Experimental_Simulations-1.7
Experimental_Simulations-1.7
Experimental Simulations
Reader
Lectures by G. Eitelberg
edited by T. van Pelt & T. Sinnige & E.M. Raijmakers
3 Wind tunnels 24
3.1 Different types of wind tunnels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.1.1 Wind-tunnel configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.1.2 Low-speed wind tunnels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.1.3 High-speed subsonic and transonic wind tunnels . . . . . . . . . . . . . . . . . . . . . 31
3.1.4 Supersonic & hypersonic wind tunnels . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2 Wind-tunnel corrections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.2.1 Blockage effect explained from conservation principles . . . . . . . . . . . . . . . . . . 38
3.2.2 Blockage in open test sections due to pressure effects . . . . . . . . . . . . . . . . . . 40
3.2.3 Potential flow approximation applied to wind tunnel interference corrections . . . . . . 41
3.2.4 Thrust and Drag corrections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4 Reynolds-number effects 57
4.1 Reynolds number effects on airfoils and high aspect ratio wings . . . . . . . . . . . . . . . . . 60
4.1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.1.2 Transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.1.3 Flow separation as a Reynolds-number effect . . . . . . . . . . . . . . . . . . . . . . . 62
4.2 Reynolds-number effect testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.2.1 Requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.2.2 Simulating the Reynolds number by the technique of tripping . . . . . . . . . . . . . . 66
4.2.3 Effect of tripping at different locations along the chord of the airfoil . . . . . . . . . . . 67
4.2.4 Conclusion for airfoil and wing testing . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.3 Highly swept wings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.4 Slender bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.5 Pseudo Reynolds-number effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.6 Summary and recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
5 Engine integration 78
5.1 Thrust generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6 Noise 95
6.1 Human hearing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
6.2 Acoustics as a wave phenomenon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
6.3 Scaling of sound . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
6.4 Measurement devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
6.4.1 Presentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
6.5 Beam forming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
6.5.1 Delay & sum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
6.5.2 Conventional beam forming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
6.6 Array performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
6.7 Wind tunnel noise measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
6.7.1 Moving sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
8 Balances 118
1 Why experimental simulations and what
is it?
The relatively brief history of aeronautical testing started with both experiments for phenomenological re-
search (L. Prandtl (1904)) as well as for product development (O. and W. Wright (1903)). The distinction
between the former (scientific curiosity driven experiment) and the latter (product development) provides
the justification for the title of the course : Experimental Simulations. It is about experimentation for science
and about simulation of flight performance in experimental facilities.
Figure 1.1: L. Prandtl with a water tunnel; visualization of the boundary layer.
The straightforward, curiosity driven testing requires the mastery of experimental techniques; it requires
understanding of the signal processing, resolution and measurement capacity.
The testing in a laboratory for the purpose of product development - for the purpose of obtaining from a
laboratory experiment information about the behaviour of a product in the real usage - requires additional
skills of designing and interpreting an experiment. The experimentation performed in aeronautical wind
tunnels for the purpose of predicting aircraft properties and behaviour, falls into this latter category. The
experiments are designed and interpreted as to represent aircraft in a particular flight situation - the ex-
periment simulates flight. This particular attention to the transferabilty of the experiment to interpret flight
performance, the required scaling and the simulation of air vehicle components to represent reality provide
the background to the title of the lecture series about experimental simulations. In this series of lectures the
focus is on aeronautical applications, although an occasional reference to ground vehicles or wind energy
might provide additional illustrations.
The first primitive testing facilities used by L. Prandtl and the brothers Wright were soon followed by more
sophisticated developments in a number of countries aspiring to establish their national capacity to fly.
Especially after the 2nd World War, in the period of striving for national independence in technologies, most
powers established their own research infrastructure for aeronautical testing. The foremost powers were
the US and the SU. The European nations F, UK, D, and NL were also aspiring to maintain independence
in this strategic endeavour.
1.1.1 Aeronautical industry and development as drivers for the contents of the
lecture series
The challenge faced by the aeronautical industry in the aircraft development phase is achieving the required
performance targets including stability and controllability of aircraft while minimizing the fuel consumption.
This scaling rule can be derived from the appropriate vorticity production equations, see relevant literature
(eg. Lighthill [15]). Here the Strouhal number, Str, represents the unsteadiness of the flow/flight. This
scaling rule shows that at high Mach numbers, much more effort goes into maneuvers than at low low Mach
numbers, even if the Reynolds number stays identical, i.e. the viscous effect is amplified.
An additional component in experimental simulation is the need to consider not only the unsteady aerody-
namics, but also the structural properties of the vehicle or experimental model as well. This is covered by
the term testing for aero-elastics.
From a technological point of view, it is important that the future developments in aeronautical materials
leading to changed behavior of the aircraft structures under wind loads can be properly simulated also in
the cruise conditions. Here the capability again will have to offer a satisfactory compromise between the
aerodynamic scaling of the flow and the strength/elastic scaling of the structures.
Further, there are a range of issues related to the integration of propulsion which will require high quality
experimental studies for clarification. A standard issue for the aircraft designer is the issue of choosing
the proper placing of high by-pass ratio engines for clean undisturbed inlet flow. The comparison of the
performance of the isolated engine with the performance of the installed engine can currently only be
reliably performed in well-designed experiments. The experimental design here is mainly concerned with
providing simulators for the aircraft engines which are representative of the real engine performance. The
requirement for being representative is fulfilled by adequate thrust and power scaling, in addition to the
geometrical scaling.
Proper thrust book-keeping in the experiment requires a well-established calibration procedure, in order to
resolve the interference effects which are significant for flight, but small when compared with the thrust or
drag alone.
The issues of ingestion of the boundary layer, pylon- or wing- wake and of ground or fuselage vortices in
unconventional configurations can be unsteady and highly unpredictable. So will the combination of thrust
generation and ground proximity. In the integration of propulsion with the airframe, conventional definitions
of efficiency and apportioning of installation losses to either the airframe or engine tend to break down.
Therefore the experiments have to be thoroughly designed and understood.
Testing for helicopter or rotary wing aircraft performance can be considered as a subset of propulsion inte-
gration testing with strong unsteady elements in it. Both the performance as well as the acoustic emissions
of the rotating parts are of significant value in the chain of design evaluation.
The acoustics requirements for airframe as well as propulsor noise will continue to accompany the aeronau-
tical experiments for the foreseeable future. Since the noise generation mechanism is a complex process
of interaction of the flow with the submerged solid surface, the predictive capability of numerical methods
encounters its limitations in complex configurations. Aeroacoustic simulation experiments rely on the capa-
bility to design situations in controlled environment which are representative of the real-life situations. This
notwithstanding the need for phenomenological aero-acoustic experimentation for the understanding of the
principles of sound generation and propagation through shear- and boundary layers.
There are only very few experimental facilities available in Europe and world-wide, where properly scaled
aero-acoustic simulation can be achieved. The evidence obtained from comparing the controlled environ-
ment with flight testing points toward the possibility to shift part of the airframe noise certification process
from flight testing into wind tunnels. This shift is justifiable on account of the wind tunnels’ better controllabil-
ity and verifiability of conditions. In the aeroacoustic testing methodology, the measurement of the noise at
∂u U2 hmi
FI ∝ u ∝ (2.1)
∂x L s2
c In case of an axisymmetric problem, the radial inertia force (centrifugal force) can also be con-
sidered separately.
u2t hmi
FIradial ∝ ∝ Ω2 L ∝ f 2 L 2 (2.3)
r s
• Pressure
P P a m3
1 ∂p m
Fp ∝ ∝ = 2 (2.4)
ρ ∂x ρL kg m s
• Friction
1 ∂2u µ U hmi
Ff ∝ µ 2 ∝ (2.5)
ρ ∂x ρ L2 s2
• Gravity
hmi
FG ∝ g (2.6)
s2
• Capillarity
1 σ hmi
Fc ∝ (2.7)
ρ L2 s2
Since all scaling factors of the forces per unit mass for the five forces (and their components) shown above
are defined, it is possible to formulate their independent ratios.These ratios are usually then given a name.
This either in honour of somebody who first observed the importance of the ratio or somebody who produced
significant results in the relevant field.
Most flight testing is based on Froude-scaled vehicles. This Froude number scaling becomes obvious,
when the need to balance the gravity force with the lift force is considered. This means that the prod-
uct of the Froude number and the lift coefficient has to stay constant. In the following discussion the
An aircraft in level flight is kept in the air by the lift force balancing the weight of the aircraft W. In a
manner analogous to the pressure coefficient a lift coefficient can be formed to describe the similarity
of lift:
W mAC Lg
CL = 2 2 2
=2 · (2.14)
ρf u L ρf L3 u2
This shows that for similarity of level flight the relative density (expressing the hydrodynamic lift force)
of the aircraft and the Froude numbers are similarity parameters. For maneuver simulation additional
similarity parameters are formed for the maneuver accelerations compared with the gravity.
A short ship building example can also be discussed:
Intuition states that the drag of a ship is dominated by wave formation, as large drag generates large
waves in the ship’s wake and in the special case that the ship does not generate drag one would
assume that no bow and wake waves would occur. In the case that viscous forces are omitted, the
parameters to be considered are the speed of the ship, u, dimension of the ship, L, fluid density, ρ
and gravity, g.
As a dimensionless ratio is wanted to describe the phenomenon, the only possibility is that density
u2
drops out as no other terms contain mass. This ratio will result in the Froude number, F r = Lg and if
this number should be kept constant in scaled testing, the ship length should vary with the square of
the velocity.
However, if the viscous forces are to be included, the Reynolds number, Re = uL ν , should be kept
constant as well, which will mean that the length scale should vary inversely with an increase in
velocity. From this it is clear that both the Froude and Reynolds numbers cannot be kept constant
simultaneously. Therefore the product of the two is often kept constant.
u2 uL u3
F r · Re = = (2.15)
Lg ν νg
• Weber number:
The Weber number, W e, is a number often used in cases where there is an interface between two
different fluids, especially for multiphase flows with strongly curved surfaces. The Weber number is
the ratio between the inertial force and the capillarity force and the calculation is shown below.
inertia FI u2 L2 ρ u2
We = → =ρ L (2.16)
capillarity Fc L σ σ
The Weber number plays a role in small scale free surface processing, e.g. crystal growth through
melting/resolidification etc.
Apart from the the ratios above, products of the ratios can be formed that yield additional similarities, such
as the example below:
Fp FI Fp p ρuL pL
Stk = Eu · Re = · = → 2 = (2.17)
FI Ff Ff ρu µ µu
This Stokes number is used for the simulation of lubrication and the characterisation of the behaviour of
particles suspended in a fluid flow.
Assuming a constant ratio of specific heats, for aerodynamic similarities, it seems that the Reynolds number
effects have to be corrected by the square of the Mach-number. This combination is used in the hypersonic
flow treatment frequently as the so-called viscous interaction parameter. It will be encountered in later
discussions again.
For similarities of dynamic (characterized by a time scale T , especially periodic with a period of T ) phenom-
ena, the ratio of the two inertia forces, the unsteady and the steady components, can be formed:
U L fL
2
= = Str (2.19)
T U U
where f = 1/T is a characteristic frequency. This ratio is called the Strouhal number, Str, and is abbreviated
with 3 letters in order to distinguish it from the Stanton number St. For maintaining dynamic similarity, the
model length scale has to be maintained in a linear relationship with the convective distance during a period
of oscillation. The Strouhal number similarity is a requirement both for periodic movements of a test object
as well as for simulation of noise effects in aeronautics.
For steady rotation another approach can also be adopted: The unsteady scaling can be compared with
steady angular motion scaling. For a not co-rotating observer, a rotating object might represent a periodic
motion. For a co-rotating observer the motion is steady but coupled to a radial acceleration as under
equation 2.3. Forming a ratio of the radial inertial forces with the axial inertia forces will result in a ratio:
FI,rot f 2 LL
S= = = Str2 (2.20)
FI U2
This suggests that for rotating systems, if we maintain the Strouhal number similarity, we also maintain the
ratio of the radial and axial inertial forces. On the other hand, in terms of momentum balance, it can be
concluded that the significance of the radial momentum flux grows quadratically with the reduced frequency
and thus with the unsteady loading.
In addition, examples of the periodicity of sound waves generated by rotating wings or airplane propellers
can be discussed.
The last two terms play a role for example in space re-entry, otherwise radiation and reactions within a
medium can usually be neglected. This leaves us with four steady energy scales and one unsteady scale.
The scales derived below are for 1-dimensional cases.
λ ∂2T
λ∆T J
ėcd ∝ ∝ (2.22)
ρ ∂x2 ρL2 s
The energy flux through conduction is the part that occurs without motion or exchange of mass, it is
due to the temperature gradient imposed here by the boundary conditions of the problem.
• convection
∂T cp u∆T J
ėcv ∝ cp u ∝ (2.23)
∂x L s
The convective part of energy flux, although requiring a temperature gradient like the conduction
as well, represents the part of energy transport achieved with the help of mass transport with a
characteristic velocity U.
• friction
2
νu2 J
∂u
ėf ∝ ν ∝ (2.24)
∂y L2 s
The friction energy flux is the energy dissipated due to friction. Although this occurs in all viscous
flows, it is neglected in the approximation of isothermal flows.
