0.
The second term in (4.21) can be expressed as:
x Z x
∂ dΓ ∂
Z
γ(ξ)dξ = + γw (ξ)dξ (4.29)
∂t −b dt ∂t b
Z x
= iω Γ̂ + iω γ̂w (ξ)dξ eiωt (4.30)
b
78CHAPTER 4. ANALYTIC SOLUTIONS FOR INCOMPRESSIBLE FLOWS
G(k)
0.5 1.0
0.0 F(k)
k= 1.0 k=0
0.6
0.05
−0.2 0.4 0.1
0.2
Integration and substitution into the expression for the pressure change across
the wake shows that the wake pressure condition is automatically satisfied.
Now we are left with just the expression for flow tangency (4.16), augmented
by an explicit expression for the wake vorticity distribution (4.28) for harmonic
flows. (4.16) evaluated for harmonic solutions can be combined with (4.28) to
obtain the following expression for flow tangency on a harmonically oscillating
airfoil:
b−ξ
1 b
γ̂a (ξ) iω eiω( U∞ )
∞
Z Z
− dξ + Γ̂ dξ (4.32)
2π −b x−ξ 2πU∞ b x−ξ
∂ ẑa (x)
= iω ẑa (x) + U∞
∂x
where za (x, t) = ẑa (x)eiωt . This integral equation has been solved analytically
for γ̂a (x) Theodorsen (1934) and Schwarz (1940), with ẑa (x) defined for a har-
monically plunging and pitching flat plate. Solutions of the integral equation
are usually expressed in terms of Theodorsen’s function, C(k):
(2)
H1 (k)
C(k) = F (k) + iG(k) = (2) (2)
(4.33)
H1 (k) + iH0 (k)
(2)
where Hn are nth-order Hankel functions of the second kind. Theodorsen’s
function is also known as the circulation function. The parameter k ≡ (ωb)/U∞ ,
is referred to as the reduced frequency. It expresses the ratio of periods for
oscillation and (wake) convection. The variation of the components of C(k) as
a function of k is shown diagrammatically in figure 4.7.
4.8. HARMONIC PITCH AND PLUNGE SOLUTIONS 79
−b b
x
ab
θ h
3 2
Lh’ Mh’
Lh’’ Mh’’
2
1.5
1
0
1
-1
-2 0.5
-3
0
-4
0 0.5 1 1.5 2 0 0.5 1 1.5 2
k k
Figure 4.9: Unsteady lift and moment coefficients for a flat plate airfoil under-
going pure harmonic plunge
4 1.5
Lθ’ Mθ’
Lθ’’ Mθ’’
1
3
0.5
2 0
1 -0.5
-1
0
-1.5
-1 -2
0 0.5 1 1.5 2 0 0.5 1 1.5 2
k k
Figure 4.10: Unsteady lift and moment coefficients for a flat plate airfoil under-
going pure harmonic pitch
6 10
∆Cp’ ∆Cp’’
4 8
h/b h/b
2 k=0.2 6
0 1 x
0 b 4 k=1.0
−2 0.6
0.6 2 0.2
−4 1.0 x
−1 0 1 b
Figure 4.11: Pressure perturbations for a flat plate airfoil undergoing pure har-
monic plunge
k=1.0
8 8
∆Cp’ ∆Cp’’
6 6
θ θ 0.6
4 4
Figure 4.12: Pressure perturbations for a flat plate airfoil undergoing pure har-
monic pitch
expressed as:
The real and imaginary perturbations to the change in pressure across the plate,
∆Cp′ and ∆Cp′′ are shown plotted as a function of chordwise position in figures
4.11 and 4.12. For the plunge case, real component ∆Cp′ corresponds to
the perturbation obtained when the plate is at its most downward position,
while ∆Cp′′ corresponds to the mid-position while moving downwards. Each
distribution shows the characteristic leading-edge singularity of flat plate steady
solutions. Examination of the pure pitch perturbations reveals that they only
resemble the steady flat plate pressure distribution at low k.
82CHAPTER 4. ANALYTIC SOLUTIONS FOR INCOMPRESSIBLE FLOWS
t* < 0 t* > 0
1.0 φ(t*)
0.8
0.6
0.4
0.2
t*
0 4 8 12 16 20
w 34 c = 0 t<0 (4.43)
−U∞ αo t>0 (4.44)
where
∞
1 C(k) ikt∗ U∞ t
Z
∗
φ(t ) = e dk ; t∗ = (4.46)
2π −∞ ik b
In incompressible flow, the lift perturbation due to an indicial plunge acts at the
quarter chord, so there is no perturbation to the moment about c/4. Wagner’s
indicial lift function can be approximated using:
∗ ∗
φ(t∗ ) ≈ 1 − 0.165e−0.0455t − 0.355e−0.3t (4.47)
4.10. VERTICAL STEP GUST RESPONSE (KÜSSNER’S FUNCTION) 83
wG = 0 x > U∞ t − b (4.48)
wo x < U∞ t − b (4.49)
wG
ψ(t∗ ):
As for the indicial plunge case, ψ(t∗ ) can be derived via a Fourier transform of
the solution to a sinusoidal gust problem. It has the form shown in figure 4.15.
Once again in incompressible flow the lift response acts at the quarter chord, so
there is no perturbation to the moment about c/4. Küssner’s step gust response
function can be approximated by:
∗ ∗
ψ(t∗ ) ≈ 1 − 0.5e−0.13t − 0.5e−t (4.51)
Note that the non-dimensional time, t∗ , can also be interpreted as the number
of half-chords travelled after the leading edge encounters the gust.
84CHAPTER 4. ANALYTIC SOLUTIONS FOR INCOMPRESSIBLE FLOWS
1.0 φ(t*)
0.8
0.6 ψ(t*)
0.4
0.2
t*
0 4 8 12 16 20
γa(x,t) γw(x,t)
x
−b b
x=b
e
γ (b,t)
CCCCCC
w
CCCCCC
CCCCCC
CCCCCC P
dx = U dt
2 ∂2φ ∂2φ
(1 − M∞ ) + 2 =0 (5.1)
∂x2 ∂y
Comparing (5.1) to the linear potential equation for incompressible flow reveals
that solutions to the former may be obtained from those of the latter if a simple
coordinate transformation is applied. Specifically, The solution of (5.1) can be
related to that of Laplace’s equation for φ̄(ξ,
pη) = βφ(x, y) in the transformed
coordinates ξ = x , η = βy; where β = 1 − M∞ 2 . Using this concept, it
2u 2 ∂ φ̄ Cpo
Cp = − =− = (5.2)
U∞ U∞ β ∂x β
where Cpo is the pressure coefficient from the solution to Laplace’s equation
(i.e. the incompressible problem). The behaviour of the corrected lift curve
slope over the complete Mach number range is illustrated in figure 5.1. The
correction was first derived by Prandtl in in the early 1920’s and more formally
87
88CHAPTER 5. ANALYTIC SOLUTIONS FOR SUBSONIC COMPRESSIBLE FLOWS
2π
Clα 1−Μ2
4
2π Μ2−1
k4
Lh = −k 2 + C(k) [2ik] + M 2 logM − 2ik 3 C(k) − 2k 2 C(k)2 (5.3)
2
Some results this approximation at a Mach number or 0.7 are shown compared
to incompressible results in figures 5.2 and 5.3. The compressible effects can be
seen to increase as a function of k. You can further investigate the influence of
Mach number and the centre of rotation on your own by visiting:
https://aerodynamics.lr.tudelft.nl/cgi-bin/lpfp
p′ = ρ∞ a∞ wo (5.7)
where wo is the speed to which the piston is accelerated to, and a∞ is the
speed of sound in the fluid. This perturbation appears anti-symmetrically on
the upper and lower surfaces so that ∆p = 2ρ∞ a∞ wo . The initial lift coefficient
of the wing is then:
These expressions provide a useful model for oscillations with very high frequen-
cies at both subsonic and supersonic Mach numbers. Formally, the theory is
accurate at large values of the product of Mach number and reduced frequency,
M k.
90CHAPTER 5. ANALYTIC SOLUTIONS FOR SUBSONIC COMPRESSIBLE FLOWS
0.5 3
0
2.5
-0.5
2
-1
Lh’’
Lh’
-1.5 1.5
-2
1
-2.5
0.5
-3
-3.5 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
k k
2 0
1.8
1.6 -0.05
1.4
1.2 -0.1
Mh’’
Mh’
1
0.8 -0.15
0.6
0.4 -0.2
0.2
0 -0.25
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
k k
Figure 5.2: Plunging case - comparison of M=0.7 (solid) and M=0 (dashed)
responses (a=-1/2).
5.3. PISTON THEORY 91
3 5
4.5
2.5 4
3.5
2 3
2.5
Lθ’’
Lθ’
1.5
2
1 1.5
1
0.5 0.5
0
0 -0.5
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
k k
1.2 0
1 -0.5
0.8 -1
Mθ’’
Mθ’
0.6 -1.5
0.4 -2
0.2 -2.5
0 -3
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
k k
Figure 5.3: Pitching case - comparison of M=0.7 (solid) and M=0 (dashed)
responses (a=-1/2).
92CHAPTER 5. ANALYTIC SOLUTIONS FOR SUBSONIC COMPRESSIBLE FLOWS
@@@@@
@@@@@
@@@@@
@@@@@
Expansion Wave
Compression Wave
wo
2 ḣo
L = πρU∞ Φc (t∗ ) (5.9)
U∞
2 ḣo
M = πρU∞ 2b ΦcM (t∗ ) (5.10)
U∞
After the initial piston theory response, there is an increasingly slow climb
to the steady state value as the Mach number is increased. This is due to
the increasing upstream propagation times of information describing the wake’s
dynamics. For cases where linearised theory is adequate, the final large-time
asymptotic value can be estimated using the Prandtl-Glauert factor √ 1 2 .
1−M∞
5.4. INDICIAL RESPONSES FOR M < 1 93
M = 0.7
1.2
φc(t*) M = 0.5
1.0
M=0
0.8
0.6
0.4 t*
0 4 8 12 16 20 24 28
(φcM)c/4(t*)
−0.3
M = 0.5
−0.2
−0.1
M = 0.7
M=0
0
t*
0 4 8 12
p5.5.12 Using piston theory and the Prandtl-Glauert correction, sketch the
M=0.5 indicial plunge function for a finite wing with an incompressible lift-curve
slope of 4. Give values for t = 0 and t = ∞.