• sources & sinks
J
ėss ∝ Q̇ (2.25)
s
The sources and sinks energy flux is governed by reactions within the medium.
• unsteady
cp ∆T J
ėunsteady: ∝ (2.26)
t s
From these energy fluxes independent ratios can be formed and are shown below.
• Péclet number:
The Péclet number is obtained when the ratio between the conductive and convective energy transport
terms is formed. It is of importance for the scaling of energy transport within a flow field. It illustrates
the similarity (or lack thereof) of the convective and diffusive processes of transport for momentum
and internal energy.
Where k is the thermal conductivity and defined as: k = ρcλp . The Péclet number can also be calcu-
lated as the product of the Reynolds Re and Prandtl, P r, numbers.
uL ν ν
Pe = = Re = Re · P r (2.28)
k ν k
In this ratio of energy fluxes, two terms can be recognized. The first component is the inverse of the
already identified Reynolds number. The second component is a ratio of the kinetic energy in the flow
to that of a characteristic enthalpy (difference) in the flow. The latter component is called the Eckert
number.
u2 kinetic energy
Ec = = (2.30)
cp ∆T enthalpy
u2 u2 T T
Ec = = = (γ − 1)M 2 (2.31)
cp ∆T cp T ∆T ∆T
When looking back to the ratio of the dissipation and convective flux it can also be represented in a
different way.
ėf Ec M2 T
= = (γ − 1) (2.32)
ėcv Re Re ∆T
2
Where M Re is referred to as the viscous interaction parameters and is often used in hypersonic flow.
We encountered the same parameter earlier when discussing the ratios of forces (momentum fluxes)
as well.
• Fourier number:
The Fourier number is obtained as the ratio between the unsteady convective energy flux and con-
ductive transport of energy.
The time scale t in the Fourier number can also be represented by the frequency, f = 1t , which leads
to the following relation for the Fourier number.
L2 f
Fo = (2.34)
k
It can be seen that for similarity, the frequency needs to be scaled with L2 , when unsteady conductive
processes are considered. This similarity requirement is missing in the ratios of forces, but when just
the steady and unsteady convective transport components are considered, then the Strouhal number
will be the same. Thus the conductive and convective unsteady similarities cannot simultaneously be
fulfilled in a scaled experiment.
An example for the use of the Damköhler number is the distance between the shock and the object in
hypersonic flows as shown in figure 2.1.
F = ma
(2.36)
[F ] = M LT −2
M :1=x
L : 1 = −3x + y (2.38)
T : −2 = −y
As can be seen these three equations are incompatible, thus the properties have independent dimensions.
Π = Φ(Π1 , . . . , Πm ) (2.44)
This, non-dimensional, expression confirms the intuitive observation, that physical laws cannot depend on
the choice of units. With the derivation above the Π-theorem has just been explained. This Π-theorem
states the following:
‘A physical relationship between some dimensional quantity and several dimensional governing parameters
can be expressed as a relationship between a dimensionless parameter and several dimensionless prod-
ucts of the governing parameters. The number of dimensionless products is equal to the total number of
governing parameters less the number of governing parameters with independent dimensions’.
In other words, the Π-theorem can result in orders of magnitude reduction of evaluation effort.
2.2.3 Examples
Pendulum
Determine the time of the oscillation period, τ , of the pendulum from figure 2.2. [τ ] = T It can be seen that
In figure 2.3 a representation of an airfoil submersed in transonic flow is shown. The governing parameters
are the speed of sound in the boundary layer, aδ , and the flow speed at the boundary layer, uδ . Furthermore
Figure 2.3: Boundary layer and shock interact at transonic flight conditions.
the boundary layer thickness, δ, the radius of curvature of the airfoil, R, the density of the medium, ρ, and
the dynamic viscosity of the medium, µ are also of importance. It is possible to play with the independent
variables. In this case both δ and R can be chosen as the independent parameter for the length scale and
both aδ and uδ can be chosen as the independent parameter for the speed. However once one is chosen
the other one cannot be described as an independent variable, since the dimension of the other parameter
can be expressed as a function of the dimension of the chosen independent parameter. At separation the
wall shear stress is 0, τw and can therefore be described as:
τw = f (uδ , aδ , δ, R, µ, ρ) = 0 (2.46)
τw δ
Π=
µaδ
uδ
Π1 = = Mδ
aδ
(2.47)
R
Π2 =
δ
ρuδ δ
Π3 = = Reδ
µ
v = f (d, ρs , ρf l , µ, g) (2.48)
Mδ R
→∞
attached δ
R
δ
<∞
R
δ
<< ∞
Reδ
Figure 2.4: Parametric (M, Re)-curve, with the non-dimensional radius of curvature R/δ as the parametric
variable.
V
Π = √ = f1 (d, ρs , ρf l , µ, g)
gd
ρs
Π1 =
ρf l
√ (2.49)
dρf l gd
Π2 =
µ
√
V ρs dρf l gd
→ √ = f2 ,
gd ρf l µ
√
dρf l gd
As Re → 0 then µ → 0, which will result in.
V ρs
√ = f3 (2.50)
gd ρf l
Momentum conservation:
∂u ∂u 1 ∂p
u +v =− + ν∇2 u (2.52a)
∂x ∂y ρ ∂x
∂v ∂v 1 ∂p
u +v =− + ν∇2 v (2.52b)
∂x ∂y ρ ∂y
∂v 0 v 0 L ∂v 0 p∞ L ∂p0
2 0
L2 ∂ 2 v 0
ν ∂ v L
u0 0 + 0
= − 0
+ 02
+ 2 02
· (2.56b)
∂x δ ∂y u∞ ρ δ ∂x u∞ L ∂x δ ∂y δ
L 0
!
v 0 L ∂ Lδ v 0 ∂ 2 Lδ v 0 L2 ∂ 2 Lδ v 0
0∂ δv p∞ L2 ∂p0 ν
u + = − + + 2 (2.56c)
∂x0 δ ∂y 0 2
u∞ ρ δ ∂x0 u∞ L ∂x02 δ ∂y 02
The equations 2.56a and 2.56c will yield identical solutions for the non-dimensional velocities and pressures,
as long as the three similarity parameters for different flow fields are same.
From equations 2.56a and 2.56c, non-dimensional forms of conservation of momentum can be further
analysed for special cases, where orders of magnitude of each term can be discussed. In the following the
Reynolds number dependency of further approximations will be discussed.
u∞ L
1. Re = ν << 1 (creeping flow)
∂u0 v 0 L ∂u0 ∂p0 1 2 0 2
∂ u + L ∂ u
2 0
u0 0 + 0
= −Eu 0
+ 02
(2.57)
∂x} δ ∂y | {z } ∂x Re |∂x
δ ∂y 02
2
| {z | {z } ? |{z} |{z} {z } |{z} | {z }
∼1 ∼1 ∼1 >>1 ∼1 ? ∼1
∂ Lδ v 0
v 0 L ∂ Lδ v 0 2 L 2 L 0 0
L2 ∂p0 1 2
∂ δ v + L ∂ δ v
u0
+ = − Eu + (2.58)
∂x0 } δ ∂y 0 |{z} δ 2 ∂x0 Re | ∂x 02 δ2 ∂y 02
| {z | {z } ? |{z} |{z} |{z} {z } |{z} | {z }
∼1 ∼1 ? ∼1 >>1 ∼1 ? ∼1
Since the left-hand side is of O(1), the terms of order greater 1 on the right have to be compensated
by the unknown terms:
∂p0 ∂ 2 u0 L2 ∂ 2 u0
Eu · Re = + (2.59a)
∂x0 ∂x02 δ 2 ∂y 02
L2 ∂ 2 Lδ v 0 L2 ∂ 2 Lδ v 0
∂p0
Eu · Re = + (2.59b)
∂y 0 δ 2 ∂x02 δ 2 ∂y 02
L 2 ∂p0 ∂p0
This will hold when δ = O(1) and Re · Eu = O(1). This is the case since ∂x0 ≈ ∂y 0 ≈ O(1).
2. Re >> 1 From equation 2.57 it can be seen that the left hand side is of O(1). This means that the sum
of all terms on the right hand side should be of the same order. In order to satisfy the no-slip condition
1 L2 L √1 . Furthermore
the wall friction should be finite. This results in Re δ 2 = O(1), which results in δ ∼ Re
p∞
the Eu = ρu ∞
≤ O(1).
Uplate = U∞ →
U∞
div ~u = 0 (2.63)
D~u
ρ = −∇p + µ∇2 ~u (2.64)
Dt
The initial conditions for this problem at t = 0 are: u = 0 in the whole field.
The boundary conditions for t > 0 are: at y = 0 → u = u∞ & at y = ∞ → u = 0. The velocity of the
flow field is dependent on the distance from the plate and the time, u = u(y, t), it does not depend on the
horizontal coordinate.
∂u ∂2u
=ν 2 (2.65)
∂t ∂y
| {z }
diffusion
∂p u
since ∂y = 0. If the velocity is non-dimensionalised with the velocity of the flat plate, u0 = u∞ the momentum
equation can be further simplified to:
∂u0 ∂ 2 u0
=ν 2 (2.66)
∂t ∂y
Including the boundary and initial conditions the non-dimensional velocity follows u0 (0, t) = 1 & u0 (y, 0) = 0.
2
The remaining parameters t, ν, y 2 can be non-dimensionalised as yνt or √yνt . As a conclusion u0 = uu∞ =
f √yνt and the solution to the momentum balance is:
0 y
u = 1 − erf √ (2.67)
2 νt
describing an continuously falling velocity with the distance from the wall and at the same time continuously
increasing velocity (to its maximum value at the wall) in the field with time.
√δ
νt
t
νt
y
√y
U∞ u
= u′ 1
U∞
Figure 2.6: The dimensional and dimensionless velocity profiles of the Stokes flat plate boundary layer.
In addition to the velocity profile, also a vorticity profile can be used describe the flow. The vorticity in the
above described boundary layer flow can be expressed by
∂u u∞ −y2
rot ~u = =√ e 4νt (2.68)
∂y πνt
From here a few observations can be made. One is that at all times ∂u ∂y → 0 as y → ∞. Another one is that
∂u ∞ ∂u
∂y → 0 when t → ∞ and y > 0, but finite. The integral of the vorticity 0 ∂y
dy = const. The fact that the
integral stays constant shows that vorticity is produced at the boundary at the instant of the acceleration of
the flat plate, and consequently diffused into the fluid. For this case no additional vorticity is generated at
the wall after the initial acceleration in this zero pressure gradient boundary layer.
DU0
1
J · n̂ = −n̂ × + (∇p)y=0 (2.69)
Dt ρ
By non-dimensionalising this equation for the vorticity flux, the relative importance of the two terms can be
analysed. The ratio of the vorticity production by the accelerated surface to that of the pressure gradient
can be expressed as a product of a Strouhal number with the square of the mach number:
f L ρu2∞
= γStrM 2 (2.70)
u∞ p
The Strouhal number could be formed with a frequency, as in the case of flapping wings, or with the inverse
of a characteristic time required for a manoeuvre. This is an ongoing a topic of research, such as e.g. in
the case of the shear layers rolling off of a rocket during launch in the transonic and supersonic phase.
There are many more tools to approach the simulation of real life in a laboratory frame - the ones described
could also be taken to further levels of refinement. The above chapter should, however, be sufficient as a
starting point for most of the frequently encountered challenges. The material provided should be sufficient
to make intelligent choices of where to start and what to look for.
In the aerodynamic testing, the laboratory frame is represented by wind tunnels. These are usually char-
acterized by their capability to simulate some velocity or force ratios, the other similarities are customarily
neglected - sometimes with good justification, sometimes without - the reader should by now be able to
judge that. The next chapter will deal with the wind tunnels as the laboratory framework of aeronautical
testing.
fl
Str = ≈ 0.2. (3.1)
u
Thus for a scaled object, the frequency scales with the scaling factor (the inverse of the length scale) of a
geometrically similar object. This holds for most noise sources in an aircraft, like landing gear, wing and
flap tips, even for the angular speed of a rotating engine.
Another noise source can be a cavity in the structure. The frequency of the (acoustic) pressure oscillations
scale with the following expression for a Helmholtz resonator:
r
a A
fH = (3.2)
2π V0 L
Where a is the speed of sound, A is the surface area of the cavity neck, V0 is the cavity volume and L is the
length of the neck. This frequency is invariant with the flow velocity, thus does not scale with the Strouhal
number, but it does still depend on the geometry in a similar proportionality as the external noise. This
invariance of frequency with the velocity lends itself to the identification of noise sources - if the Strouhal
number stays constant, it is a protrusion; if the frequency stays constant (i.e. the Strouhal number varies
with the velocity), it is a cavity.
In the context of scaling reality, the Helmholtz resonator cannot be left as a dimensional quantity - it too has
to be scalable. the following scaling of Helmholtz resonators enables one to simulate them in scaled wind
tunnel models:
r
3 D
f D = const. D V (3.3)
a L
There are yet other mechanisms for generating noise in the flow, like periodic shear layer impingement etc.
which follow yet other scaling laws. This means, that while it is possible to study the mechanism of noise
generation in an experimental facility, it is not always possible to generate a similar noise to that observed
in real life.