Chapter 6
In this chapter we will consider solutions to the linear potential equation for
supersonic flows. We will start by considering an isolated airfoil (figure 6.1).
This case is quite different from the solutions we have considered up to now, as
convecting vorticity in the wake has no influence on the airfoil’s forces due to
the former’s restricted domain of influence. This allows us to concentrate on the
analysis of time-delay effects. Note that in more general supersonic problems,
such as those with highly swept or multiple lifting surfaces, convecting vorticity
must again be taken into account. This will be discussed in more detail at the
end of the chapter.
Although the unsteady supersonic airfoil problem was first solved by von
Borbely in 1942, we will consider the later solution by Garrick and Rubinow
(1946), because of its more physical interpretation. As for the incompressible
case, we will exploit the linearity of the governing equations and solve the prob-
lem using a distribution of potential flow singularities. Once again, since each
of the singularities is a solution to the governing equations, to solve the prob-
lem we only need to find the combination of their strengths which satisfies the
boundary conditions.
The boundary conditions we need to consider are simpler than those of the
incompressible case. Firstly, as the wake has no influence on the isolated airfoil,
we may ignore Kelvin’s theorem and the zero-pressure-change-across-the-wake
condition. Secondly, although in the far-field the disturbance is finite and a
function of the solution, we need not take it explicitly into account as it will be
automatically provided once the singularity strengths are determined (this was
also true for the incompressible case). This leaves us with only the flow-tangency
condition to enforce.
Note that although the perturbation in the far-field is non-zero, for an earth-
fixed observer, it is only non-zero for a finite time. This can be appreciated
by imagining the airfoil in figure 6.1 passing over you. Initially you will feel
95
96 CHAPTER 6. ANALYTIC SOLUTIONS FOR SUPERSONIC FLOWS
µ
x
−b b
the compression after you are passed by the wave emanating from the leading
edge, and then expansion back to ambient pressure after you are passed by the
wave emanating from the trailing edge. If you are sufficiently far away, the
perturbation you feel due to the wake will be negligible.
As for the incompressible case, the airfoil will be located in the z = 0 plane, so
that the linearized flow tangency condition can be expressed:
∂φ
= wF M (x, t) (6.3)
∂z z=0
A(ξ, ζ, T )
φs (x, z, t) = p (6.4)
a (t − T ) − [(x − ξ) − U (t − T )]2 − (z − ζ)2
2 2
Disturbed
Region
(x,z)
a(t−T)
(ξ,ζ)
U(t−T)
(x−ξ)
the term “pulse disturbance front” to refer to the suface which separates points
which are aware of the pulse from those who are not. The pulse disturbance
front is a sphere (a cylinder in 2D problems) which expands with the speed of
sound a. Since the airfoil is moving, however, the sphere is also convected down-
stream with a speed U . At a certain time, the disturbance front will expand
and convect enough to reach (x, z). After this time (x, z) is influenced by the
pulse. Since U > a, however, after a certain time the sphere will be swept past
(x, z). After this time the effect of the pulse at (x, z) disappears. Note that
points above the dashed line in figure 6.2 never feel the effect of the pulse. The
dashed line therefore defines the limit of the pulse’s zone of influence. The cone
obtained by revolving the dashed line around the x-axis is called the Mach cone
(this becomes a wedge in 2D problems).
As explained above, the the effects of our component solutions are only felt over
limited extents of time and space. Thus, in order to complete step 1, we must
first establish formal limits of integration for the potential at (x, z).
φ (x,0,t)
x
−b b
(x,z)
aτ2
a τ1
(ξ,ζ)
Uτ1
U τ2
Mathematically, the interval within which the influence of the pulse is felt
at (x, z) corresponds to when the expression for φs (x, z, t) is real. Defining
τ = t − T , we can determine the times τ1 , τ2 between which the pulse is felt at
6.3. SPATIAL INTEGRATION LIMITS 99
2
µ= tan−1 [1/(M −1) ]
1/2
Forward−looking
Mach cone
(x,z)
−b ξ1
With a little bit of geometry, we can establish the spatial limits of integration
for sources on the airfoil which can influence the point (x, z). If (x, z) is above
the airfoil, then the limits are:
p
ξ = −b to ξ1 = x − z M 2 − 1 (6.6)
if (x, z) is below the airfoil, then the limits are:
p
ξ = −b to ξ1 = x + z M 2 − 1 (6.7)
which, noting that τ1 , τ2 are the roots of the denominator of the source function,
can also be expressed:
Z ξ1 Z τ2
1 Â(ξ, 0, t − τ )
φ(x, z, t) = √ p dτ dξ (6.9)
2 2
U − a −b τ1 (τ − τ1 )(τ2 − τ )
√
where we have chosen to re-express the integral in terms of  = U 2 − a2 A. If
we use an angle transformation defined in terms of an angle γ as :
(τ2 − τ1 ) (τ2 + τ1 )
τ= cosγ + (6.10)
2 2
then after changing variables from τ to γ and substituting for dτ the denomi-
nator cancels perfectly, allowing us to simplify the expression to:
φ(x, z, t)
ξ1 π
1 (τ2 − τ1 ) (τ2 + τ1 )
Z Z
= √ Â ξ, 0, t − cos γ − dγdξ(6.11)
U 2 − a2 −b 0 2 2
ξ1 Z π
1
Z
= √ Â(ξ, 0, T ) dγdξ (6.12)
U − a2
2
−b 0
where
p
M (x − ξ) (x − ξ)2 − (M 2 − 1)z 2
T =t− − cosγ (6.13)
a(M 2 − 1) a(M 2 − 1)
xo
This expression can be used to find the force response of the airfoil to any
given wF M , by integration and substitution into the linearised pressure relation.
In general it is integrated numerically, but it is also possible to integrate it
analytically for the case of harmonic motion, as discussed in the next section.
After some manipulation the potential at the airfoil can then be expressed as:
Z xh
2b i
φ(x, t) = ± √ U θ + ḣ + 2b(ξ − xo )θ̇ I(ξ, x, M, ω̄)dξ (6.21)
M2 − 1 0
h ω̄ i ωb 2M 2
where: I(ξ, x, M, ω̄) = e−iω̄(x−ξ) Jo (x − ξ) and ω̄ = (6.22)
M U M2 − 1
The details of the analytic procedure can be found in NACA report R846 by
Garrick and Rubinow, or in the text by Bisplinghoff. The harmonic case can
102 CHAPTER 6. ANALYTIC SOLUTIONS FOR SUPERSONIC FLOWS
also be derived very efficiently using a Laplace transform, as shown in the text
by Fung.
The explicit expression for the potential can be differentiated with respect to
time and space, and the results substituted into the linearised pressure relation
to determine the pressure perturbation. This can be integrated along the chord
to arrive at expressions for the airfoil’s lift and moment, which are usually
written compactly as:
" #
2 2 iωt ĥ
L = 4ρbU k e (L1 + iL2 ) + θ̂(L3 + iL4 ) (6.23)
b
" #
2 2 2 iωt ĥ
M = −4ρb U k e (M1 + iM2 ) + θ̂(M3 + iM4 ) (6.24)
b
(6.25)
where L1..4 = L1..4 (M, ω̄) and M1..4 = M1..4 (M, ω̄) are coefficients dependent
on ω̄ and M . Alternatively, the lift and moment can be expressed in terms
of the Lh , Mh , Lθ , and Mθ coefficients introduced previously, which are then
functions of k and M (note that ω̄ is essentially a modified reduced-frequency
parameter). Figures 6.7 and 6.8 show the variation of these coefficients with
Mach number and reduced frequency. Both the real and imaginary components
of the response can be dramatically affected by these parameters. You are free
to experiment for yourself using the program at:
https://aerodynamics.lr.tudelft.nl/cgi-bin/lpfp
0.5 3
0 2.5
-0.5 2
Lh’’
Lh’
-1 1.5
-1.5 1
-2 0.5
-2.5 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
k k
1.2 1.4
1 1.2
0.8 1
0.6
0.8
Mh’’
Mh’
0.4
0.6
0.2
0 0.4
-0.2 0.2
-0.4 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
k k
Figure 6.7: Plunging case - comparison of M=2.0 (solid), M=4.0 (dotted) and
M=0 (dashed) responses (a=-1/2). (Supersonic values multiplied by π)
104 CHAPTER 6. ANALYTIC SOLUTIONS FOR SUPERSONIC FLOWS
2.5 3
2.5
2
2
1.5 1.5
Lt’’
Lt’
1 1
0.5
0.5
0
0 -0.5
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
k k
1.2 2
1 1.5
1
0.8
0.5
Mt’’
Mt’
0.6
0
0.4
-0.5
0.2 -1
0 -1.5
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
k k
Figure 6.8: Pitching case - comparison of M=2.0 (solid), M=4.0 (dotted) and
M=0 (dashed) responses (a=-1/2). (Supersonic values multiplied by π)
P P
p6.9.13 How long does a point (x, z) feel a pulse at (ξ, ζ, T ) if (ξ, ζ) is on the
forward-looking Mach cone of (x, z) ?
p6.9.14 Briefly outline how you could use linearised theory to determine the
lift response of an airfoil to a sharp-edged gust in supersonic flow. Use equations
from the notes.
√
p6.9.16 A wing-tail combination is travelling at a Mach number of M = 5,
when it is subjected to a step plunge with amplitude ḣo at t = 0.
e
d
c
P
U f
CG
c/ 4
x
Arbitrary Unsteady
Aerodynamic Responses
107
108CHAPTER 7. ARBITRARY UNSTEADY AERODYNAMIC RESPONSES
∆τ
α(t)
α(t) ∆τ ::::
::::::::::::: ::::
::::::::::::: ::::
::::::::::::: ::::
::::::::::::: ::::
:::: t
t ::::
::::
τ ::::
∆α τ δα :::: α(t)
∆α = dα ∆τ ::::
Step
:::::::::::::
dτ Impulse
::::
Input
:::::::::::::
Input
::::
::::::::::::: t :::: t
:::::::::::::
dα A(t−τ) ∆τ
∆L dτ δL
Impulse α(t) h(t−τ) ∆τ
Step
Response Response
t t
ρU 2 S ∞
Z
L(t) = α(ω)CLα (ω)eiωt dω (7.6)
4π −∞
We can think of taking the direct transform of α(t) as determining the harmonic
components of α, and the inverse transform as a summation of each component
multiplied by the CLα at the corresponding frequency. Although we will consider
analytic expressions here, in practice the procedure is often performed using fast
Fourier transform (FFT) techniques.
where k = ωb/U∞, and w 43 c (t) can be computed for any combination of pitch,
θ, and plunge, h . If we focus on the second term, the frequency-domain repre-
sentation of its input is:
Z ∞
∗
W (k) = w 34 c (t)e−ikt dt∗ (7.8)
−∞
Since our aerodynamic theory is linear, we can replace the second term in
the lift expression with the sum of the responses to each one of the sinusoids:
Z ∞
∗
L(t) = πρb2 [ḧ(t) + U∞ θ̇(t) − baθ̈(t)] + ρU∞ b C(k) W (k) eikt dk (7.9)
−∞
This allows us to use the analytic expressions for C(k) in the computation of
arbitrary lift responses.
Since for a → 0
∞
a
Z
dω = π (7.13)
−∞ a2 + ω2
1
πδ(ω) + (7.14)
iω
To speed manipulation, however, we will consider the step function to be defined
by (7.11) where a is very small but finite, so that its approximate transform is:
1
(7.15)
iω
This approximation allows us to express the indicial plunge lift response of a
flat plate in incompressible flow as:
∞
1 C(k) ikt∗
Z
φ(t∗ ) = e dk (7.16)
2π −∞ ik
where we have changed to the non-dimensional variables k and t∗ , and have also
ignored the initial infinite response peak arising from the apparent mass terms.
7.3. STATE-SPACE REPRESENTATIONS 111
0.165k 0.335k
C(k) ≈ 1 − − (7.18)
k − 0.0455i k − 0.3i
Note that this approximation contains spurious poles at k = 0.0455i and k =
0.3i.
Compact Models of
Structural Systems
113
114 CHAPTER 8. COMPACT MODELS OF STRUCTURAL SYSTEMS
θ
3
L2
L3
2
(Xo,Yo)
L1
1
In figure 8.1 (left) the triangle is visualised as nodes connected by rigid bars.
In order to describe its motion, we use the particle coordinates of the nodes as
unknowns, and Newton’s second law to describe their evolution. This leads to
six equations of motion, as there are six particle coordinates:
Since the bars are rigid, we also need to apply three constraints between the
nodes of the form:
the equations of motion to three. In many cases, however, constraints are non-
holonomic and can be difficult to apply (e.g. inequalities forcing a fuel particle
to remain in the fuel tank). By choosing our coordinates carefully, we can
apply the constraints implicitly, thereby reducing the degrees of freedom while
avoiding the need to deal with difficult constraints directly.
Using the chain rule, one may express a small change in a particle coordinate,
δrp , as:
N
X ∂rp
δrp = δqi (8.4)
i=1
∂qi
Similarly:
N
X ∂rp ∂ ṙp ∂rp
ṙp = q̇i from which: = (8.5)
i=1
∂qi ∂ q̇i ∂qi
We may write the virtual work expression for the system of particles as:
M
X
(mp r̈p − Fp ) · δrp = 0 (8.7)
p=1
N X M
X d ∂ ṙp ∂ ṙp
= (mp ṙp · ) − mp ṙp · ( ) δqi (8.10)
i=1 p=1
dt ∂ q̇i ∂qi
116 CHAPTER 8. COMPACT MODELS OF STRUCTURAL SYSTEMS
N X M
X d ∂ ∂ mp
= − ṙp · ṙp δqi (8.11)
i=1 p=1
dt ∂ q̇i ∂qi 2
N
X d ∂T ∂T
= − δqi (8.12)
i=1
dt ∂ q̇i ∂qi
where
M
X mp
T = ṙp · ṙp (8.13)
p=1
2
is the total kinetic energy of the system. So the first term of the virtual work
expression for the system can now be expressed entirely in terms of the gener-
alised coordinates as the kinetic energy of the system is a scalar that does not
depend on what coordinates it is expressed in. The second term of the virtual
work expression for the system can be written:
M M N N
X X X ∂rp X
Fp · δrp = Fp · δqi = Q̃δqi ≡ δW (8.14)
p=1 p=1 i=1
∂qi i=1
Substituting into the virtual work expression, and noting that by definition δqi
are arbitrary and independent, we can write:
d ∂T ∂T ∂U ∂D
− + + = Qi , i = 1, N (8.17)
dt ∂ q̇i ∂qi ∂qi ∂ q̇i
where:
T : Kinetic energy
U: Potential (strain) energy
D: Rayleigh’s dissipation function
qi : Generalised coordinate
Q: Remaining generalised forces
Equations (8.17) are known as Lagrange’s equations of motion.
8.3. LAGRANGE’S EQUATIONS FOR A FLEXIBLE AIRCRAFT 117
r=s+d (8.18)
r′ = r′o + r (8.19)
(8.20)
from which
dr
ṙ′ = ṙ′o + ω × r + (8.21)
dt
x, y, z : coordinates in local frame
x′ , y ′ , z ′ : coordinates in space-fixed frame
r′o : vector giving position of the local coordinate origin
r′ : vector giving position of an element in space-fixed frame
r: vector giving position of an element in local frame
s: undeflected position vector of an element in local frame
d: deformation vector of an element in local frame
ω: angular velocity of local frame
r
x
y
z
r’ r’
o
x’
y’
z’
Where S is the surface of the aircraft and F is the force acting on the surface.
Lagrange’s Equations
Lumped-parameter Rayleigh-Ritz
method method
? ?
System of N exact System of N approximate
equations for the equations which converge to
lumped parameters an exact structural represen-
which approximate the tation as N → ∞
system
l 2
m dw
Z
T = dy (8.26)
0 2 dt
l 2
EI d2 w
Z
U = dy w = w(y, t) = bending deflection (8.27)
0 2 dy 2
l 2
I dθ
Z
T = dy (8.28)
0 2 dt
l 2
GJ dθ
Z
U = dy θ = θ(y, t) = torsional deflection (8.29)
0 2 dy
120 CHAPTER 8. COMPACT MODELS OF STRUCTURAL SYSTEMS
h
α
1. Express the total kinetic, potential, and dissipative energies of the system
in terms of the generalised coordinates
1
T = [m1 ẋ21 + m2 ẋ22 ] (Kinetic) (8.30)
2
1
U = [k1 x21 + k2 (x1 − x2 )2 ] (Potential) (8.31)
2
1
D = [c1 ẋ21 + c2 (ẋ1 − ẋ2 )2 ] (Dissipative) (8.32)
2
8.5. LUMPED PARAMETER METHOD 121
9999 k1 k2
9999
9999
9999
9999
9999
F2(t)
m2
9999
m1
9999
9999
9999
9999 c1 c2
x1 x2
Then we need to consider the virtual work done by the force Fp . From the
expression for δW on slide 8-7, we can express the generalised force as:
M
X ∂rp
Qi = Fp (8.33)
p=1
∂qi
While for i = 2:
You can verify these equations using a direct application of Newton’s second
law.
122 CHAPTER 8. COMPACT MODELS OF STRUCTURAL SYSTEMS
γ1 γ2
1
0
0
1
0
1
0
1
0
1
0
1
0
1 w(y,t)
0
1
0
1 y y
0
1 y
0
1
0
1
0
1
In this case, the deflections are expressed in terms of a finite set of assumed
modes, the amplitudes of which are the generalised coordinates:
N
X
w(y, t) = γi (y)qi (t) (8.37)
i=1
where w(y, t) is the local deflection and γi (y) is the assumed mode. To derive
the equations of motion:
2. Express the total kinetic, potential, and dissipative energies of the system
in terms of the assumed modes and generalised coordinates
This system can be considerably simplified if the chosen modes have the property
of orthogonality:
Z l
γi (y) γj (y) dy 6= 0 for i = j (8.39)
0
= 0 for i 6= j (8.40)
9999
9999
9999
9999
9999
9999
w(y,t)
9999 y
9999
9999
9999
9999
l
9999
Figure 8.6: A Bending Wing
The kinetic and strain energy of the beam are given by:
1 l
Z
T = m(y)ẇ2 (y, t)dy (Kinetic) (8.42)
2 0
1 l
Z
U= EI(y)(w′′ )2 (y, t)dy (Strain) (8.43)
2 0
d
Here ˙ denotes the time derivative, and ′ denotes dy . The above equations can
be re-expressed in terms of the assumed modes as:
N N
1 l
Z X X
T = m(y)( γi (y)q̇i )( γi (y)q̇i )dy (8.44)
2 0 i=1 i=1
N N Z l
1 XX
= q̇i q̇j m(y)γi (y)γj (y)dy (8.45)
2 i=1 j=1 0
N N l
1 XX
Z
= Mij q̇i q̇j ; where Mij = m(y)γi (y)γj (y)dy (8.46)
2 i=1 j=1 0
Finally, we define the ith generalised force by considering the work done by
L(y, t) for a generalised displacement δqi :
Z l
Qi (t)δqi (t) = (L(y, t)δw(t))dy (8.48)
0
Z l
= L(y, t)γi (y)δqi (t)dy (8.49)
0
or:
Z l
Qi (t) = L(y, t)γi (y)dy (8.50)
0
Substituting these expressions into Lagrange’s equation yields the following sys-
tem of N equations:
N
X N
X
Mij q̈j + Kij qj = Qi (t) (8.51)
j=1 j=1
::::
!!!!!!!!!!! k 1
::::
!!!!!!!!!!!
:::: !!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!!
:::: !!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
θ1
::::
!!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
::::
!!!!!!!!!!!
:::: !!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!!
:::: !!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!! i1
!!!!!!!!!!
!!!!!!!!!! k2
!!!!!!!!!!
!!!!!!!!!! !!!!!!!!!!θ2
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!! !!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!! !!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
i2
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
!!!!!!!!!!
Figure 8.7: Lumped-parameter model
::::
!!!!!!!!!!!
::::
!!!!!!!!!!!
::::
!!!!!!!!!!!
::::
y=0
!!!!!!!!!!!
::::
!!!!!!!!!!!
::::
θ(0) = 0
!!!!!!!!!!! θ(y)
::::
!!!!!!!!!!!
!!!
!!!
y = !!!