In this example, part of the reasoning behind not scaling automobiles in wind tunnels can be found. We
shall look at the Ma and Re-number considerations separately.
The principles of the behaviour of wind tunnels can be described by the isentropic relations given below.
These equations are valid for both the subsonic as well as the supersonic behaviour of the flow in the wind
tunnel circuit, as long as there are no significant source or sink terms involved. The sources sinks could
be heat exchangers or flow straighteners in some sections of wind tunnels, for example. Some times there
are also actual sources, like engine simulators, or sinks like windmills in the tunnels. These objects will be
treated separately, when simulation of thrust or rotor propulsion is discussed. Here we deal with the wind
tunnels themselves.
T 1
= γ−1 (3.4a)
T0
1 + 2 M2
ρ 1
= 1 (3.4b)
ρ0 1+ γ−1 2
γ−1
2 M
p 1
= γ (3.4c)
p0 1+ γ−1 2
γ−1
2 M
Nevertheless, in the case for subsonic testing, the Bernoulli equation together with the mass conservation
is often sufficient to estimate the relationship between the nozzle contraction and the flow velocity. How to
use these equations to find the behaviour and limits of wind tunnels will be described in more detail later in
this chapter.
Figure 3.1: Wind tunnel classification in terms of flow regime. The drive system is only indicative; many
tunnels do not have a continuous drive. [7]
Test section
Flow field about a model
simulates conditions of flight Model
Controlled
airstream Motor
Balance
Test instrumentation
Return duct
in the corner along the diagonal than it is in the perpendicularly joining sections of the tube, infinitely thin
turning vanes will maintain this variation in the stream tube cross-sectional area. This change in cross-
section causes the flow to either expand or compress at the corners, both of which cause energy losses
and affect the uniformity of the flow. This phenomenon can be seen in figure 3.4. One can see that the
turning vanes depicted on the right keep the cross-section constant and therefore keep the pressure in the
flow constant.
Figure 3.4: The importance of turning vane design. In this case the gap width is held constant. Illustration
from an internal project in DNW. [12]
Example: As mentioned before, operating an open-return wind tunnel requires more energy than a closed-
return type as the flow needs to be constantly accelerated. But how much power does an open-return wind
tunnel actually need to operate? In order to find out how much this is the energy balance of the flow is
needed, which is shown below.
1
P = Ė = e + u2 ṁ
2
(3.5)
1
= e + u2 ρuA
2
Assuming that the internal energy of the flow does not vary, which is an assumption of reasonable accuracy
for low speed tunnels, the power required per square meter of wind tunnel cross-section is found to be
Preq = 12 ρu3 .
In case of the DNW-LLF wind tunnel a speed of 100 m/s can be achieved in a 6x6 m2 test section. If this
tunnel would have been open circuit tunnel it would require 22 M W of power to be delivered to the flow,
even more to the fan when accounting for its limited efficiency.
Fortunately, this wind tunnel is of a closed-return type, thus part of the kinetic energy of the flow is recovered
and about half that power is actually needed to drive the fan.
Another characteristic that needs attention, when utilizing the closed test section configuration is the rise of
a boundary layer along the wind tunnel walls. As the flow is constrained at the walls, a boundary layer is
created that not only decreases the effective cross-section but it also interferes with wall mounted objects,
such as testing a 2D wing in a test section where the wing covers the entire width of the tunnel. Ways to
improve this is by introducing boundary layer blowing (or suction, although this means additional removal of
momentum, which the the BL does already) in which ‘new’ air is introduced at the walls of the wind tunnel
which will make the flow more uniform as shown in figure 3.6.
Figure 3.6: Boundary layer blowing to reduce the wall effect. [1]
dA du
Au = constant ⇒ =− (3.6)
A u
C 2 psettl − pm ∆p
u2m = 2 ≈2 (3.7)
C2−1 ρ ρ
ρ 1 M2
≈ 2 + ... ≈ 1 − + ... (3.8)
ρ0 1 + M2 2
Apart from compressibility affecting the measured aerodynamic parameters in a subsonic (low speed) wind
tunnel, there are other effects that need to be considered as well. This could be the span when testing 2-
dimensional airfoils. In the subsonic regime both the effect of the span width and the effect of compressibility
can be accounted for with the help of linearised potential theory and they show up together as a sort of
compound effect. For the limiting case of zero compressibility and infinite span,
span b
M∞ → 0, σ = chord = c → ∞,
the relation for the lift coefficient is known to be Cl = 2πα, where α is the angle of attack in radians. The
effect of the span in the wind tunnel for an elliptic lift distribution changes the above relation for the lift
coefficient to be:
2πα
Cl = (3.10)
1 + σ2
For Mach numbers smaller than 1 another linearised correction for the effect of the Mach number on the lift
coefficient needs to be taken into account as shown below.
2πα 1
Cl = p · (3.11)
1−M 2 1+ √ 2 2
| {z ∞} | σ 1−M∞
Mach effect
{z }
span effect
The first term describes the influence of the Mach number upon the lift coefficient at infinite aspect ratio, the
second term corrects for finite aspect ratio. This means that for profile testing the Mach number influence is
not negligible, even at very low compressibility! Increase in Mach number is equivalent to reduction in span
width. Thus it is preferable to obtain Reynolds number variation independently of Mach number variation.
Therefore for fixed wing testing, M ≈ 0.3 should be seen as a limit.
T? 2
= ≈ 0.833 (3.12a)
T0 γ+1
γ
p?
γ−1
2
= ≈ 0.528 (3.12b)
p0 γ+1
Using the area relations and substituting the isentropic relations provided in the beginning of the chapter
one can find the relationship between the nozzle area ratio and the Mach number in the nozzle, of which a
crude derivation is given below.
ρAu = const = ρ? A? u? = ρ? A? a?
A ρ? a? ρ? ρ0 a? (3.13)
?
= =
A ρ u ρ0 ρ u
All the three ratios in the product of the right hand side of the lower equation in 3.13 are known from the
isentropic relationships given before. The multiplication results in the following expression:
2 γ+1
γ − 1 2 γ−1
A 1 2
= 2 1+ M (3.14)
A? M γ+1 2
Expression eq. (3.14) means, that the Mach number in the expanding part of the nozzle, when the sonic
conditions where obtained at the throat, is only a function of the area ratio as can be seen from fig. 3.8.
Figure 3.8: Relationship between the Mach number and the area ratio in an expanding flow channel. [2]
To accelerate the flow to supersonic conditions a convergent-divergent nozzle is required. The convergent
part is required to reach the sonic conditions. To reach the sonic conditions, an overpressure of approx-
imately 2 is required. After the sonic conditions are reached, additional expansion to yet higher Mach
numbers can only be reached in a diverging flow channel called a diverging nozzle. In consequence, the
Mach number achieved in a test section in a convergent-divergent wind tunnel is solely related to the ex-
pansion ratio ( AAthroat
test
), so the effect of pressure won’t influence the Mach number, provided the pressure to
achieve sonic conditions in the throat is provided.
In transonic wind tunnels the walls of the test section are often slotted or perforated. Since the transonic
testing is usually required for high speed cruise condition testing at free stream Mach numbers just slightly
below 1, the model may represent such a restriction in the test section as to create sonic conditions between
the wind tunnel wall and the model. This is called blocking or chocking the flow in the transonic wind
Figure 3.9: Perforated/slotted wind tunnel. Note that the reentry of the flow into the tunnel circuit is not
shown, but must occur in a closed loop configuration. [26]
Another way to deal with shocks reflected from the wind tunnel walls, is to design the experiment in such
a way that the shocks generated by the model do not interfere with the tests. In the transonic testing for
civil transport aircraft, the shocks closing the local supersonic regions above the wings stay far from the
walls. In supersonic or hypersonic testing, the testing is frequently performed in free jet conditions and the
reflections of the shocks return as expansion waves far downstream of the test articles, so the reflections
sketched here for solid walls, do not influence the test results, figure 3.10.
Wall-reflected
Model bow shock
shockwave wave
M = 1.5
M = 1.3
M = 1.1
Figure 3.10: Experimental rhombus variation with the Mach number. [23]
From this relationship it can be seen that the static temperature in the wind tunnel drops with the square
of the Mach number for high Mach numbers. This explains why the reservoirs of hypersonic wind tunnels
are often heated. The same behaviour observation can be explained without the exact equation; just by
re-formulating the square of the Mach number as a dimensionless ratio between the kinetic energy and the
thermal/internal energy of the fluid.
u2 Ėkin
M2 = ∝ (3.16)
a2 Ėinternal
Here, too, the internal energy (expressed as a product of temperature with specific constants or alternatively
as the square of the speed of sound) is inversely proportional to the Mach number. Looking at the equation
above one can see that for high Mach numbers the kinetic energy dominates the thermal energy of the fluid
and this makes it clear why the temperature drops significantly for high Mach numbers. In other words, in
order to achieve hypersonic flow conditions in the test section the internal energy in the reservoir needs to
be very high as all this internal energy will be converted into kinetic energy by the time it reaches the test
section. The high internal energy is provided by heating the fluid in the reservoir. However, this causes a
technical problem as it requires a lot of energy to simulate some flows. For instance, if one would simulate
a reentry vehicle entering the atmosphere one would require to simulate a flow at 100 km altitude with
a velocity of 8 km/s. This would result in tunnel temperatures of 9000 K, which is above the melting
temperature of most materials used to contain the gas. [8].
Pressure control
High-pressure tank
Exit
Support
Spreader Screens
√
a∼ T (3.17)
So one can relate the different speeds of sound for the reservoir and throat by using the following relation
obtained from the fundamental relationships for adiabatic expansion.
a?
r
T?
= ≈ 0.91 (3.18)
a0 T0
For a homogeneous (i.e. ignoring the wave propagation), but low rate of change in reservoir conditions
we require that the information about the outflow be spread effectively and instantaneously overall in the
reservoir vessel. In this approximation the pressure and temperature (p, T ) remain uniformly distributed in
the vessel. The characteristic time of spreading information is aD0 , in which D is the diameter of the reser-
voir. This is the time required to equilibrate the conditions at one part of the reservoir with another, e.g. the
conditions everywhere in the reservoir with those at the exhaust location.
The characteristic time in the discharge scales with the inverse of the mass flux ρ? A1? a? ( kg s
) out of the
reservoir through the sonic contraction. This time is also proportional to the mass of gas contained in the
reservoir ρ0 D3 (kg). Dividing the total mass in the reservoir by the mass flow rate provides an estimate for
the time it takes to drain the tank. Thus, in order to be able to assume a reservoir condition for the outflow, it
remains important that the time scale of the spreading of information is much smaller than the time it takes
to drain the tank, which results in the following:
D
a0 A
ρ0 D 3
∝ << 1 (3.19)
D2
ρ? A? a?
If this condition, which means that the reservoir is large compared to the nozzle throat, is met it means that
the time it takes to equilibrate the reservoir condition is small compared to the time required to empty the
vessel.
It can be assumed that under these conditions and a sufficient pressure ratio that the exit Mach number will
remain constant and determined by the area ratio of the discharge nozzle. Still, the Reynolds number will
vary according to the pressure and temperature drop in the reservoir.
In the case that the above condition is not met, or if the size of the reservoir cannot be described by the
one diameter only, i.e. there are multiple length scales to be considered, wave propagation has to be
t [s]
p4 p [s]
p1
x [m] x [m]
(a) (x, t)-diagram showing the moving shock, contact surface (b) (p, x)-diagram of a shock tube.
and expansion waves.
By providing for a controlled contraction and expansion through a well designed nozzle, the shock tube
becomes a Ludwieg tube, visualised in figure 3.13. In the case for a Ludwieg tube, the tunnel is designed in
such a way to make sure to maintain the sonic conditions at the throat of the nozzle. That means that there
is a standing characteristic, vertical in the t-x plane at that location. This is the tail of the expansion wave
in the above diagram. Because the speed of a wave in the centred expansion fan is given by the following
relation.
γ+1
c = a4 + u (3.20)
2
In which c is the speed of the shock wave (0 m/s at the throat) and a4 is the speed of sound at location 4 in
fig. 3.13. Since the shockwave at the throat is a standing wave it follows that.
2
u= a4 (3.21)
γ+1
It is this combination of the steady and unsteady processes, which leads to the term ‘effective total pres-
sure’. In the case when the nozzle throat is equal to the charge tube diameter the last expansion char-
acteristic will remain in the throat, i.e. the steady expansion sets in from M = 1. At that condition, the
pressure has dropped to 0.28p4 , corresponding to an effective total pressure for quasi-steady experiments
0.28
of 0.528 = 0.53p4 .
Note that the main reason for the implementation of the Ludwieg tube is because all characteristics in the
nozzle are steady. Thus, even though the discharge time is in the order of seconds, the conditions are
favourable as they don’t change in time, at least not up to the point that the reflection of the expansion wave
reaches the nozzle again. This is why Ludwieg tubes are quite long as this increases the testing time. In
the case of a steady expansion, the Mach number is steady throughout time, but other factors, such as
the Reynolds number are not, however these types of wind tunnels can operate for longer periods of time,
which has its advantages as well.
Dnozzle ≈
α 3Lmodel
Lnozzle
sin α ≈ tan α ≈ α can be assumed to hold. The nozzle expansion angle can then be related to the Mach
1
number as α ≈ M and with these relations it is possible to find the minimum required nozzle length, Lnozzle ,
to be,
3Lmodel
Lnozzle ≈ M (3.23)
2
At hypersonic Mach numbers this will have to be complemented by the boundary layer displacement effect.
An approximate estimation of the BL displacement effect can either be obtained from the so-called Mirel’s
u2 γ p u2 a2
+ · = + = const (3.25)
2 γ−1 ρ 2 γ−1
In a blow-down wind tunnel the fluid is accelerated from the rest, thus the first term of the above equation
will be zero since this the initial velocity is zero. This leads to the following.
a20 u2 a2
= + (3.26)
γ−1 2 γ−1
The maximum possible velocity in the nozzle exit plane is then obtained by setting the speed of sound in
that plane to zero:
r
2
umax = a0 (3.27)
γ−1
In the case for air (diatomic gas with constant γ ) this will lead to a maximum velocity of umax ≈ 2.2a0 .
The above observation again illustrates the need not to mix non-dimensional and dimensional parameters
in a single discussion. Although there is a limit to the velocity of the flow at the exit of the expansion, there
is no limit to the achievable Mach number. The Mach number is therefore no expression of velocity. For
simulation of hypersonic flight in atmospheric conditions (such as re-entry from space orbit) this means
that the realistic supersonic conditions can be generated for the ratio of velocities to the local speed of
sound, and the correct ratios of pressure viscous forces are obtainable. The correct energy relationships
for the ratios of the kinetic energy to the molecular excitation energies may however be restricted due to
the restriction of the obtainable velocities. In order to obtain sufficient kinetic energy in the flow, the level
of internal energy in the reservoir has to be raised. The special techniques available for that are subject to
specialist literature.
Question: Explain the plausibility of the above observation in terms of the Mach number as representing
the fractional analysis of energy (fluxes) as done in equation 3.16. Assume that the Mach number is zero
in the reservoir and infinity at the end of the expansion.
u ∞ , p∞ u 2 , p2
d
L2
u2 = u∞ (3.28)
L2 − d2
Once the velocity is determined, the Bernoulli equation can be used to determine the reduction of the back
pressure due to acceleration of the flow at the location of the wake blockage:
2 !
L2
ρ
∆p = u2∞ −1 (3.29)
2 L − d2
2
This pressure is assumed to be uniform over the whole downstream cross-section of the wind tunnel,
including the wake.
The conservation of momentum will, in turn, deliver the drag of a displacement body in the wind tunnel,
where the sum of the momentum flux and pressure forces is equal to the drag D of the model.
This shows, that even in the inviscid and incompressible approximation, a displacement body produces
drag in a closed wall test section. For small blockage ratios, this drag coefficient approaches the blockage
ratio. This blockage drag is the correction that needs to be applied to the measured drag in a closed test
section wind tunnel. In the free air case, where the blockage ratio goes to zero, we obtain the potential flow
solution of zero drag at this level of approximation.
Similar expressions can be obtained adopting different approaches. In the literature (see e.g AGARDograph
336) it is customary to discuss the difference between the measured and the corrected coefficients. For
incompressible flow in a closed test section this may take the following form.
CDunc − CDcorr d2
= 2.5CDunc 2 (3.32)
CDcorr L
1
In the case that the requirement for tunnel accuracy for development tests of 10000 in drag coefficient, it can
be concluded that the correction is non-negligible.
Yet, another formulation for the correction in drag coefficient, now accounting for the compressibility effect
is given in literature as eq. (3.33), where is the blockage ratio, which often is defined in terms of velocities.
Figure 3.17: Static pressure distribution along the tunnel axis. (Figure from DNW)
• Collector Blockage Effects: There is a effect on a bluff model when its wake enters the collector as
the flow is transformed from an open jet boundary condition to a closed wall boundary condition.
The entry of the bluff-body wake into the collector may result in a closed-wall, wake-induced velocity
increment at the model due to the changed constraint on the wake.
• Wake-Induced Effects: When compared to the other effects this is the odd one out. This effect causes
a incremental increase in drag on the model as the solid wall of the collector has its effect on the
flow, which causes a additional drag on the model. Till this time there is no correction found for this
phenomenon.
In the case of an open test section there is no ‘direct’ blockage such as in a closed test section as the air
is free to move, however there is a similar correction as found in equation 3.32. In this case the correction
has to be subtracted rather than added and leads to the following equation [9],
CDunc − CDcorr d2
= −1.3CDunc 2 (3.34)
CDcorr L
mx
φ= (3.36)
x2 + y 2
representing a circular cylinder with infinite span. Without a wake, this represents no loss of momentum
and thus also no drag, though.
In order to explain the lift induced correction, it is useful to recall the mechanism of lift generation. In the
thin airfoil approximation it is the distribution of circulation γ along an airfoil, in the simplest form the lift
generating infinite span airfoil can be approximated by a potential vortex with circulation Γ which generates
the lift L per unit span:
L = ρu∞ Γ (3.37)
This lift can be normalised with the dynamic pressure and a length scale l (area = l · 1), yielding a lift
coefficient
Γ
Cl = (3.38)
u∞ l
No drag is generated in this model (paradox of d’Alambert).
The essence of the lift generation mechanism is the pressure distribution as a consequence of the curvature
of the streamlines introduced by the superposition of the free stream with the circulation. The interference
of the wind tunnel walls/shear layer is then obvious, since the walls stay straight, as a rule. The above
simple observation is valid for a lifting wing of infinite span; an airplane can be approximated by a so called
horseshoe vortex, where the wing tip vortices extend to infinity downstream of the aircraft model.
There are a lot of misconceptions circulating around the generation of lift. One of the most persistent ones
is that a downdraft is required to produce lift. This misconception leads to difficulties in the practice and
interpretation of wind tunnel correction applications.
First, it is important to realize that in the case of level flight, there is no net vertical displacement of mass.
Figure 3.19: Reflection of the circulation on the floor and ceiling.(I. Philipsen, DNW)
superposition of the unperturbed free stream with the model induced and wall induced flow as in equation
3.39.
The boundary conditions then are formulated for the sum of the two perturbations. For the solid, imper-
meable walls, where there is no flow perpendicular to the wall at the wall and thus the wall-perpendicular
gradient of the potential has to remain zero:
∂φ
=0 (3.40)
∂n
∂φ
=0 (3.41)
∂x
This means that there is no pressure gradient along the free jet boundary. The pressure in the plenum is
assumed to remain unaltered along the undisturbed boundary of the free jet. This boundary condition has
to be satisfied at the location of the undisturbed free jet. This can be somewhat misleading in practice. It is
easy to understand the plausibility of the solid wall boundary condition, but it is more difficult to do so with
the free shear layer. In principle, this BC is assumed to hold on the undisturbed boundary, while it actually
allows for the boundary to be displaced.
The boundary conditions can be satisfied by introducing additional potential flow elements outside the flow
field in a symmetric manner for the solid walls and in an anti-symmetric manner for the free jet boundaries.
The reflection of the circulation at the walls can be illustrated by simple schematic drawings, as shown in
figure 3.24. Rigorous solutions exist for only simple cases; the rest will have to be calculated from the
known information.
(a) Curved free stream path due to a lifting surface. (b) (CL − α)-curves for the experiment, the CFD model and
the free air case.
Figure 3.20: Illustration of the deviation of the free stream in an open test section and the resultant difference
from (numerical) free air lift coefficient. [21]
The open test section testing, although in the strict sense of determining the boundary conditions, is inferior
to the testing in closed test sections, is still quite popular for practical reasons. these are ease of access,
lower blockage, acoustic transparency etc. For lifting bodies it might not be sufficient to satisfy the far field
boundary conditions at the location of the undisturbed velocity jump as described above, since the displace-
ment might not be negligible, as shown in the figure 3.20.
In figure 3.20, we also note that the upstream and downstream deflection of the flow are not symmetrical, as
one would expect to be the case of potential flow approximation where, in order to satisfy the constant pres-
sure boundary condition, an updraft of equal value is generated upstream as the downdraft is downstream.
This is illustrated in the following Figure, and can also be deduced from Figure 3.24 b.
The reason for the lack of symmetry in Figure 3.20 can be found in the upstream boundary condition that
is present in the open-jet wind tunnel, which differs from the model used in the previous discussion. The
flow field produced in the wind tunnel is a jet of finite extent and thus does not extend from −∞ to +∞ as
assumed in the previous description. In the open jet wind tunnel case, the flow potential and its gradients
(i.e. velocity components) are prescribed (determined) at the exit plane of the nozzle contraction.
Approximating a 2D lifting surface in those conditions as a potential flow singularity, expressed through
circulation Γ, requires consideration of the distance between that lifting surface and the exit plane of the
nozzle contraction. This gives rise to a correction on the effective angle of attack of the lifting surface, as is
explained in the following sequence of approximative treatment.
We us the standard expression for sectional lift in 2D potential flow, as
L
= ρu∞ Γ (3.42)
s
with s denoting the spanwise distance, and obtain the standard description for the induced flow ui at any
th
distance r from the singularity; in our case from the 14 chord location of the airfoil.
Γ
ui = (3.43)
2πr
On the other hand, one can find the relationship between the value of the circulation and the lift coefficient
cL of a thin airfoil (not accounting for separation):
L cl ρu2∞ sc 1
Γ= = = cl cu∞ , (3.44)
sρu∞ 2sρu∞ 2
In order to relate the lifting performance of the lifting surface in the open-jet wind tunnel to that in the ideal
situation in an unbounded flow field, the change in inflow conditions imposed by the boundary condition at
the nozzle exit plane needs to be determined. The situation is sketched in Figure 3.22.
In an unbounded flowfield, the velocity induced by the lifting surface at the location of the nozzle exit plane is
obtained by combining equations 3.43 and 3.44, making use of the thin airfoil approximation in 2D, cl = 2πα:
cL c α c
ui = · · u∞ ∼
= · · u∞ (3.45)
4π r 2 r
The distance r denotes here the distance between the nozzle exit plane and the lifting object in the open
test section and the angle of attack α is the one in the potential flow frame, in unbounded flow the geometric
angle of attack.
In the open-jet wind tunnel, however, the flow is straight at the nozzle exit plane. Accounting for this nozzle
exit plane condition as a requirement for the entrance plane of the control volume, and matching that to the
potential flow model of the lifting surface, requires a rotation of the effective inflow that the lifting body is
exposed to. This rotation is equal to the induced angle of attack that follows directly from equation 3.45:
ui α c
αcorr ≈ = · (3.46)
u∞ 2 r
This last expression allows one to determine the effective angle of attack of the airfoil – it is the geometrical
angle of attack minus the induced angle of attack imposed by the boundary condition at the nozzle entrance:
The correction for the induced angle of attack (equation 3.46) is directly proportional to the geometrical
angle of attack; hence the linear deviation of the two cl curves in Figure 3.20 it is also inversely propor-
tional with the distance of the airfoil from the nozzle exit plane. In the case of the nozzle exit plane being
at infinite distance upstream, the standard potential flow solution will be obtained. The last observation re-
sults in a clear recommendation not to test too close to the nozzle exit in open jet test section configurations.
As an illustration of the principle of the correction, the above Figure 3.21 has to be split into two. The first
part represents the nozzle exit plane where the inflow condition to the control volume is still horizontal and
homogenous. The second part adapts the potential flow field to that boundary condition.
The above approximation was simplified for a lifting body with zero lift at zero angle of attack. For cambered
airfoils like the one illustrated in the Figure 3.20 above the zero lift angle of attack will have to serve as the
reference instead of α = 0. Furthermore, the modelling relies on the small perturbation approximation, like
all corrections. This means that the rotation angles remain small, that the sin(αcorr ) = αcorr and tan(αcorr )
= αcorr and that cos(αcorr ) = 1.
In practice, it is customary to express the interference through the non-dimensional parameters of
ui
ε= (3.48)
u∞
(a) Reflection of a lifting vortex on the wind tunnel floor with a resultant horizontal
velocity component.
(b) Anti-symmetric reflection of the lifting vortex on the free shear layer with a
resultant vertical velocity on the undisturbed boundary.
Figure 3.24: Lift corrections for closed and open test section. [9]
where ui is the axial flow component induced by the wall/shear layer boundary condition, and
wi
∆α = (3.49)
u∞
where wi is the wall perpendicular (vertical) velocity component induced by the boundary conditions. Or
∂δ
δ1 = (3.51)
x
∂ βL
√
where β is the compressibility factor β = 1 − M 2 . Further, the same potential flow approximation can
be applied to the displacement effect of the model, where the solid displacement can be obtained by a
judicious application of source/sink doublets as shown in the illustrations in figure 3.25 below.