1
θ’(1) = 0
Flutter Calculation
Methods
where:
[M ] : Mass matrix
[D] : Structural damping matrix
[K] : Stiffness matrix
{q} : Vector of Generalised Coordinates
{QA } : Generalised aerodynamic forces
127
128 CHAPTER 9. FLUTTER CALCULATION METHODS
U
Figure 9.1: Brute-force amplitude analysis
(e.g. the flight speed U ) and the frequency, ω, of F (t). Typically this results
in system response amplitude plots such as that shown in figure 9.1. Although
this procedure is analogous to those used in flight flutter testing, it is quite in-
efficient from the numerical point of view as it requires harmonic time-domain
simulations for each point in the U − ω plane.
Fortunately there are a number of more efficient methods, the most common
of which will be be described in this chapter. Each of these avoids the costs of
the brute-force approach by directly generating an approximate flutter diagram.
The first two methods considered, the k method and p-k method, assume that
only harmonic aerodynamic data are available. This is sufficient for the predic-
tion of flutter boundaries, as we know that at each flutter point, the response
will be purely harmonic. The last method we consider, the p method, is more
general in that it can also predict the response away from the flutter boundary,
but requires enough aerodynamic data to construct an approximate model for
QA as a function of p.
9.1 Non-dimensionalisation
For the remainder of this chapter, we will employ a non-dimensional expression
for p:
b
p= (σ + iω) = δ + ik (9.2)
U
9.2. K METHOD 129
Ut
and assume solutions of the form {q} = {q̂}ep b . The dynamic equations of
motion can then be written as:
2
U 2 U 1
[M ]p + [D]p + [K] {q̂} = ρU 2 [A(p)]{q̂} (9.3)
b2 b 2
Where A(p) is also a function of the Mach number M∞ and the Reynolds number
Re. In general, A(p) can be non-linear, and need not be expressed in terms of
explicit functions.
9.2 k method
The k method arose during the 1940’s as a technique to efficiently make use of
pure harmonic aerodynamic data during flutter analysis. The idea is simply to
add a ficticious structural damping term to the equations, and then to determine
its required value in order to keep the response purely harmonic. The flutter
boundary is then determined by finding the point at which the required ficticious
damping becomes zero. Note that since the response is always kept purely
harmonic, the use of harmonic aerodynamic data is justified throughout the
analysis.
9.2.1 Procedure
If we assume a ficticious structural damping, g, and a purely harmonic response
p = ik, the equations of motion become:
1
−ω 2 [M ] + (1 + ig)[K] {q̂} = ρU 2 [A(k)]{q̂}
(9.4)
2
We can then set up the following eigenvalue problem:
1 1 + ig
[K]−1 [M ] + ρb2 [A(k)/k 2 ] {q̂} = {q̂} (9.5)
2 ω2
or:
1 + ig
[B(k)]{q̂} = λ{q̂}, λ= (9.6)
ω2
where[B(k)] and λ = λ′ + iλ′′ are, in general, complex. Now to generate a point
on the k method’s “flutter diagram”, we follow the following procedure:
1. Choose an altitude, which results in a value for ρ
2. Choose k and calculate [B(k)], which will include the aerodynamic data
at frequency k
3. Compute the eigenvalues for each mode λ1 ...λN
p ′
4. Compute the frequency of each mode ωi = 1/ λi
130 CHAPTER 9. FLUTTER CALCULATION METHODS
This procedure is repeated at the same altitude for different k’s , which generates
diagrams similar to that shown in figure 9.2 (note that a constant k is a straight
line on the U − ω plane). An estimate for the flutter speed, UF , can be obtained
by interpolation using two values near gi = 0.
k2
ω kF
k1
k2
k1
U
+
g
U
UF
The k method is efficient when the aerodynamic forces do not vary signifi-
cantly with M∞ or Re, as it then allows the flutter point to be found after a
small number of order N eigenvalue problems, where N is the number of gen-
eralised coordinates. It is worth noting that at the flutter point, no additional
sources of error are present, so the flutter solution is exact.
When the aerodynamic forces do vary significantly with M∞ or Re, addi-
tional iterations are required. This is because the flight velocity is determined
from the computation after A(k) has been specified. Additionally, when flight
control systems are involved for which the frequency parameter is ω instead of
k, an additional interpolation step has to be made to match both parameters.
This can make the application of the k method unwieldy in general design cases.
Outside of these efficiency issues, the k method suffers from practical problems
in implementation and interpretation. These include difficulties associated with
connecting the solution pairs in the flutter diagram to form sensible frequency
and damping curves, especially when the curves merge. In some cases it is
9.3. P-K METHOD 131
+
g
The full equations of motion are then used to determine the correct value of p
for each mode at a given flight condition.
132 CHAPTER 9. FLUTTER CALCULATION METHODS
9.3.1 Procedure
With our assumed form of the aerodynamic forces, the equations of motion can
be written:
2
U U 1
2
2
[M ]p + [D]p + [K] {q̂} = ρU 2 [A(k, M∞ , ...)]{q̂} (9.8)
b b 2
or
3. Update p using: p2 F1 − p1 F2 ;
p3 =
F1 − F2
4. Setp1 = p2 and p2 = p3 , then repeat from step 1 until converged
5. Choose the next flight condition (ρ, U , M∞ , Re, ...)
6. Choose two new values of p based on extrapolation from the previous flight
conditions
7. Return to step 1 and repeat to find the new p at the next flight condition
Once the mode is tracked through its desired range, a different mode is selected
and the entire procedure is repeated. The formula for p3 in step 3 can be
derived using the diagram shown in figure 9.4. Note that only one determinant
evaluation is required per iteration as after the procedure is started, as F1 can
be taken from the previous iteration.
Although the process of finding a values of p can be costly for complex F ,
this must be compared with the high cost of finding eigenvalues for large systems
as required by the k method. Additionally, no further iteration is required by
the p-k method to incorporate Mach or Reynolds number effects. An additional
9.4. P METHOD 133
F1
p3 p
0 2 p
p
1
F
2
9.4 p method
If we want flutter diagrams which are accurate over the complete range of condi-
tions, we will be forced to construct an approximate expression for [A(p)] which
is accurate for decaying or growing motions. One of the major complicating fac-
tors in doing so is representing the time lags inherent in unsteady aerodynamic
responses.
Once [A(p)] is specified, the resulting system can be solved using the same
determinant iteration procedure used in the p-k method. Alternatively, A[(p)]
can be expressed in the time domain using an inverse Laplace transform, and the
solution carried out as an eigenvalue problem in state space. The dimension of
state-space problem is usually higher, as new state variables have to be introduce
to properly represent the aerodynamic lag modes.
Approximate models for [A(p)] are typically of the form:
L
" #
2
X A(l+2)ij (M∞ )p
A0ij (M∞ ) + A1ij (M∞ )p + A2ij p + (9.10)
βlij + p
l=1
• The first three terms are instantaneous in nature and are often referred to
as aerodynamic stiffness, damping and mass terms
It is instructive to compare the lag effect terms above to the approximation for
Theodorson’s function given in chapter 7 (Note that the latter approximation
accounts for the circulatory part of the response only).
The main advantage of the p method is that it can accurately represent non-
harmonic aerodynamic responses, including converging, diverging and aperiodic
motions. In its state-space form, it is also effective for control system design. Its
disadvantage is that it can require a complex procedure for accurately approx-
imating the aerodynamic forces, and its state-space approximations can lead
to very large eigenvalue problems. Also, the p method will find a slightly dif-
ferent value of flutter speed than determined using k and p-k methods, due to
additional interpolation errors in the aerodynamic forces.
k=1 A(k) = 4 + 2i
k=2 A(k) = 2 − 4i
p9.5.21 For the single DOF system described above, use the p-k method to
estimate p = δ + ik at U=0.4. Use an assumed value of δ1 = −0.1 with the
experimental A(k) for k = 1, and δ2 = 0.1 with the experimental A(k) for k = 2.
Perform a single step of determinant iteration.
Chapter 10
Dynamic Responses
Tail−
Root
Bending
Moment
Rigid T a i l
135
136 CHAPTER 10. DYNAMIC RESPONSES
Analysis of cases with less severe consequences can also be useful. Passenger
comfort, for example, is affected by the details of response to atmospheric turbu-
lence. On the other hand, non-critical dynamic responses can also be measured
and used by aeroelasticians in order to validate the numerical models they have
developed.
From the analysis point of view, dynamic response problems are more gen-
eral than dynamic instability problems, as we can no longer exploit the sim-
plifications inherent in assuming undamped harmonic solutions. In fact, the
complexity of modern aircraft structural and control systems, coupled with the
difficulties of accurately predicting unsteady aerodynamic flows, means that
dynamic response problems can pose a significant challenge.
1 2
[M ]{q̈} + [D]{q̇} + [K]{q} = ρU {A(t, q, v, M∞ , Re)} + {E(t)} (10.1)
2
[M ] : Mass matrix
[D] : Structural damping matrix
[K] : Stiffness matrix
{A(t..)} : Aerodynamic forces
{q} : Vector of generalised deformation coordinates
{v} : Vector of external aerodynamic excitations
{E(t)} : External mechanical forces (e.g landing gear)
Here {A(t)} defines the combined unsteady aerodynamic forces arising from
both excitation (gusts) and deformation. For a truly general analysis, this must
10.2. DYNAMIC RESPONSE EQUATIONS 137
? ?
Deterministic
- Control deflections Stochastic
- Prescribed gusts - Atmospheric turbulence
- Initial landing impact - Ground roll, water waves
wg wg
? ?
Time Frequency ?
Domain Domain Frequency
- Numerical - Fourier Domain
integration Transform - PSD Technique
1 2
[M ]{q̈} + [D]{q̇} + [K]{q} = ρU ([A(t)]{q} + [Ae (t)]{v}) + {E(t)} (10.2)
2
z
U x
y
U
wg αg
x
wg wg
In practice fast Fourier transforms are often used to perform these steps.
The equations of motion for our standard 2 DOF (h,θ) representative section
with a low-frequency aerodynamic model (see chapter 2) can be written:
where
and:
1 0 0 0
0 m 0 Sθ
[M ] =
0 0
(10.12)
1 0
0 Sθ 0 Iθ
0 −1 0 0
Kh CLα qS/U CLα qS 0
[C] =
0
(10.13)
0 0 −1
0 −CLα qSec/U Kθ − CLα qSec 0
We will consider some dynamic responses for a section with the following prop-
erties:
ρ = 0.53kg/m3
e = 0.2, c = 6 m
m = 400 kg/m
Sθ = 180 kg m/m
Iθ = 200 kg m2 /m
Kh = 1 × 105 N/m /m,
Kθ = 3 × 105 N m/m
For reference, the associated flutter diagram is shown in figure 10.5.