Figure 3.25: Image reflection to satisfy the boundary condition for displacement effect. [9]
For the 2D lift interference, when the wing is approximated as a line vortex with the potential
Γ βz
φ= atan (3.52)
2π x
the integration over all reflections and from infinity upstream to infinity downstream can be performed to
obtain the results for the upwash interference as graphically presented in figure 3.26. Note, that the upwash
is negative upstream of the singularity and positive downstream. This is of importance, when not only the
lift coefficient, but also the pitching moment needs to be corrected, because the lift correction introduces
another error in the moments. Note the importance of positioning of the model representation. In the 2D
case, the area ratios are replaced by the ratio of the wind tunnel height and the chord length. For the
streamwise interference a remarkable result is obtained as there is no interference along the centreline of
the tunnel, if that is the location of the lifting surface! The strongest interference occurs at the location of
the tunnel walls. This can be explained by the opposing signs of the induced velocities of the reflected
circulations, see figure 3.27. Further cases can be obtained from the literature. All these rely on the
Figure 3.27: Interference in free stream velocity normalised with the lift coefficient. [9]
principles of linear superposition, which means that only small perturbations are permissible. An example
of the importance of corrections for a 2-D measurement, where a more than 50% error is shown to be
correctable is illustrated by Barlow, Ray and Pope in figure 3.28.
Question: what happens to the stability mechanism that is represented by the tail plane?
1 h Γ Γ
uinterf erence ≈ − (3.53)
2π l l l
The total interference behaviour as a function of the tunnel aspect ratio can be plotted as in figure 3.29. The
same principal approach as applied to the 2D corrections can also be applied to finite width lifting bodies.
The line singularities are then replaced by horseshoe vortices. The handling of those requires a more
complex mathematical approach, but the principles of reflections are maintained. Corrections for massively
separated wakes rely on the quadratic relationship between drag and lift (See Barlow Rae and Pope) and
make use of the relationship
Figure 3.29: Interference factors for classical correction for a small model in the centre of the wind tunnel. [9]
which can be plotted as a straight line against the quadratic lift coefficient. Where the measured relationship
deviates from the above relationship, separation (stall) is assumed to be responsible and a strong wake
ensues. This wake can be modelled as one caused by a combination of sources and sinks. The velocity
Q
increment in a tunnel of width B and height h can be expressed as ∆u = 2Bh , where Q is the source
strength. This can be related to the drag and the wing aspect ratio Span/Chord (S/C):
∆u S
ε= = CDuncorrected (3.55)
u 4C
In the literature, this is further split into the components due to the constant viscous drag, the induced drag
quadratic in lift, and the wake blockage:
S 5S
εwb = CD0 + (CDuncorrected − CDi − CD0 ) (3.56)
4C 4C
The term in the brackets vanishes, when there is no separation, leaving us with one of the standard wake
blockage corrections. There are other similar approaches, but this, known as the Maskell correction, is
widely used.
(a) Control volume for analysis of a propeller. (b) (u, x) and (p, x)-diagrams.
conservation (while adding momentum) and therefore the continuity equation remains valid, thus:
S∞ u∞ = Sp up = Se ue (3.57)
If a momentum source or sink is introduced in the wind tunnel measurement inconsistencies will occur, as in
normal operation there are no sources and sinks within the wind tunnel measurement volume. Because of
these the flow within the tunnel is either increased, in the case of a momentum source, or decreased, in the
case of a momentum sink. There are two models that are often used for these thrust and drag corrections.
The first goes back to Glauert, and relies on obtaining an equivalent free stream velocity for the given thrust,
where the thrust is estimated from the momentum equation for the free stream on the one hand and the
combination of pressure and momentum consideration in the closed wall wind tunnel on the other. For the
conventional actuator disk model, the ratio of the corrected velocity over the tunnel velocity is expressed as:
S
ucorr Tc p
= 1 − √ Swt (3.58)
u 2 1 + 2Tc
The last expression is derived for the conventional actuator disk model, whereas in the further modelling the
contraction behind the propeller will be neglected. For large Tc ’s the above expression yields a correction to
the free stream velocity that is proportional to the product of the area ratio and the square root of the thrust
coefficient.
The other method, somewhat simpler but yielding the same result views the propeller as a singularity. For
the far field, the propeller acts as a sink with strength Q:
Q
φ= ln(r2 ) (3.59)
4π
Resulting in velocity correction
Q
∆u∞ = (3.60)
4πr · r
r !
8
Q = u∞ (S∞ − Sp ) = u∞ Sp 1 + Tc − 1 (3.62)
π
From the last expression, a correction velocity to the far upstream field can be calculated as
q
u∞ Sp 1 + π8 Tc − 1
∆u∞ = (3.63)
4πB · H
Here the velocity correction is assumed to be an additional component to the free stream velocity in a
constrained cross section of dimensions H and B, and the free stream is assumed to be uniform over that
cross section. Again, as in the Glauert method, for large Tc the product of the area ratio and the square root
of the thrust coefficient determine the correction. The above approach means, that in a constrained wind
tunnel, the upstream velocity at a given thrust is higher than it would be in an unconstrained inflow, thus the
corrected velocity for the measured thrust for unconstrained space is
ucorr = u − ∆u
ucorr ∆u (3.64)
=1−
u u
It is customary to plot this relationship in the inverted way, as shown in figure 3.31. This corresponds to
the perception that for the measured thrust will be lower than that in an unconstrained space. This can be
u
( ucorr )
evaluated to provide a corrected free stream velocity, which for a given value of thrust results in the same
axial velocity at the propeller location as that observed in the wind tunnel. Here, the corrected velocity ucorr
is the sum of the tunnel velocity and the area ratio is a parameter. Yet another alternative approach can
be taken to compare the pressure effect upon the back plane of the propeller. Due to the acceleration of
the flow in the slipstream, the conservation laws require the flow outside of the slipstream to slow down
to maintain continuity between the tunnel walls. Again, since the slipstream velocity is a function of thrust
(effectively, it is a consequence of the inflow stream tube contraction), a relationship between the additional
The inverse approach can be taken to correct for a windmill test in a closed test section. In a windmill case
one can assume that the inflow is unadulterated; the windmill acts as a source for the downstream part of
the actuator disk, leading to an expanding wake. If we adapt the symmetrical approach, then the additional
(positive) velocity far downstream of the windmill will be the same as in the inlet of the propeller:
q
u∞ Sp 1 + π8 Dc − 1
∆u∞ = (3.65)
4πB · H
On the other hand, we know that the drag of the windmill can be expressed as the product of mass flux with
the velocity increment, here deceleration.
D = ṁ∆u (3.66)
For a given windmill drag, with a fixed inflow condition determined from the free stream parameters, one
can express the velocity increment as
where both the last terms are functions of the (measured) drag. The second term is the slipstream velocity
that would establish itself in the unconstrained space; the third term is that correction of the slipstream
which is due to the constrained space. For a given drag, this would also mean that the measured drag
corresponds to a free stream with a velocity
All of the derivations above have been performed for a closed test section, however it can also be done for
an open wind test section. For a propeller, with a contraction in the stream tube upstream of the propeller
plane, this means that ambient air has to be entrained from the plenum area, where the air is assumed to
remain still. The amount of air is determined from the strength of the sink, as was described in the case of
the closed wind tunnel:
r !
8
Q = u∞ (S∞ − Sp ) = u∞ Sp 1 + Tc − 1 (3.69)
π
We now assume, that the work done to entrain that fluid can be ascribed to the inflow stream tube of the
propeller actuator disk, and thereby the stream tube averages between the entrainment form the ambient
still air and the air that exits the wind tunnel contraction. In reality, there is also a layer of wind tunnel air
between the inflow stream tube and the propeller. The simplest approximation accounting for that is to
assume that the inflow velocity condition in the stream tube is a weighted average:
Stunnel
uinf low = u∞ q
Stunnel + Sp 1 + π8 Tc − 1
(3.70)
1
= u∞ q
Sp
1+ Stunnel 1 + π8 Tc − 1
T = ṁ∆u (3.71)
ṁ = ρu∞ Sp (3.72)
This thrust can be compared with the thrust that would correspond to the unbounded domain, where ∆u =
ue − u∞ . From this comparison, the velocity correction or the thrust correction can be obtained: for a given
thrust, the tunnel inflow velocity has to be compensated for the sink effect of the propeller. We measure
thrust that corresponds to a lower inbound velocity than provided by the tunnel. Or, inversely, the measured
thrust in an unbounded free jet would be smaller than it is obtained in a bounded free jet wind tunnel, see
figure 3.32. Tunbounded < Tbounded
u∞
Figure 3.32: (Tc , J)-diagram for a propeller. (J = ωD )
Next, the question can be turned around for a windmill. Here the assumption is then made of an unadulter-
ated inflow and the source effect in the wind turbine wake slipstream. For the windmill case in a bounded
free jet, we return to the stream tube approximation again. In this approximation we ignore the upstream
influence of the wind turbine and assume that all the work is again done by the suction side. From experi-
mental evidence we conclude that this approximation is about as far from reality as is the approximation of
even distribution of the deceleration before and after the actuator disk, see figure from MEXICO project [5],
th th
where approx. 13 of the deceleration is done before the turbine and 23 is done downstream of it. Impor-
tant is that the assumption of velocity continuity across the disk is maintained and the drag is given by the
pressure jump across the turbine plane as actuator disk. In this case, we can apply the Bernoulli equation
from the rotor downstream, assuming that in the far wake the ambient pressure prevails.
u2∞ u2
ρ + p2 = ρ e + p∞ (3.74)
2 2
Together with the incompressible mass flux conservation, we can express the slipstream area far down-
stream as
1
Se = √ Sp (3.75)
1 − Dc
u2
Where the drag coefficients are defined as Dc = SpDq∞ , with q∞ = ρ 2∞ . Note that the thus defined drag
coefficient Dc cannot exceed 1! (For similarity with the propeller, where the expanded part is the reference
p u2
Se = 1 − Dce Sp , where qe = ρ 2e .) The second expression is just to demonstrate the similarity between
the propeller producing thrust and the turbine producing drag; it is not customary to normalise drag with
the velocity in the slipstream. With the above equation, the source strength for the wind turbine can be
determined:
Q = u∞ (Se − Sp ) (3.76)
This is the flow that is forced out of the bounds of the open jet configuration into the nominally still air at
ambient conditions. It is obvious that this is more work than forcing the air into co-flowing stream. When
we again, form a weighted average of the slipstream and the ambient flow (actually just assume that the
stream-tube has been widened by the influence of the source) then we can define a corrected free stream
velocity
1
ucorr = u∞ (3.77)
Sp
1+ √ 1 −1
H·B 1−Dc
This corrected velocity can be used for determining the corrected tip speed ratio as well as the corrected
drag coefficient belonging to the corrected tip speed ratio. This corrected free stream velocity accounts for
the fact, that we require the wind tunnel circulation to conserve mass. In the end, all the flow exiting the
wind tunnel has to again enter it. We account for that by reducing the flow rate out of the contraction to
match the flow rate entering the collector. In reality there is also a change in the plenum pressure due to
the spreading (source) effect of the wind turbine that has to be accounted for in the flow reference system,
but for a first approximation the proposed methodology will do. The new tip speed ratio will then defined as
πωD
λ= (3.78)
ucorr
and the new drag coefficient will be
D
Dc = u2corr
(3.79)
Sp ρ 2
This means, that the measured drag appears lower than it would be in an unbounded flow field, Dunbounded <
Dbounded , see figure 3.34.
CD ∼ D
1
λ∼ u
Figure 4.1: (L, M )-diagram for transonic testing. Dashed line represents the operating conditions and the
solid line represents the limits.
It can be seen that near the cruise condition the drag rise and the buffeting (irregular oscillations of part of
an aircraft due to periodic variation of the aerodynamic forces) phenomena need to be considered. Both
phenomena are related to the local supersonic flow being terminated by a shock. On the other hand the
maximum lift coefficient, CLmax , is the limiting factor for lower velocities such as during take-off and landing.
As the behaviour of a aircraft in free flight is different from the model in the wind tunnel it is necessary to
clarify how these limiting boundaries translate from wind tunnel test conditions to free flight. To do this an
understanding of scaling effects is required.
When testing aircraft models in a wind tunnel it is important to identify different wing regions and profile
characteristics as different regions react differently to the flow as proposed in figure 4.2.
The Figure 4.2 separates the wing and the wing root into regions that may have different dependencies upon
the Reynolds number. Whereas perts of the wing can be treated almost as 2-dimensional flow cases where
the transition Reynolds number dominates the flow behaviour, the tip and root regions are dominated by the
formation of root and tip vortices responding to different Reynolds number dependencies. So extrapolation
to flight conditions is different for different regions. Different components of the aircraft may exhibit different
sensitivities to the variation of the Reynolds number. This sensitivity analysis is most important for the
determination of the CLmax , the cruise drag and the drag rise that occurs when the aircraft is flying close to
the speed of sound.
In case of the determination of the CLmax one can see that the unsteadiness of the flow to increases
when measurements are performed close to the maximum lift coefficient as illustrated in figure 4.3. This
unsteadiness can be observed at low speed as well as in the transonic regime when it occurs as buffeting.
Figure 4.2: Isolation of wing in various regions with varying profile characteristics.
Which detailed phenomena are involved and how this is affected by the global Reynolds number, will be
discussed later, here a characteristic illustration of the unsteadiness of the stress measured by a strain
gauge at the root of the wing is used as an illustration of the boundaries that need to be considered.
The second of those limiting conditions is the drag rise or drag creep. This is the limiting effect for cruise
performance and thus indirectly also for the fuel efficiency of an aircraft. The determination in an experiment
of the drag creep is not always straightforward and is usually based on a number of indicators. The indicators
used to determine the drag creep boundary is illustrated in the figure below for three indicators. These are
listed below and visualised in figure 4.4.