60
50
40
30
20
10
0
0 50 100 150 200
Figures 10.6 and 10.7 show the response of the system to a sharp-edged gust with
a magnitude of 10m/s encountered at t = 0. Initially there is increased damping
10.5. STOCHASTIC RESPONSE ANALYSIS - PSD TECHNIQUE 141
0.08 0.2
0.06 theta theta
h 0.15 h
0.04
0.1
0.02
0 0.05
-0.02 0
-0.04 -0.05
-0.06 -0.1
-0.08
-0.15
-0.1
-0.12 -0.2
-0.14 -0.25
0 0.5 1 1.5 2 0 0.5 1 1.5 2
0.3 80
theta theta
h 60 h
0.2 40
0.1 20
0
0 -20
-40
-0.1 -60
-0.2 -80
-100
-0.3 -120
0 0.5 1 1.5 2 0 0.5 1 1.5 2
the speed is increased further however, the damping decreases until the system
becomes unstable. Depending on the maximum allowed strain of the structure,
the 100 and 110 m/s cases may already be critical, as these produce relatively
large excursions.
The mean square value can also be expressed in terms of the Power Spectral
Density (PSD) function:
Z ∞
x¯2 = Φx (ω)dω (10.16)
0
where:
2πX(iω)X(−iω)
Φx (ω) = (10.17)
T
and
T
1
Z
X(iω) = x(t)e−iωt dt (10.18)
2π −T
The function Φx (ω) may be thought of the distribution of power across the
frequency domain. Expression (10.16) can be derived using the autocorrelation
function for x assuming it to be defined by a processes which is statistically
stationary (its statistical properties are independent of time) homogeneous (its
statistical properties are independent of space) and isotropic (its statistical prop-
erties are independent of direction) [see “Aeroelasticity” by Bisplinghoff et al,
appendix C ].
If we assume that the response to the excitation is also a random function, then
it can be shown that for a linear system:
where:
Dryden Spectrum
Von Karman Spectrum
Gust PSD function
Frequency
1.2
|H|^2
Phi g
1 Phi r
0.8
PSD function
0.6
0.4
0.2
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency
p10.6.22 An wing with mass m enters a stochastic gust field with a power
spectral density, Φg (ω), approximated by:
A
Φg (ω) =
B + ω2
The wing moves with a single degree of freedom, h, and is supported by a spring
Kh . Its lift may be approximated by:
!
ḣ wg
L(t) = CLα qS +
U U
L(t)
wg
h K
h
111111111111
000000000000
000000000000
111111111111
Compute the mechanical admittance, H(ω), and the PSD of the response,
Φr (ω). If A and B are positive, will increasing Kh increase or decrease the
maximum response amplitude? Use sketches of H(ω) and Φg (ω) to illustrate
your answer.
p10.6.23 Consider the single DOF plunging airfoil of the previous problem.
Using the frozen gust assumption and results from unsteady incompressible flat-
plate theory, express the response PSD as a function of reduced frequency for
an arbitrary gust PSD function, Φ(k). Assume that the lift due to a harmonic
gust with amplitude ŵ can be expressed:
Lg = 2πρU b [S(k)] ŵ
146 CHAPTER 10. DYNAMIC RESPONSES
Chapter 11
Introduction to
Computational
Aeroelasticity
147
148CHAPTER 11. INTRODUCTION TO COMPUTATIONAL AEROELASTICITY
11.1 Treatment of CA
Our look at computational aeroelasticity over the next three chapters (CA) is
summarised in the following diagram:
11.1. TREATMENT OF CA 149
Computational structural
dynamics
Figure 12.1:
151
152 CHAPTER 12. COMPUTATIONAL STRUCTURAL DYNAMICS
y
node i
Figure 12.2:
φi
element e
ui ui+1 Nodal Displacements
x Shape Functions
0 L
Βi+1 = 1/L
x Derivatives of
0 Shape Functions
Βi = − 1/L
Figure 12.3:
L
1 ρA
Z Z
2
T = ρu̇ dV = (u̇i φi + u̇i+1 φi+1 )2 dx (12.2)
2 v 2 0
M = ρAL
6
Mei
4 1
1 4 1
1 4
Mei+1
where the ith row of the system expresses Lagrange’s equation for the generalised
coordinate ui :
d ∂T ∂T ∂U
− + = Qi (12.9)
dt ∂ u̇i ∂ui ∂ui
mü + ku = 0 (12.10)
The
p solution of this system is a constant-amplitude oscillation with frequency
k/m.
Now assume we will advance the system in time using the Explicit Euler
integration technique, which is defined by a simple first-order Taylor series ex-
pansion in time. In the following, we will assume that the solution is advanced
with a constant time-step ∆t. We denote the current time level by n so that
tn = n∆t. Considering a step from n to n + 1 with the Explicit Euler technique,
the discrete version of our system can be written:
or:
n+1 n
u u 1 ∆t
= [C] where [C] = (12.14)
u̇ u̇ −∆t(k/m) 1
156 CHAPTER 12. COMPUTATIONAL STRUCTURAL DYNAMICS
Using [X], the matrix of right eigenvectors of [C], we can also express this system
in “modal coordinates” w1 , w2 as:
n+1 n+1 n
w1 u u
= [X]−1 = [X]−1 [C][X][X]−1 (12.15)
w2 u̇ u̇
n
σ1 0 w1
= (12.16)
0 σ2 w2
This method is also first-order accurate, but in this case, | σ1,2 |< 1 so that
the amplitude is decreased each time step. One has to be aware of this type of
error during a computation, as this damping introduced by the time march can
stabilise an otherwise unstable mode.
In general, all numerical integration methods produce either amplitude er-
rors (signal growth or dissipation) or phase errors (time leads or lags) or both,
depending on their design. We may express the amplitude error per time step
(local amplitude error) as:
eamp = 1− | σ | (12.19)
for such a case it is clear that the eigenvalues which characterise the be-
haviour of the numerically time-integrated solution will be functions of [M ] and
[K]. Therefore, the type of spatial discretisation used also effects the result-
ing amplitude errors (stability) and phase errors of the solution. Multiple-DOF
systems produce multiple modes, each with an associated oscillation period.
If the oscillation periods are close to each other, then conditionally-stable
explicit methods may be preferred, since they require far less work per time
step then implicit methods. If the oscillation periods are far apart in magnitude
(i.e. the matrix resulting from the discretisation is stiff), then unconditionally
stable implicit methods are generally preferred. This is because typically only
the lower frequency modes are of interest, and implicit methods permit the use
of large time steps. In contrast, a conditionally-stable explicit method would be
forced to march at the time-step required to prevent instability of the highest-
frequency mode.
and two Taylor series expansions in terms of the anticipated value of acceleration
at the next time step:
∆t2
un+1 = un + ∆tu̇n + [(1 − 2β)ün + 2β ün+1 ] (12.24)
2
u̇n+1 = u̇n + ∆t[(1 − γ)ün + γ ün+1 ] (12.25)
This generalised eigenvalue problem can be solved for the eigenvalues, ω 2 , and
the eigenvectors (or natural modes) Φi . The latter are normalised so that
[Φ]t [M ][Φ] = I, where [Φ] is the matrix of eigenvectors (orthonormal modes).
We can then use orthogonal properties of the modes to decouple the dynamic
system, by noting the matrix of orthonormal modes, [Φ] has the following prop-
erty:
2
ω1 . . .
. ω22 . .
[Φ]t [K][Φ] = [Λ] = (12.27)
. .
. . . .
2
. . . ωN
(12.28)
while by definition:
Overall, using modal superposition can substantially reduce the cost of nu-
merical integration. Furthermore, we can now choose to retain only the low-
frequency modes. This approach to reduced-order modelling procedure allows us
to perform compact computations which retain the most important behaviours
present in a complex geometry.
160 CHAPTER 12. COMPUTATIONAL STRUCTURAL DYNAMICS
p12.7.24 Derive the element mass and stiffness matrices for a bar element
loaded in torsion using linear shape functions.
p12.7.25 Compute the torsional normal modes of the cantilever bar shown in
the figure below. Use two finite elements with linear shape functions.
!!!!!!!!!!!
!!!!!!!!!!!
::::
!!!!!!!!!!!
::::
!!!!!!!!!!! Ιθ(y) = Constant
::::
!!!!!!!!!!!
::::
!!!!!!!!!!!
::::
!!!!!!!!!!!
::::
!!!!!!!!!!!
GJ(y) = Constant
::::
!!!!!!!!!!!
::::
!!!!!!!!!!!
::::
!!!!!!!!!!!
::::
y=0
!!!!!!!!!!!
::::
!!!!!!!!!!!
::::
θ(0) = 0
!!!!!!!!!!! θ(y)
::::
!!!!!!!!!!!
!!!
!!!
y=1!!!
θ’(1) = 0
p12.7.26 Compute the forced response θ(y, t), of the bar in question 2 to a
torque T = Asin(ωt) applied at y = l. Use the results of question 2 and the
modal superposition approach.
mü + ku = 0
un+1 = un + ∆t
2 (u̇
n+1
+ u̇n )
u̇n+1 = u̇n − ∆tk
2m (u
n+1
+ un )
Compute the phase and amplitude errors produced by this method per time
step when k/m = 1 and ∆t = 0.1.
Chapter 13
Figure 13.1: Flutter speeds predicted with different aerodynamic models (from
Schuster et al, J. of Aircraft (40) 5 , 2003)
161
162CHAPTER 13. COMPUTATIONAL METHODS FOR UNSTEADY FLOWS
better predictions of classical flutter modes, non-linear methods can also be used
to consider responses which are purely dependent on phenomena which cannot
be computed with linear methods. Examples include aileron buzz and stall
flutter, although the truly accurate computation of these phenomena remains a
considerable challenge.
In our description of non-linear methods, we will first consider the small-
disturbance potential equation, and then move on to the Euler and Navier-
Stokes equations, for which we will also discuss the analysis of numerical errors.
We will not discuss methods for the full-potential equations, as these are similar
in form to methods for the Euler equations. Figure 13.2 presents an overview
of the models, concepts, and methods considered in this chapter.