• The lift divergence observed at constant angle of attack for increasing Mach number.
• The angle of attack divergence at constant lift coefficient.
• The change in the trailing edge pressure indicating separation.
The divergences of the angle of attack and of the lift have to do with the transonic conditions and the
change of the shock position as a result of the boundary layer characteristics. Some of the ways to avoid
misinterpreting the experimental results due to the scaled Reynolds numbers will be covered in the following
sections. First the limits and the governing phenomena for the 2D airfoils and the large aspect ratio wing
Figure 4.4: Occurrence of lift divergence and drag creep close to sonic conditions.
configurations will be treated. These airfoils and high aspect ratio wings behave in a manner analogous to
the 2D phenomenology. Later on low aspect ratio phenomenology will be introduced.
4.1 Reynolds number effects on airfoils and high aspect ratio wings
4.1.1 Background
The Reynolds number effects of airfoils and high aspect ratio wings are grouped together as in both cases
the flow experienced over the airfoil or wing is practically two-dimensional or very weakly three-dimensional
(such that local strip theory, essentially the negligibility of the 3rd dimension, still applies). In figure 4.5 the
influence of the boundary layer character, and thus also the Reynolds number, upon the flow over an airfoil
in transonic speed range is schematically illustrated.
The Reynolds number affects the relative location of the transition of the boundary layer from laminar to
turbulent and consequently the thickness of it under the shock wave as well as at the trailing edge of the
airfoil. The dependence of these three geometrical parameters upon the Reynolds number are termed as
direct Reynolds number effects. The consequent changes in the flow field, like the position of the shock
as a consequence of the change in the circulation due to the thickness change of the trailing edge, are
referred to as indirect Reynolds number effects. In this figure separation and stall are not shown, nor are
the different mechanisms depicted in detail. In all of these direct Reynolds number effects the Mach number
is another governing parameter as it determines the pressure effects. See the discussion of the viscous
interaction parameter and the SWBL interaction in the chapter about scaling and similarities. Some of these
more detailed issues will be dealt with in the following sections.
4.1.2 Transition
When designing aircraft, knowing where and how transition occurs is important because it affects the drag,
and thus fuel consumption, of an aircraft significantly. In figure 4.6 the skin friction coefficient as a function of
Reynolds number is described for both laminar and turbulent flows. In this figure the effect of the Reynolds
number on the skin friction coefficient has been shown for a flat plate without longitudinal pressure gradients.
From the graph it can be seen that there is a range of Reynolds numbers with a simultaneous possibility of
having both laminar and turbulent flow. Factors other than the Reynolds number determine whether either
laminar or turbulent flow will occur. It can be seen that the turbulent friction coefficient is significantly higher
than the laminar friction coefficient which means that if possible, laminarity should be maintained for the
longest stretch possible in order to reduce the viscous drag over an object. Since it is inherently difficult
to maintain laminar flow for large Reynolds numbers and thus for a long stretch, only high quality testing
conditions allow testing the full extent of the laminar BL on an airfoil.
Typically a laminar boundary layer up to 60% of chord is possible with appropriate design of the airfoil for
the cruise Mach numbers. In good quality wind tunnels it is possible to demonstrate the laminarity of the
boundary layer, on specially designed laminar wing profiles, up to the shock induced transition.
The presence of the domain of simultaneous flow regimes is indicative of the importance to obtain and
quantify information about the turbulent characteristics of the free stream in a wind tunnel; in the flight
conditions the free stream is assumed to be disturbance free.
The above classical illustration was obtained in the 2D parallel straight flow conditions. The transition to
turbulence occurs via the Tollmien-Schlichting mechanism of instability. In the swept wing condition, the
flow can be much more complicated, as illustrated in figure 4.7. The component of the velocity parallel to
the leading edge of the wing for the swept wing is not always contributing to the pressure change, although
it can significantly contribute to the transition, through the cross-flow mechanism as will be discussed later.
In figure 4.7 the external ‘potential flow’ streamline is shown as smooth curve with no strong gradients in
it. In the boundary layer adjacent to the airfoil surface a completely different situation can be observed.
The streamlines can be ‘attached’ or ‘separated’ and their direction can strongly differ from the mean flow.
Not only is the flow direction affected by the sweep of the airfoil, also the transition to turbulence can
occur via different dominant mechanisms. A number of mechanisms is illustrated in figure 4.8. There is
enough evidence, though, that with modern civil transport aircraft, with wings at 37◦ sweep, the transition
is dominated by cross-flow instabilities or by the stagnation line contamination. The theoretical limit for the
Tollmien-Schlichting instability as a transition mechanism for non-zero pressure gradient boundary layers
(the Granville criterion) is preceded by other mechanisms in the case of sufficiently high Reynolds numbers.
Even for the Tollmien-Schlichting mechanism the Reynolds number is not the only transition criterion, also
the leading edge curvature, describing the rate of acceleration at a given Reynolds number, influences the
transition mechanism.
In figure 4.8 four Reynolds numbers are shown (Rx ) each of which signify different transition mechanisms
briefly described below.
R1 : At R1 the transition follows directly from the Tollmien-Schlichting instability.
R2 : At R2 leading-edge contamination followed possibly by re-laminarisation and then transition through
Tollmien-Schlichting instability.
R3 : At R3 leading-edge contamination followed possibly by re-laminarisation and then transition through
cross-low instability.
R4 : At R4 transition follows from cross-flow instability if leading-edge contamination were absent.
In realistic cases of current generation of airfoils, however, the transition in free flight occurs at approximately
5% of the chord length. Since the transition determines the characteristics of the boundary layer over the
rest of the chord length it is important to treat that part of experimental simulation with particular care. It
also affects the maximum lift performance, the stall characteristics, the buffeting and the separation at the
transonic shock location.
(a) Graphical interpretation of various types of stall for (b) Experimental data of various types of stall for the
different airfoils. same airfoil.
In the figure a clear distinction emerges between a thin and a thick airfoil behaviour. Thin airfoils tend to
stall at the leading edge first, though the definition of thin and thick itself is a Reynolds number dependent
category, as shown in figure 4.10. At low Reynolds number the chosen airfoil behaves like a thin one and
also displays a rather low value of maximum lift. As the Reynolds number is increased by an order of
magnitude from the baseline, the same airfoil displays all the characteristics of a thick airfoil with the stall
being initiated at the trailing edge. At the intermediate Reynolds numbers the CLmax increases continually
and the separation characteristics vary.
Furthermore, it can be observed, that for the different stall mechanisms, not only the Reynolds number
along the airfoil, but also the characteristic radius of the leading edge of an airfoil plays a role. Although
the radius of curvature can be used as the relevant length scale in the formulation of a second Reynolds
number, the ratio of the radial acceleration with the linear inertia force might be a more appropriate scaling
factor. The similarity of that is automatically given for geometrical similarity of models. In the figure below,
this similarity (nose bluntness) is expressed in terms of the ratio of the airfoil thickness at 1.25% of the chord
to the chord itself. One can observe that not only the pure forms occur, but also combinations thereof can
occur as well.
In figure 4.10 the domains of different separation regimes are indicated in a plot of leading edge curvature
vs. free stream Reynolds number. This figure is an impressive illustration of the fact that it is not the global
Reynolds number alone that determines the behaviour of the flow field.
The separation will become even more of an intricate interaction when the flow becomes transonic. The
shock wave-boundary layer interaction phenomenon was discussed in the lectures for dimensional analysis.
The dimensional analysis provided information and delineation about the separated domain boundary, but
not about the reattachment, i.e. the possible existence and size of a separation bubble. The shock wave-
boundary layer interaction is illustrated further in figure 4.11. The pressure rise downstream of the shock is
communicated upstream via the boundary layer, leading to an advance compression of the free stream and
a thickening of the boundary layer upstream of the shock. Thus it is a combination of viscous (Reynolds
number) and inviscid (Mach number) effects.
(a) Type A at expanding bubble stage. (b) Type B at early stage with localised (c) Type B at later stage with one sep-
shock bubble and rear separation. aration extending from shock to trailing
edge.
In the figure about the shock induced speration, the separation is split up into two classes. The type A
and the type B, both eventually leading to flow breakdown. A type A is one where the separation starts as
a bubble under the shock and eventually grows to a length exceeding the trailing edge distance from the
shock. This type separation start is a weak function of the Reynolds number and more dependent on the
Mach number.
In type B separation the shock induced separation starts with a bubble either simultaneously with a trailing
edge separation or is wholly dominated by a trailing edge separation which reaches upstream to the shock
position. Many rules exist for predicting the start of the flow breakdown. A rather simple one correlates the
pressure gradient at the trailing edge with the momentum thickness of the boundary layer at that location.
where Θ is the momentum thickness of the boundary layer and thus Reynolds number dependent. This flow
breakdown criterion indicates that a given profile will experience an earlier breakdown at lower Reynolds
numbers. This means that the wind tunnel test will tend to over-predict the separation. The position of the
shock itself can react in a very sensitive manner to variations in boundary conditions both upstream as well
as downstream of it. Whereas for standard supercritical airfoils the shock position can be expected in the
vicinity of the 60% chord, in low Reynolds number (thick boundary layer) experiments the shock position
tends to move forward.
to be checked for the lowest Reynolds number and usually at max lift conditions. A well optimised transition
strip should trigger the transition from a laminar to a turbulent boundary layer without additional detrimental
effects like:
• An extra increase in boundary layer thickness (an unwanted deterioration of the boundary layer).
• The generation of streamwise vortices similar to the effect of vortex generators that re-energise the
boundary layer
A rule of thumb for trip application is to use the trip Reynolds number, Rek .
kulocaledge
Rek1 = (4.2)
ν∞
Where k is the size of the grid used. In general good tripping conditions occur when the trip Reynolds
number is approximately 600, Rek ≈ 600. A more refined and also more physical version of the grid size
requirement is given in equation 4.3. In this definition of the trip Reynolds number the skin friction coefficient,
Cf plays a role.
The velocity has to be the free stream velocity. Note that the criterion using the square root of the skin
friction coefficient has an approximate value of the square root of the criterion 4.2. This is further explained
by Barslow [4].
q
Cf
kuedge local 2
Rek2 = ≈ 25 (4.3)
νlocal
An example of the effectiveness of the grit size is shown in figure 4.13, where the measured drag is com-
pared with a theoretical estimate for a fully turbulent boundary layer. In this figure, it is obvious that at higher
Reynolds numbers for the free stream, smaller trips become efficient. Large transition strips on the other
hand generate their own drag, which should be accounted for. It should be noted, that in addition to its ben-
eficial effect, the transition due to tripping can affect the result of a drag measurement rather significantly
as shown in figure 4.14.
4.2.3 Effect of tripping at different locations along the chord of the airfoil
The previous examples were used to demonstrate the effect of tripping the airfoil boundary layer near the
leading edge of an airfoil, around the location where it would occur in free flight Reynolds number condition-
s. An interesting alternative to tripping the boundary layer at the location where it occurs in free flight is the
possibility to trip at such a location as to obtain the relative boundary layer thickness corresponding to free
flight at the trailing edge of the model airfoil. The principle of this method, also termed aft fixation, is illus-
trated in figure 4.15. This approach can be recommended, when the trailing edge-dominated phenomena
are of significance, and when the Reynolds numbers are too low to trust the effectiveness of the leading
edge tripping in the experiment. We have applied that approach in the small tunnels of the TUD (M-Tunnel),
to avoid trailing edge separation.
The positioning of the strip forcing transition can lead to apparent shifting from one type of separation to the
other one as illustrated in figure 4.16, where the dependency of the maximum lift as well as the characteristic
shape of the lift curve upon the tripping location is shown for three different Mach numbers. If the maximum
lift is also influenced by the occurrence of the buffet phenomenon, then a more complicated development of
the maximum lift can be observed, see figure 4.17. For the lowest Mach number the shock is closer to the
leading edge and the turbulent BL has a detrimental effect.
Apart from changing the maximum lift coefficient of wings, tripping also affects the shock location. In figure
4.18 the sensitivity of the shock location to the location of tripping is also displayed. There are two different
locations obtained for the free transition case and three locations for forced transition case. In the forced
transition it is shown that tripping too far upstream will move the shock wave location upstream as well.
The mechanism responsible for the forward movement of a shock is the thickening of a boundary layer. If
Figure 4.16: Effect of transition fixation on lift development on a supercritical airfoil for various Mach num-
bers.
the boundary layer is tripped the thickness of the boundary layer increases as turbulent boundary layers
are thicker, with respect to laminar boundary layers. Apart from being more prone to separation, these
thicker boundary layers cause the effective shape of the airfoil to change which reduces the circulation.
This phenomenon is often referred to as decambering.
However, one should alway pay attention to the set up, as shown in figure 4.19. It shows a comparison
between free flight data at Reynolds numbers of 20 · 106 in comparison with transition fixing at 30% of chord
when tested at Reynolds numbers an order of magnitude smaller. In this comparison two different angles
of attack are shown. One, where the transition tripping produces a pressure distribution very much like the
Figure 4.18: Effect of transition point changes on the pressure distribution of a supercritical airfoil.
one from the corresponding free flight test, whereas at a higher angle of attack (and thus more circulation)
the agreement significantly deteriorates.