Small-Distubance Linearised
Finite Difference
Potential moving-wall B.C.s
2
1 ∂2φ ∂2φ
2 2 ∂ φ
∇ φ− M∞ − 2 + 2U∞ =0 (13.1)
∂x2 a∞ ∂t2 ∂x∂t
(13.2)
∂φ ∂φ
p − p∞ = −ρ + U∞ (13.3)
∂t ∂x
∂ 2 ψ̂ M∞ ∂ ψ̂ ω2
2
∇2 ψ̂ − M∞ 2
− 2iω + 2 ψ̂ = 0 (13.5)
∂x a∞ ∂x a∞
p − p∞ = −ρψ̂ (13.6)
The first equation is similar to the velocity-potential form for harmonic motions.
The second equation is now much more convenient for the application of wake
boundary conditions. In this case the ∆p = 0 wake condition can be simply
expressed as ψ = 0 on the wake surface.
z y,η
za
x,ξ
S
Figure 13.3: Basic wing geometry
where:
Z
ŵ(x, y) = ∆ψ̂(ξ, η)K(x − ξ, y − η, k, M∞ )dξdη (13.8)
S
where fi (x) and gj (y) are assumed continuous functions which satisfy the pres-
f(x)
g(y)
sure boundary conditions on the edges of the wing (singular at leading edge,
13.1. LINEARISED POTENTIAL EQUATION 165
zero at the tips and trailing edges). Example functions are illustrated in figure
13.4. The âij coefficients can be found by enforcing flow tangency condition at
n × m control points using the integral equation for w and the assumed form
for ∆ψ̂(x, y). This leads to a system of (n × m) equations for the unknown
coefficients. Note that the wake need not be included in the integration since
∆ψ̂wake = 0. This is because all of the assumed harmonic wake-lag effects are
already encapsulated in the definition of K.
Since the main equation is non-linear, superposition cannot be used. This means
that the entire flow domain must be discretised (figure 13.7). As we are con-
sidering small disturbances, however, it is customary to linearise the boundary
conditions. The latter is an extremely extremely convenient simplification, as
then the application of flow tangency on the airfoil surface does not require dis-
tortion of the grid in order to account for structural deformations. A popular
Far−Field Boundary
j,k+1
j,k j+1,k
j−1,k
j,k−1
Airfoil Wake
course, as they do not include the effects of viscosity. In theory this means that
13.3. EULER AND NAVIER-STOKES EQUATIONS 169
they also do not enforce the Kutta condition, but in practice this is enforced
automatically due to the necessary presence of finite amounts of numerical dis-
sipation.
When other viscous effects must be considered, the Reynolds-averaged Navier-
Stokes equations are normally used. This raises an important issue concerning
the fidelity of turbulence models used to predict unsteady flows. A lack of un-
steady data makes such models difficult to calibrate, and often models calibrated
using steady data are used. Another problem occurs if the turbulent fluctuations
of the largest eddies approaches the time scales of the structural vibration. In
this case the assumptions underlying the Reynolds-averaged approach are com-
promised, and one formally should switch to large-eddy simulation. Due to
resolution requirements, however accurate LES is currently not possible at the
Reynolds numbers of interest to aircraft applications.
In contrast to the small-disturbance equations, in most cases it is inap-
propriate to linearise the boundary conditions for the Euler and Navier-Stokes
equations. This adds considerable complication to the solution procedure, as the
discretisation must be general enough to treat deforming domains (figure 13.9).
Aside from the complication of implementation, it has been demonstrated that
Mesh Generation
Fully−Discrete Approach
Semi−Discrete Approach Spatial Discretization
Time Integration
Fully−Discrete System
(Algebraic Equations)
Due to computational cost the domain of the problem must remain limited,
which means that important influences of physical phenomena (such as con-
vecting wake vorticity) can be lost as they leave the domain. This contrasts
with steady problems, which typically have uniform far-field conditions. The
latter can be be represented by relatively simple procedures, provided solution-
error transients can be properly expelled.
discrete solution values as a function of time. These are integrated using either
explicit or implicit time-integration methods.
For unsteady problems the solution accuracy is a function of both the spatial
discretisation and time integration method. This is different from when steady-
flow problems are considered, for which the accuracy of the time integration
method normally used to solve the non-linear problem is unimportant.
In general, explicit time-integration methods (e.g. MacCormack, Runge-
Kutta) are restricted for stability reasons to small time steps, on the order of
the time necessary for an acoustic wave to traverse the smallest cell. Implicit
time-integration methods (e.g. Trapezoidal) are stable for much larger time
steps but require more work per time step than explicit methods. Implicit time-
integration methods are often preferred, as the time scales of fluid-structure
interactions are usually much greater than the acoustic scale of the smallest
mesh cell, particularly for viscous calculations.
n
V
V : Control volume
S: Surface of control volume
x
n: Local surface normal vector
x: Local surface velocity vector
dS
d
Z Z
W dV + (F~ i − F~ v − ~ẋW ) · ~n dS = 0 (13.13)
dt
V S
d
Z Z
∗
W dV + (F~ i (W ∗ ) − ~ẋW ∗ ) · ~n dS = 0 (13.15)
dt
V S
13.6. EXAMPLE: A SEMI-DISCRETE FINITE-DIFFERENCE METHOD173
In this case, the viscousRfluxes F~ v (W ∗ ) = 0, and the inviscid fluxes are constant.
Since for a closed cell, ~n dS = 0:
S
d
Z Z
dV = ẋ · ~n dS GCL (Integral form) (13.16)
dt
V S
ξ=const
y η
η=const
x ξ
Wt + Wξ ξt + Wη ηt + fξ ξx + fη ηx + gξ ξy + gη ηy = 0 (13.18)
(Note that dissipative terms which vanish in the limit of zero mesh size are often
added to (13.18) to improve the stability of the numerical solution procedure).
To evaluate the metrics of the transformation, note:
dt 1 ξt ηt dτ
dx = 0 ξx ηx dξ (13.19)
dy 0 ξy ηy dη
and:
dτ 1 xτ yτ dt
dξ = 0 xξ yξ dx (13.20)
dη 0 xη yη dy
F
CCCCCCCCCCCCCCC
CCCCCCCCCCCCCCC
CCCCCCCCCCCCCCC
CCCCCCCCCCCCCCC
CCCCCCCCCCCCCCC
CCCCCCCCCCCCCCC
S
CCCCCCCCCCCCCCC
Wi
CCCCCCCCCCCCCCC
CCCCCCCCCCCCCCC
CCCCCCCCCCCCCCC
CCCCCCCCCCCCCCC
CCCCCCCCCCCCCCC
CCCCCCCCCCCCCCC
The conservation laws can applied directly to each cell (figure 13.14), resulting
in a system of ODEs for the time evolution of the average values:
M
d
(F~m − Wm~ẋm ) · S
~m
X
(Vi Wi ) = − (13.25)
dt m=1
n+1
n+1 Xb
Xa
t
Xa(t) Xb(t)
x n
n Xb
Xa
Normally it is chosen to compute Vin+1 and Vin exactly from the given mesh
deformation data, and then formulate the right-hand side in a way which sat-
isfies the DGCL. How this is done depends on both the spatial and temporal
discretisation. Consider a single (flat) side of a 2D cell as shown in figure 13.15,
combined with a first-order accurate time-integration scheme, and define:
Z tn+1 Z b
Iab = ẋ · ~n ds dt (13.27)
tn a
n+1 n+1
Xa Xb
Xa(t) Xb(t)
n
n Xb
Xa
n+1
Xb
n+1
y
n+1 Xc
x Xa
t Xa(t) Xc(t)
n n
Xa Xc
x
V x
Vo
(ξi , t) for which ∂V = J∂Vo (figure 13.18). The original ALE form then becomes:
d
Z Z
W J dVo + (F~ − ~ẋW ) · ~n dS = 0 (13.35)
dt
Vo S
d
Since Vo is fixed, we can bring dt inside the integral. By employing the diver-
gence theorem, the following weak form can be derived:
d(W J)
Z
w ~ ~
+ J∇ · (F − ẋW ) dVo = 0 (13.36)
dt
Vo
t n
@@@@@@@@@
@@@@@@@@@
@@@@@@@@@
@@@@@@@@@
V
@@@@@@@@@
@@@@@@@@@
@@@@@@@@@
S
x
Figure 13.19: A fixed control volume in space-time coordinates
Q n+2
t
Q n+1
Qn
Q n−1
x
Ωn
where f, g, and h are the x, y and z components of the space flux vector, F~ , the
conservation laws can be written:
Z
˜ · F̃ dṼ = 0;
∇ (13.39)
Ṽ
or
Z
F̃ · ñ dS̃ = 0 where ñ = (nt , nx , ny ) (13.40)
S̃
To advance the solution, we note that due to causality, only one part of the
space-time domain needs to be considered at a time. The space-time domain
is therefore split into slabs Qn , n = 1..N with constant-time boundaries Ωn at
t = tn , as shown in figure 13.20. The solution can then be advanced in time
by solving the slabs sequentially. If the mesh topology does not change in time,
the number of unknowns in each slab will be comparable to those in standard
time-marching techniques.
If a doubling of unknowns can be afforded, one can also choose to make the
boundary between slabs discontinuous in time, as shown between the first and
180CHAPTER 13. COMPUTATIONAL METHODS FOR UNSTEADY FLOWS
Q
t y
t +∆t
t
R
Ωo
x
second slabs of figure 13.20. The information from the previous slab is then
imposed as a weak boundary condition on Ωo , the boundary between Qn+1 and
Qn :
Z Z
˜ · F̃ dQ +
wh ∇ wh (uh+ − uh− ) = 0 (13.41)
Q Ωo
This gives enormous flexibility for the generation of moving meshes, as the mesh
in one time step can be completely unrelated to the one in the next time step
(the total number of elements, for example, can be arbitrarily changed from one
time step to another). In order to avoid evaluating the gradient of the flux,
(13.41) is usually expanded to:
Z Z
˜
−F̃ · ∇w dQ + wh F̃ · ñ dΓ
h
(13.42)
Q Γ
Z
+ wh (uh+ − uh− )dΩ = 0 (13.43)
Ωo
meshes which are coarser that what would be desirable. Consequently, the esti-
mation of discretisation errors due to finite space and time spacings is crucial,
especially when considering phenomena as sensitive as flutter.
log(ε)
Asymptotic
Region
Non−Asymptotic
Region
n
1
log(∆x)
negligible and the slope on a log-log plot becomes n. The error is then said to
be in the asymptotic region.
Results from standard CFD calculations must be in the asymptotic region
to be reliable. This can depend on the definition of ǫ, as results for drag may
require a finer discretisation than those for lift. In order to have confidence in
your results, variations with both h = ∆x and h = ∆t must be checked.
In practice, one does not have ue . Therefore plots similar to figure 13.22 are
produced by comparing coarse-mesh solutions to those of the finest mesh in space
and time. A straight-forward way to estimate error can then be constructed by
considering solutions computed on a sequence of grids with different h.
Consider uh computed with element size h, and u2h computed with element
size 2h, for a method which is second-order accurate (n = 2). g can be eliminated
182CHAPTER 13. COMPUTATIONAL METHODS FOR UNSTEADY FLOWS
Exact
solution
eamp
t
Numerical
x1 x2 solution epse
Vortex travelling relative to mesh
express:
• The convection of entropy with speed u
• The convection of vorticity with speed u
• The convection of acoustic perturbations with speeds u + a and u − a
where u is the local flow speed, and a is the local speed of sound. By analogy,
solutions for the linear convection equation:
∂u ∂u
+c =0 (13.47)
∂t ∂x
are often used to analyse and compare the performance of different numerical
methods for the Euler and Navier-Stokes equations.
13.11. PHASE AND AMPLITUDE ERROR ANALYSIS 183
m=6
m=5
∆x
m=4
Periodic 1D Domain
c∆t
epse = βj − arg(σj ) (13.54)
∆x
184CHAPTER 13. COMPUTATIONAL METHODS FOR UNSTEADY FLOWS
1
0.8 j=2
j=4
0.6
0.4
Amplitude
0.2
0
-0.2
-0.4
-0.6
-0.8
-1
1 2 3 4 5 6 7 8 9 10
Node index (M=10)
Where in this case, βj = 2π(j − 1)/M is the frequency of the eigenvector (wave)
associated with the eigenvalue σj . For the linear convection equation, the phase
and amplitude errors produced by a numerical method are a function of both
the wave frequency and the Courant number = c∆t/∆x, where ∆t is the time
step and ∆x is the spacing between nodes. In figures 13.26 to 13.28, phase
and amplitude errors for three different numerical methods are given, along
with numerical solutions for the convection of a block wave. In the first
case, the combination of a second-order accurate finite-difference method with
first-order accurate implicit time integration produces a method which is high
in phase error at high frequencies (beta) but also has very little damping at
high frequencies. The result is a sawtooth effect due to the incorrect speeds of
the high-frequency portions of the solution. In the second case, a second-order
accurate explicit time integration method is used, with coefficients designed
to produced better high-frequency damping, resulting in an improved solution.
Finally, results from a third-order accurate space-time method are shown, to
which high-frequency damping has been added. The remaining oscillations near
the discontinuous part of the solution can be further controlled using operators
tuned for shock capturing.
The above observations will also apply to more general flow features con-
vected by similar schemes for the Euler and Navier-Stokes equations. If you
think of a flow feature (such as a vortex) as being made up from waves of dif-
ferent frequencies, you can anticipate that its high-frequency components will
normally exhibit large phase errors. For a well-designed scheme, these compo-
nents will also be strongly damped. A side effect will be the smoothing out of
the flow feature as it is convected. Marching with higher Courant numbers will
tend to produce greater errors. It can also be seen that higher-order schemes
can be really worthwhile, provided that their cost is not excessive.
13.11. PHASE AND AMPLITUDE ERROR ANALYSIS 185
1
Courant=1.5
Courant=1.0
Courant=0.5
0.8
0.6
e_amp
0.4 1.5
Initial
Exact
0.2 Numerical
1
0
0 0.5 1 1.5 2 2.5 3
0.5
u
beta
3 Courant=1.5
Courant=1.0 0
Courant=0.5
2.5
2 -0.5
0 0.2 0.4 0.6 0.8 1
e_pse
1.5 x
0.5
0
0 0.5 1 1.5 2 2.5 3
beta
Figure 13.26: Second-order accurate centred finite difference method with first-
order accurate implicit time integration: 100 nodes, 80 time steps.
Over the last few years methods for formally estimating the magnitude of local
error contributions to quantities of interest, such as lift and moment, have de-
veloped rapidly. These involve the solution of dual or adjoint problems which
express the sensitivity of the error in the desired quantity to local values of the
residual. There is great potential for such methods in the field of aeroelasticity,
but their exists two basic complicating issues. The first of these is that the dual
problem evolves in the reverse time direction, making its simultaneous compu-
tation with the main (primal) problem a bit awkward. The second more serious
problem arises from the moving interface between the flow and the structure,
which poses serious difficulties for the implementation of the dual problem. The
reader is referred to the recent TU Delft thesis of Kris van der Zee for more
information and successful strategies.
186CHAPTER 13. COMPUTATIONAL METHODS FOR UNSTEADY FLOWS
1
Courant=1.5
Courant=1.0
Courant=0.5
0.8
0.6
e_amp
0.4 1.5
Initial
Exact
Numerical
0.2
1
0
0 0.5 1 1.5 2 2.5 3
0.5
u
beta
3 Courant=1.5
Courant=1.0
Courant=0.5 0
2.5
2
-0.5
e_pse
0.5
0
0 0.5 1 1.5 2 2.5 3
beta
Figure 13.27: Second-order accurate centred finite difference method with ex-
plicit second-order accurate Runge-Kutta time integration, 100 nodes, 80 time
steps (additional dissipation added for high frequencies).
1
Courant=1.0
0.8
0.6
e_amp
0.4 1.5
Initial
Exact
Numerical
0.2
1
0
0 0.5 1 1.5 2 2.5 3
0.5
u
beta
3 Courant=1.0
0
2.5
2
-0.5
e_pse
0.5
0
0 0.5 1 1.5 2 2.5 3
beta
-0.6 -0.6
-0.4 -0.4
-0.2 -0.2
0 0
Cp
Cp
0.2 0.2
0.4 0.4
Exper_Upper Exper_Upper
0.6 Exper_Lower 0.6 Exper_Lower
TSD Euler
Full_Potential Euler + BL
0.8 0.8
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c
-25 -25
-20 -20
-15 -15
-10 -10
Re(dCP/Theta)
Im(dCP/Theta)
-5 -5
0 0
5 5
10 Exper_Upper 10 Exper_Upper
15 Exper_Lower 15 Exper_Lower
TSD TSD
20 Full_Potential 20 Full_Potential
Euler Euler
25 25
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c
p13.13.29 Given two solutions uh and u2h , estimate the error of uh if both
solutions were computed using a first-order accurate method.
Computing Fluid-Structure
interactions
191
192 CHAPTER 14. COMPUTING FLUID-STRUCTURE INTERACTIONS
Examples of the dominant natural mode shapes for a finite wing are shown in
figure 14.1.
Figure 14.1: Examples of the first four mode shapes (φi , i = 1..4) for a wing.
(Taken from Raveh, Levy, and Karpel, AIAA paper 00-1325)
+ ... (14.3)
1
Z
Qis (t) = ǫφi · G(t)dS (14.4)
ǫ S
where G(t) is the computed response of the surface fluid forces due to a step
change in the structure geometry of shape φi and magnitude ǫ. An arbitrary
response can then be computed using discrete convolutions for the generalised
forces:
N =t/∆t
X
Qi (t) = Qi (0) + Qis (t) q˙i (t) ∆t i = 1, 2, ... (14.5)
n=1
If it is not sufficient to retain only h0 and h1 , then more general expressions for
Qi (t) must be constructed which include terms with multiple mode displace-
ments.
An alternative to the convolution approach is make a reduced- order model
of the aerodynamic operator. This can be done by computing the eigensystem
of the discrete CFD matrix which expresses the change in the fluid state over
a time step. The eigenvectors with largest magnitude are then used to make a
simplified model of the fluid response (see, for example Dowell, AIAA Journal
Vol 34, No 6, 1996). Although powerful in concept, this approach must be used
with care, as due to the multi-scale nature of fluid phenomena, mode truncations
can significantly reduce the range of behaviours which can be predicted.
194 CHAPTER 14. COMPUTING FLUID-STRUCTURE INTERACTIONS
F
M
S
Figure 14.2: Coupled integration as a three-field problem
Where q are the unknowns associated with structural deflections, W are the un-
knowns associated with the fluid state, and x are the unknowns associated with
the fluid mesh position. A visualisation of how the fields are coupled appears
14.3. PARTITIONED TIME INTEGRATION TECHNIQUES 195
in figure 14.3. Note that the choice of M̃ , K̃, and D̃, are in principal arbitrary,
Interface moving
displacement domain
Interface forces
as they only affect the dynamics of the fluid-structure system indirectly by de-
termining the quality of the mesh on which the fluid solution is computed. One
can choose M̃ and D̃ to be zero, for example, to avoid introducing complex un-
steady motion of the mesh. K̃ is normally constructed so that fluid elements in
critical regions are stiffer, which prevents them from becoming overly distorted
in response to boundary deformations.
Assuming that the mesh responds instantaneously to the structural defor-
mation, we can make the following classification of time integration techniques
for the remaining fluid-structure system:
Partitioned
A partitioned method advances the solution in time using one (fully-converged)
fluid solution and one (fully-converged) structure solution per time step. Such
methods are asynchronous in nature (first fluid, then structure or first structure,
then fluid). Extrapolation of previous results is sometimes used to reduced the
errors incurred by asynchronicity.
Monolithic
A monolithic method either fully or partially converges the complete fluid-
structure interaction system at the end of each time step by using multiple
fluid and structure solutions. Depending on the level of convergence, errors in-
curred by asynchronicity are eliminated.
Mixed
Mixed methods use monolithic techniques on low-order representations of the
system (e.g. coarse meshes) and partitioned techniques on higher-order repre-
sentations (e.g fine meshes). These have low but significant errors incurred by
asynchronicity.