The domain, where aft fixing might be recommended is given in figure 4.20, where the domain is separated
into 3 regions for forward, middle and aft fixing of transition. This sort of graph should only be used as an
indication and not a prescription. The experimenter will always have to verify the strategy and compare with
alternatives.
Figure 4.19: Comparison free flight data with simulations where transition fixing at 30% of the chord was
applied.
namics delta wings represent a class of flying vehicles which can be described as having low aspect ratio
wings with a highly three-dimensional flow development characterised by free vortex flow, highly skewed
shock waves and/or three-dimensional separated regions. In this section experiments will be discussed in
(a) (Cp − η)-curve, α kept constant. (b) (Cp − α)-curve, η kept constant.
Figure 4.23: Reynolds number effect on the vortex generation of delta wings.
figure 4.23a that the peak in −Cp increases for higher Reynolds numbers. This decrease in Cp is caused
by vortex generation at the wing’s leading edge. As the vortex introduces and additional velocity compo-
nent the static pressure decreases as, according to Bernoulli the sum of the static pressure and dynamic
pressure must be constant. Though the vortex position is mainly determined by the free stream conditions
and the separation location by the curvature of the leading edge, a Reynolds number dependency is still
found in the size of the vortex. The size of the vortex is evidenced by the change in the pressure distribution
characteristics on the upper surface of the delta wing.
However, not all measured effects can be explained with the term Reynolds number effect as not all of
the highly complicated phenomena scale directly with the Reynolds number. In the case of delta wing
testing at transonic speeds this is shown in figure 4.24 It can be seen that doubling the Reynolds number
leads to fundamental changes in the flow features in the opposite direction that was expected according
to the previous figures. It was anticipated that an increase in Reynolds number would cause a suction
peak, whereas in this case suction almost entirely disappears as the Reynolds number increases, with the
exception of the suction peak at the leading edge of the wing.
Figure 4.25: Typical flow conditions visualised by the hydrogen bubble technique.
It can be seen that under high angles of attack the flow around a slender body behaves similarly to a
cylinder. This means that if a slender body is assumed to behave like a cylinder the several flow regimes
can be discerned that are shown in figure 4.26, where RD is the Reynolds number based on the diameter of
the cylinder. In figure 4.27 other flow characteristics about a circular cylinder are shown. In addition to the
cross flow effect occurring on all cylinders, the aspect of vortex separation on the side edges resembles that
of the highly swept wings. The same sort of transverse pressure gradient on both sides of the symmetry
plane produces a component of downstream directed vorticity on the upper surface with the opposite sign
to the equatorial plane vorticity, leading to converging streamlines between those two lines on the body.
The radius/diameter of the relevant geometry is a suitable length scale for the Reynolds number scaling
in experiments, if at all. The ratio of the angular to linear acceleration might have more influence on the
performance.
(a) Lift fluctuations (r.m.s.). (b) Strouhal number of the lift fluctuations.
transonic behaviour with turbulent transonic boundary layer behaviour, as the following two examples illus-
trate, see figure 4.29. From figures 4.29b and 4.29c one can see that the shock-boundary layer interaction
for a laminar boundary layer results in different flow conditions than for a turbulent boundary layer. As the
pressure distribution of the laminar profile is heavily dependent on Reynolds number, whilst for the turbulent
boundary layer a change in Reynolds number hardly has an effect on this pressure distribution.
(b) Shock-laminar boundary layer interaction (Reynolds number (c) Shock-turbulent boundary layer interaction (Reynolds number
effect). effect).
This equation states that the net momentum flux across the boundaries of a control volume has to be
balanced by the sum of the external forces, see figure 5.1. The external forces in general are both the
forces applied to the mass as well as the forces applied to the surface. The latter include the pressure
force, which is sometimes forgotten.
ṁ
u∞
u∞ usl
u∞
ṁ
Figure 5.1: Illustration to the application of the balance of momentum across a control volume including a
propulsor.
The illustration above can be used to derive the expression for thrust, without any specific knowledge about
the thrust generation mechanism inside the control volume. The inflow plane is described by uniform flow;
the outflow plane includes the accelerated slipstream emanating from the propulsor. It is assumed that
the entrance plane and the exit plane are in sufficiently far field, such that the static pressure in those
planes is that of the ambient air. The pressure forces on the top and bottom (symmetrical) planes of the
sketched control surface can be ignored, because of the antisymmetric geometry. In the thus specified
Here S∞ and Ssl are the cross-sectional areas of the stream tube through the propulsor upstream and
downstream respectively. This mass flux produces a momentum force according to the above integral
equation. The component parallel to the free stream, the x-component, of this force is
By summing all the momentum forces acting upon the control volume in the x-direction and including the
thrust force T , which compensates for holding the engine in place, it is possible to write:
Thanks to the continuity equation it is known that ρ∞ u∞ S∞ = ρsl usl Ssl = ṁT as the mass flow through the
stream tube must remain constant, which leads to:
Although the above equation was obtained from the momentum balance across all surfaces of the control
volume, the result contains only inflow and outflow conditions. It is customary to define the inflow part,
ṁT u∞ , as the ram drag and the outflow part, ṁT usl , as the gross thrust. This is a somewhat arbitrary
separation of the two components, because one would not exist without the other. In the case that a
turbofan is generating the thrust, the controlling parameter is the Fan Pressure Ratio (FPR), which will be
defined later.
If it can be assumed that the above description of thrust remains valid even when the engine is mounted
on an aircraft, i.e. the interaction between the propulsor and the airframe is negligible. In such a case, and
only in such a case, the propulsion efficiency η can be defined. This efficiency is the ratio of the work done
to overcome the drag of the aircraft without the engine and the amount of work done for to accelerate the
free stream flow to the slipstream velocity in the engine. If it is assumed that the thrust generated by the
propulsor is equal to the drag of the airframe, then the efficiency becomes:
T · u∞ 2u∞ 2
η= = = (5.6)
ṁT
2 (u2sl − u2∞ ) u∞ + usl 1 + uu∞
sl
Note that by applying this formulation to any other situation, such as to a Boundary Layer Ingestion (BLI)
or also wake ingesting configuration, will lead to confusing and even wrong interpretations of the engine
performance. The deviation from this idealised behaviour is usually defined as the engine interference and
is included in the drag bookkeeping for the aircraft.
When the control volume, as described above, is shrunk to an extent where the static pressure cannot be
considered equal to the free stream pressure along its boundaries everywhere, then the pressure forces
have to be included in the thrust calculations. This may be the case when the engine is mounted in a
accelerated/decelerated flow field or when the engine is viewed as an infinitely thin disk alone. The latter
case, the actuator disk model deserves a separate explanation.
For an isolated propeller with no or negligible forward motion, J = 0, and constant cT , the Tc will approach
infinity, which results in a large axial inflow factor, a >> 1. When this is combined with x >> R,q which
means that the measurement is performed far away from the disk the contraction ratio is found to be 12 as
shown below.
r r
Rsl 1+a 1
= ≈ ≈ 0.7 (5.12)
R 1 + 2a 2
This requires a rapid contraction close behind the actuator disk. This approximation can be used to de-
termine the wing area wetted by the propeller slipstream when correcting for the “superlift” in wind tunnel
testing for propeller aircraft.
T = Sp (p2 − p1 ) (5.13)
Where Sp = π4 D2 is the cross-section of the actuator disk, p1 is the pressure upstream and p2 is the pressure
downstream of the actuator disk. Note that with this representation a positive thrust is against the direction
of the flow. For determining the thrust, no assumption about the distribution of the pressure jump across the
disk needs to be made. In the classical actuator disk model half of the pressure change occurs upstream
of the disk, also known as the suction side. The other half of the pressure jump occurs downstream of the
disk, or the pressure side. However, equally well, the same thrust can be ascribed to a suction side only, so
that behind the propeller the free stream static pressure is recovered immediately.
This approximation corresponds to the case where the pressure directly behind the actuator disk equals the
free stream pressure. Under such condition, neglecting the contraction of the stream tube downstream of
the actuator disk can be assumed a reasonable approximation. With these assumptions the jump conditions
at the location of the propeller could be applied.
p2 = p∞ (5.14)
Thus the pressure immediately ahead of the actuator disk can be formulated as below:
T
p1 = p∞ − (5.15)
Sp
When the incompressible Bernoulli equation is applied to the stream tube from far upstream to the plane
immediately in front of the actuator disk, we obtain :
1 1 T
p∞ + ρu2∞ = p∞ + ρu21 − (5.16)
2 2 Sp
or simply
1 2 1 T
ρu∞ = ρu21 − (5.17)
2 2 Sp
Further, assuming continuous incompressible flow up to the actuator disk, one can obtain the tube contrac-
tion ratio ahead of the propeller plane from the continuity equation.
S∞ u∞ = Sp u1
S∞ (5.18)
u1 = u∞
Sp
Inserting this result in equation 5.17 will result in:
2
1 2 1 S∞ T
ρu∞ = ρu2∞ − (5.19)
2 2 Sp Sp
From which the contraction ratio in terms of areas can be obtained.
2
S∞ T
=1+ (5.20)
Sp Sp ρu2∞
It should be noted that the last term represents the thrust coefficient Tc , thus:
r r
S∞ 8 8 CT
= 1 + Tc = 1 + (5.22)
Sp π π J2
When comparing the above equations one can see that the area ratio is obviously related to the axial inflow
factor discussed earlier in this chapter in equation 5.8. For small advance ratios, the contraction ahead
of the propeller plane becomes large. In the case of u∞ = 0, S∞ → ∞. This trend is in agreement with
empirical observations. The limiting value for the contraction ratio could be obtained from the observation
that there is a limit to the thrust by achievable aerodynamic performance of the propeller blades. Equations
5.21 and 5.22 could be used for a scaling exercise for the elevation of the propeller above ground to such a
position that the outer limit of the stream tube just reaches the ground at ∞ upstream.
r r
h∞ 4 8 4 8 CT
= 1 + Tc = 1 + (5.23)
rp π π J2
u0∞ 2Tc α1
=1− q (5.24)
u∞ π 1 + π8 Tc
Sp
Where α is the ratio of the of the disk area to the tunnel cross section area, α = St . A more extensive
discussion of this is included in the chapter about wind tunnels and wall corrections.
The actuator disk approach above is kept generic enough to be applicable to compressible and incompress-
ible flow situations. In the section dealing with the particular testing technique of turbofan engine effects,
the discussion will be refined to account for compressibility effects as well.
A number of the interference effects also remain present when the aft fuselage mounted configuration is
chosen. Additionally the slipstream interaction with the tail-plane and control surfaces will then need to be
considered.
One way of splitting up the aerodynamic interference effects is looking at the geometry effects and the
power setting effects separately. Both of them affect the flow around the aircraft, and may amplify or
dampen each other. But they are easier to understand, when viewed separately. As geometry effects, the
channel geometry, the surface area and the cross section area of the engine might be classified.
The channel flow effect is present in all versions of podded engine configurations; both on the wing as well
as on the fuselage. An example for the fuselage mounted configuration is shown in figure 5.3. Here the
‘surface streamline’ pattern of the pylon flow is made visible with the help of a fluorescent dye visualization,
where viscous fluid embedding fluorescent paricles is spread in a thin layer on the solid surface. The oil
spreads according to the wall shear stress vector field forming a so-called surface streamline pattern. In the
current example the channel flow is generated with the help of the simplified engine simulator, a through
flow nacelle (TFN). This particular image has been obtained in DNW-HST. Part of the pattern could be
interpreted as indication of a separation bubble in the conditions representing flight idle settings.
Figure 5.3: Surface flow visualisation in the channel between the engine cowling and the fuselage. Possi-
bility of a separation bubble on the pylon fuselage junction.(Figure from DNW)
The ‘surface area effect’ can be estimated by calculating the engine cowling and pylon surface areas, and
multiplying those with an appropriate local skin friction coefficient Cf . For a flat plate in turbulent flow this
(a) Engine simulators at DNW. (b) Effect of the bypass ratio on the en-
gine/wing interaction in case of constant verti-
cal distance between the wing and engine axis.
Figure 5.4: Engine simulators at DNW and their wing interference. (Figure from DNW)
The second, also fundamental power effect is the loss of lift as an interaction between the wing-mounted
engine and the wing circulation. This is shown in figure 5.5. In this figure the different configurations of
engines (or simulators of the engines) are shown with their effect upon the vertical wing loading, i.e. the lift
and compared with the clean wing, which is the dashed black line.
A special case of interference occurs in the case of wing mounted propeller engines in a tractor or puller
configuration. In that case part of the wing is submerged in the propeller slipstream, figure 5.6, experiencing
flow with a total pressure higher than that of the free stream. This exposure to the increased total and
therefore also the increased dynamic pressure affects the behaviour of the propeller aircraft and needs to
be accounted for in the wind tunnel testing. As the propeller is mounted upstream of the wing, the wing will
experience accelerated flow in the slipstream of the propeller. This causes the lift to increase locally. This
can both be beneficial or disadvantageous depending on the application.