Pn
Xn+1 3
1
Structure 2
tn tn+1
Fluid,
Fluid Mesh 2
Xn
1 3
Pn+1
Structure 4
tn tn+1
2. Redistribute the fluid mesh and advance the fluid from tn to tn+1
In this case, the forces in the fluid match those applied to the structure at each
time tn , while the displacements of boundary do not match at tn . In practice the
latter mismatch is not not necessarily as disturbing as it sounds, as the fluid and
structure interface meshes often differ in interface refinement, and thus already
do not match in the steady case.
1. Find the change in energy of the fluid, ∆EF , and the change in energy of
the structure ∆ES , during one time step
198 CHAPTER 14. COMPUTING FLUID-STRUCTURE INTERACTIONS
F S
CCCCCCCCCCCCCCCCCCCCCCC
CCCCCCCCCCCCCCCCCCCCCCC
CCCCCCCCCCCCCCCCCCCCCCC
CCCCCCCCCCCCCCCCCCCCCCC
P
CCCCCCCCCCCCCCCCCCCCCCC
CCCCCCCCCCCCCCCCCCCCCCC
X
2. Estimate the the changes in energy over one period of the oscillation,
(∆EF )T and (∆ES )T , using a continuous approximation for the n = T /∆t
discrete contributions
3. Determine how accurately the energy balance (∆EF )T + (∆ES )T = 0 is
satisfied by the partitioned scheme under consideration
For simple time-marching techniques, we may express the transfer of energy to
the fluid during a single time step ∆t as:
Z tn+1
∆EF = − ẊF APF dt = −(XFn+1 − XFn )AP̄F (14.9)
tn
where :
the energy change over a period can be expressed as the piecewise integration
of ∆E ∆E
∆t (figure 14.7). As ∆t represents an average value, This can be related to
the integration of a continuous function via the midpoint rule (figure 14.8). We
can therefore use the approximate identity:
2π
∆E
Z ω
(∆E)T ≈ dt (14.17)
0 ∆t
14.3. PARTITIONED TIME INTEGRATION TECHNIQUES 199
55555
55555
55555
∆E 55555
55555
55555
∆t 55555
55555
55555
55555
5555555555 5555555555 55555
5555555555 5555555555 55555
5555555555 5555555555 55555
55555
5555555555
55555 55555
5555555555
55555 55555
55555
5555555555 5555555555 55555
5555555555 5555555555 55555
5555555555 5555555555 55555
Figure 14.7: Energy change over a period of oscillation
∆E
∆t
Example Problem
Compute the energy-conservation accuracy of a coupled numerical solution pro-
cedure for the 1D model piston problem shown in figure 14.6. Consider the
case of a volume-discontinuous partitioned solution technique, combined with
trapezoidal time integration of both the fluid and the structure.
Solution
We will assume that the piston is oscillating harmonically, so that the discrete
solution can be described by:
Step 1: For trapezoidal time integration, the change in the energy of the fluid
during one time step may be expressed:
(PFn+1 + PFn )
∆EF = −(XFn+1 − XFn )A (14.20)
2
Similarly, for trapezoidal integration of the structure:
(PSn+1 + PSn )
∆ES = (XSn+1 − XSn )A (14.21)
2
200 CHAPTER 14. COMPUTING FLUID-STRUCTURE INTERACTIONS
Thus, the change in energy of the fluid and structure during a single time step
is:
P AX
∆EF = − [cos(ωtn ) − cos(ωtn − h)] [cos(ωtn + h + φ) + cos(ωtn + φ)]
(14.26)
2
P AX
∆ES = [cos(ωtn + h) − cos(ωtn )] [cos(ωtn + h + φ) + cos(ωtn + φ)]
(14.27)
2
Step 2: To compute the change in energy over one period of oscillation, we use
the approximate identity:
2π
∆E
Z ω
(∆E)T ≈ dt (14.28)
0 ∆t
h3 2h2
(∆EF )T ≈ kπ −h + + dπ −1 + (14.30)
3 3
h2
dπ
(∆ES )T ≈ [2sin(h)] ≈ dπ 1 − (14.31)
2h 6
14.4. MONOLITHIC TIME INTEGRATION TECHNIQUES 201
Step 3: The energy balance over one period of oscillation is given by:
h3
2
h
(∆EF )T + (∆ES )T = kπ −h + + dπ (14.32)
3 2
Note that although we have used (rather expensive) O(∆t2 ) accurate trape-
zoidal integration for both the fluid and the structure, the use of a partitioned
solution technique has reduced the time accuracy to O(∆t). For lower frequen-
cies, the displacement is out of phase with the change in pressure (pressure is
low for large X). Thus φ ≈ π and the change in energy of the complete system is
positive (note h > h2 > h3 ). As a result, this partitioned technique is formally
unstable for this undamped problem, and as a consequence the amplitude of the
oscillation will grow slowly in time. In the context of aeroelastic instabilities,
this means that the flutter speed of a more complex configuration might be
underpredicted by such a method.
This value is then transferred to the fluid in the first part of the partitioned
sequence. Using the analysis presented in the previous section, α1 and α2 can
be chosen in a way which improves the energy-conservation accuracy to O(∆t)3 .
In practice, this can increase the allowable time step by a factor of 10-100.
A similar idea can also be applied to volume-continuous partitioned methods,
by transferring an extrapolated estimate for P n+1 to the structure in the first
step of the sequence. When using extrapolation for the forces, however, some
care must be taken near moving discontinuities in the flow.
Where F F , M S, etc denote sub-matrices coupling the fluid, fluid mesh, and
structural fields. The complete system is not normally solved simultaneously,
however, due to the requirement of having a single program describe both the
fluid and the structure, and due to the global stiffness introduced by the very
different natures of each sub-matrix. The development of optimal decomposition
and preconditioning procedures for the complete matrix remains an interesting
research topic, however.
Thus to avoid the issues mentioned above, monolithic methods are usually
implemented as an iteration loop involving sequential fluid and structure solu-
tion steps performed until the system reaches a specified level of convergence.
Since the fluid problem is normally solved by a subiteration procedure within
each time step, the communication with the structure can be included in each
fluid subiteration with little extra cost.
The prime advantage of monolithic methods is that energy conservation
errors at the fluid structure interface can be kept arbitrarily small, or eliminated.
Like partitioned methods, it is possible to use extrapolation at the interface
to speed the rate of the converge, but this does not influence the final energy
conservation property. In principle, energy conservation errors can be controlled
even if the accuracy of the solution representation is very different on the fluid
and structure sides of the interface (see for example, Van Brummelen, E.H. et
al.: Energy Conservation Under Incompatibility for Fluid Structure Interaction
Problems, Computer Methods in Applied Mechanics and Engineering, Volume
192, Number 25, 20 June 2003, pp. 2727-2748(22)).
-3
-4
log10(error)
-5
-6
-7
-8
-9
1 2
log10(fluid-structure solutions )
conservative monolithic
non-conservative monolithic
partitioned withprediction
-2
-3
-4
log10(error)
-5
-6
-7
-8 2
1
-9
-2 -1
log10(deltaT)
influenced by:
− fluid and structure integration methods
− communication timing
− extrapolation procedures
CCCCCCCCCCCCCCCCCCCCCCC
CCCCCCCCCCCCCCCCCCCCCCC
CCCCCCCCCCCCCCCCCCCCCCC
CCCCCCCCCCCCCCCCCCCCCCC
CCCCCCCCCCCCCCCCCCCCCCC
CCCCCCCCCCCCCCCCCCCCCCC
Angle formulae
h4
Series expansions cos2 (h) ≈ 1 − h2 +
3
h2 h4 h4
cos(h) ≈ 1− + sin2 (h) ≈ h2 −
2 24 3
h3 h5 h3
sin(h) ≈ h− + cos(h)sin(h) ≈ h−2
6 120 3
(During an exam, all of these formulae will be made available when necessary
for completing a question)
Chapter 15
Aeroelasticity in Aircraft
Design
15.1 Requirements
There are a large number of design requirements associated with aeroelastic
phenomena. The basic bounds for these requirements are set by certification
standards. For large aircraft, these can be found in section 25 of the air regu-
lation documents of the relevant country or region (e.g. EASA CS-25, FAR-25,
CAR-525 etc). Aeroelasticity is mentioned most explicitly in section 629. Some
of the CS-25 aeroelasticity-related standards include:
• The aeroplane must be designed to be free from control reversal and from
undue loss of longitudinal, lateral, and directional stability and control
within the previously described VD/MD envelope.
207
208 CHAPTER 15. AEROELASTICITY IN AIRCRAFT DESIGN
Pre-design stage
• Analytic modelling
• Wind tunnel testing to support modelling of new aeroelastic features
Prototype stage
• Ground vibration and flexibility tests to provide final validation of struc-
tural models
• Analytic calculations to prepare for flight testing
• Flight-test program
Figure 15.1: Full aeroelastic model in tunnel (left) and a typical underlying
structural model (right)
Soft Supports
Shaker
• Aerodynamic vanes
- Good for low frequencies
- force varies with airspeed
- can distort aerodynamics
• Atmospheric turbulence
- No change to aircraft
- Excites both symmetric and asymmetric modes simultaneously
- Usually amplitude is too small to provide accurate damping estimate
- Flight time lost looking for it
A comparison of the measured response levels achieved via a rotating vane
and atmospheric turbulence is shown in figure 15.6. Strain gauges have been
traditionally used for the measurement of in flight responses, but piezoelectric
accelerometers now much more common as they are compact and light, and have
good frequency response. The current trend is towards self-contained “stick-on”
devices which can transmit data wirelessly to recorders on the aircraft or directly
to the ground station.
from range of speeds was then used to estimate the zero-damping point before
the next condition was considered. This approach considerably improved safety,
although testing in areas with non-linear damping trends remained risky.
Modern on-line flutter prediction methods make use of robust analysis tech-
niques, where a model with uncertainties is constructed using complex numerical
aerodynamic and structural representations. As the flight progresses, the flight
data is used to conservatively identify the uncertainties in the model, thereby
improving the accuracy of the flutter prediction (see figure 15.7). The result is
then transmitted back to the aircraft, where it can be shown, for example, as a
continuously-adjusted instrument showing the distance from the flutter point.
The latter display device is referred to as a “flutterometer” (figure 15.8).
15.5. FLIGHT TESTING 215
Robust Analysis
Method
Flutter Prediction
Figure 15.7: Robust analysis for improved on-line flutter prediction (From Lind,
R.C, NASA TM-97-206220)
216 CHAPTER 15. AEROELASTICITY IN AIRCRAFT DESIGN