The propeller does not only cause the flow in the slipstream to locally accelerate, it also introduces swirl
to the flow which affects the wing lift distribution. The general rule of thumb is that the downward stroke
reduces the wing lift in the slipstream and the upward stroke increases it. The two propeller interference
effects are shown in figure 5.7. For the correction of lift values, it is usually sufficient to account for the
increased dynamic pressure in the slipstream only. The swirl effects on the downstream and upstream
stroke are approximately equal (when the wing is operating in the linear range of the lift curve) and can be
neglected in the accounting for required corrections.
Figure 5.8 gives a general estimation of the ’degree of difficulties’ when choosing the position oft the engine
relative to the wing. Here the difficulty is mainly determined by the desire to avoid detrimental interaction
between the airframe and the engine.
All the interactions in their sum can be counterproductive to its original purpose of increasing propulsive
efficiency by increasing the bypass ratio. The general rule is that higher the bypass ratio, the higher the effi-
ciency of the isolated engine. The installation interference effects can cancel this benefit. The contradicted
(a) Engine mounting parameters. (b) Plot showing the challenge of certain engine configura-
tions.
tendencies are illustrated in figure 5.9 in a qualitative manner. In this graph, the qualitative amount of fuel
consumption of an engine is plotted against its bypass ratio, and for an isolated engine this is the monoton-
ically falling curve. At the same time, the installation drag increases at such a rate that at a certain point
the total fuel consumption can start increasing again. This consideration is valid not only when designing
a new aircraft where detailed measures can be applied to remedy the situation, but also when re-engining
an existing aircraft, when in general fewer design degrees of freedom are present. In both cases, a careful
evaluation of the quantitative amount of interference can be obtained from suitably designed experiments.
Overall, the increase in efficiency of propulsion achievable through increasing the mass flow through the
fan and reducing the pressure jump across it can become counterproductive when the interactions are not
properly dealt with, as indicated in figure 5.9. Note that the graph is only qualitative since manufacturers
jealously guard their data.
Note: Flow non-uniformities on the entrance plane affecting the performance of the engines are usually not
considered in civil aircraft applications. It is assumed that the engines are placed in undisturbed flow. This
is not the case in military aircraft engine integration, where a lot of effort is dedicated to reducing the losses
in the inlet ducts.
Where f f stand for the free flight case and wt stands for the wind tunnel model. The thrust coefficient of
the engine simulator will also have to be equal to that of the real motor, thus:
T
Tc = 1 2
(5.27)
2 ρu Sp
It can be seen that the thrust coefficient is computed in the same manner as the lift and drag coefficient.
with the exception of the reference surface. When computing thrust for isolated propulsors the most used
reference area is the cross-sectional area of the propulsors as shown above. Though airframe designers
quite often choose to reference the thrust to the wing surface area as well, for convenience sake. It is
important to remember this definition when different sizes or types of a propulsor are compared with each
other. It is clear from the definition of the thrust coefficient, that the thrust requirement for the simulators
scales linearly with density and Sp and quadratically with velocity. From this follows that, when assuming
that the testing is performed at the same velocity and the same atmospheric pressure, then the power
requirement scales with the square of the size of the model, Dp .
3 2
ρwt uwt Dwt
Pwt = Pf f (5.28)
ρf f uf f Df f
This expression will have a slightly modified form, when we discuss the open rotors in a more specific
manner. While the requirement for power decreases with the area of the propulsor, the power density of
Here the powered model data has to be ‘cleaned’ of the thrust values themselves in terms of the thrust
book-keeping, as will be discussed later.
2. Power effect, where the power setting dependency of interference is evaluated.
In the power effect testing, the question of behaviour of the aircraft at different power setting is de-
scribed, without a direct look at the detail of the interactions. The installation effect is already account-
ed for.
3. Jet interference effect.
This deals mainly with the effect of the slipstream on the control surfaces in flight conditions. it is not
dissimilar to the previous category, except that a more detailed analysis of the interaction is aimed at.
The categories are arbitrary, any other categorisation that is well structured, can be introduced.
5.4 Simulators
The simplest form of engine simulation in the wind tunnel is a through flow nacelle (TFN); an example has
already been shown in figure 5.3. This is an aerodynamic shape corresponding to the outer shell of the
engine cowling and providing acceleration inside the hollow inner contour. The use of a TFN corresponds
approximately to the simulation of flight with engines set on idle. In order to be able to simulate the effects
of different power settings, powered engine simulators are required. The most advanced of those are the
air turbine powered simulators (TPS), shown in figure 5.10. In this figure the shaded areas represent the air
Figure 5.10: Turbine powered simulator with the performance monitoring planes identified.
(a) Engine: all air enters through the fan-intake. (b) TPS: turbine air enters through pipe-line.
Figure 5.11: Differences in air mass flow through turbine and TPS (M.Laban et al.).
Other effects to be considered are the shaping of the spinner and the outer contour. The outer contour
performance is affected by the character of the boundary layer developing on the nacelle and the propensity
of the lower Reynolds number boundary layer in the wind tunnel to separate earlier at high angles of attack.
The same AIAA 2005-3703 paper [10] describes in some detail the methodologies of reshaping engine
nacelles for TPS testing in wind tunnels.
In the case of simulating propulsion by propeller engines, actual scaled propellers are used in wind tunnels.
These can be driven either by electric, hydraulic or, in a manner similar to the TPS, by air motors. The
drawback of the current electric motors is their lower power density compared to pneumatic and hydraulic
motors. This often leads to larger nacelle diameters for electrically powered propulsors, Furthermore, the
larger heat generation complexifies measurements with internal load cells, and the high current supplied
to the motor generates an unfavorable testing environment in terms of electromagnetic interference. On
the other hand, compared to the traditional pneumatic and hydraulic motors, the use of electric motors has
advantages in terms of flexibility and simplicity of installation and operation of the engines, and improved
quality of the control of the operational conditions of the engine (like rotational speed, phase angle).
Gross thrust: FG
Measured
Figure 5.12: Force balance for determining the interference drag. The calculation is based on ideal expan-
sion and calibrated correction values.
The only directly measured parameter is the total force (in the direction of the thrust, Fx ) measured by
the balance. For determining the interference, the drag of the aircraft with installed power units has to
be measured and compared with that of the clean configuration. For that the thrust produced by the air
turbine and the thrust produced by the fan have to be determined. The drag of the TPS inlet has to be
subtracted from the gross thrust to yield the net thrust. The difference between the net thrust and the
balance measurement is the drag including the interference effects.
The parameters ram drag (also known as inlet momentum), fan thrust and turbine thrust are determined
indirectly with the help of measured thermodynamic parameters pressure and temperature in addition to the
calibration results from the engine calibration facility. For the discussion of the thermodynamic parameters
and their use, the definition of different reference planes is given in the figures 5.13 and 5.14. The most
important reference planes are the upstream far field entrance plane 0 and the fan exit plane 19. The turbine
exit plane is denoted as the plane 9.
In the following, the procedure of the determination of single components of force bookkeeping as was
illustrated in figure 5.12 is described in some more detail with the help of figure 5.14. With the given free
stream conditions in the upstream reference plane the outflow conditions have to be determined from the
measured parameters. Instead of the more global evaluation of the thrust generation in the beginning of
this chapter, here also the compressibility effects will be accounted for. The basic accounting uses the
isentropic relationships.
T 1
= γ−1 (5.33)
T0
1 + 2 M2
p 1
= γ (5.34)
p0 1+ γ−1 2
γ−1
2 M
The deviations from the isentropic expansions will be accounted for in the calibration process.
p
u19 = M19 a19 = M19 γRT19 (5.35)
From the isentropic expansion relationship for pressure above, the Mach number for the fan exit plane can
be expressed as:
v " γ−1 #
u
u 2 p0 γ
M19 = t −1 (5.36)
γ−1 p
With this the fan velocity can be expressed as a function of free stream paprameters and parameters
obtained in the slipstream measurements:
v
u
u 2γRT0
" γ−1 #
p19 γ
u19 (p, T0 , p0 ) = t 19
1− (5.38)
γ−1 p019
It is assumed that the flow is perfectly expanded, such that p19 corresponds to the free stream static pressure
p∞ , and the total pressure and temperature are conserved from the measurement station 15 to the exit plane
19 of the control volume, p015 = p019 and T015 = T019 .
In the case that the jet in the fan stream expands into a flow field with a static pressure different from the
free stream static pressure, extra correction for the pressure difference is required. This could be the case
when the fan exhaust is under the wing and the pressure side has a non-zero pressure coefficient Cp . This
would result in a different mass flow through the channel.
In the above expressions no allowance is made for the variation of the ratio of specific heats in the flow field.
In the practice, for the benefit of accuracy, those values are also determined according to local conditions
and include humidity in the ambient air.
For the pressure this is only satisfied when local supersonic conditions are not reached anywhere in the
expansion, which would have to be decelerated by shocks. That would be the case if p∞15 ≥ 1.89p∞19 , but
is not generally encountered in our simulations of engines with fan pressure ratios (FPR) of 1.3 to 1.5.
Even with the lack of shock induced losses, the ideal isentropic expansion results have to be corrected for
‘imperfections’ by calibration, as will be described later.
The fan mass flow is a product of the velocity with the density and the cross sectional area of the stream
tube:
In this case A19 is determined by the geometry and ρ19 needs to be determined at station 15 by using
p
ρ = RT . Since the total temperature is conserved, the temperature at the location 19 is determined by the
isentropic expansion. The area A19 is taken to be equal to the area at the fan exit plane. Maintaining this
approach consistently in the calibration as well as in the test evaluation assures that any error introduced
by this interpretation is absorbed in the calibrated values.
It is assumed that p8 = p9 = p∞ . For the turbine mass flow, there is an additional option available for
determining it from the supply line evaluation instead of measurements at the location 15. Usually sonic
venturis are used to determine the mass flow. These are chocked devices that determine the maximum
mass flow through a contraction.
5.7 Calibration
In order to account for any imperfections and deviations from the isentropic, inviscid expansion relation-
ships, a calibration procedure has to be applied. In order to obtain an independent measurement for those
parameters that have to be determined in the wind tunnel test in an indirect way, a calibration facility and a
corresponding procedure is required. A schematic representation of the DNW Engine Simulator Calibration
facility is given in figure 5.15. It is essentially a vacuum tank with controlled through-flow. The inlet of this
tank is formed by the TPS or components thereof; the outlet is formed by a selection of sonic venturi noz-
zles. These allow a precise determination of the mass flow, required for the thrust calculations. The precise
mass flux determination is also required in order to calibrate the calculations of the ideal flow through the
TPS to the real values.
Figure 5.15: Calibration facility featuring a bellmouth intake, a TPS mounting mechanism equipped with
load cells and a set of venturi nozzles for mass flux determination.
In order to obtain controlled mass flow through the fan, a bellmouth shape contraction is attached to the
inlet plane of the TPS in order to avoid possible separation effects caused by the suction around the intake
lip.
As mentioned, the evaluation of test results relies on calibration of the TPS-s, because the drag, as well
as the lift and other forces and moments, are measured by one central balance integrated into the airplane
fuselage. This is the case even in half-model testing with a balance external to the wind tunnel as it is still
integrated with the body of the aircraft. In the calibration procedure the net thrust has to be subtracted from
the measured drag and comparing that result with the drag of a clean configuration will yield the installation
drag, as discussed earlier.
ṁ19 ṁBM
Cd = = (5.41)
ṁ19isentropic ṁ19insentropic
ṁtankventuri − ṁBM
CdBM = (5.42)
ṁBMinsentropic
Where
In a similar approach, in order to correct for the fan exit velocity inaccuracy obtained from using the isentropic
approximation, needed to determine the gross thrust of the fan, a velocity coefficient has to be determined
from the calibration procedure. The required calibration for the fan velocity coefficient is defined in an
analogous way:
u19 ṁ19 u19
Cu19 = = (5.45)
u19insentropic ṁ19 u19insentropic
The numerator on the right, which is the term for the fan thrust, can be obtained from the balance measure-
ment in combination with the known turbine conditions:
Fbalance − Fturbine Fbalance − ṁ9 u9insentropic
Cu19 = = (5.46)
ṁ19 u19insentropic ṁ19 u19insentropic
The discharge coefficients typically have values between 0.9 and 0.95 accounting for the losses in the
boundary layers and elsewhere.
Stapes
attached to
val window)
Semicircular
Canals
ncus
Vestibular
Nerve
Cochlear
Nerve
Cochlea
External
Auditory Canal Ty
Ca
When discussing sound the frequency domain is often used. Depending on the age and hearing quality of
a person the human hearing system can detect pressure fluctuations between 20 Hz-20 kHz, although the
upper limit decreases with age. The sensitivity of the frequency is not uniform as seen from figure 6.2. In
general it is most sensitive for frequencies around 3-4 kHz, which is also approximately the frequency of
the human voice.
As seen from figure 6.2 the sensitivity of the human hearing system is expressed in dB; the dB is the unit
(actually, it is not a unit at all as shown below, although it is usually used as such) generally used to express
the Sound Pressure Level (SPL), which is one of the methods to describe the ‘strength’ of the sound. The
other measure is the Sound Power Level (SWL). Both the Sound Pressure and Sound Power levels can
provide useful information about sound but what they represent is different. the Sound Power Level is used
to describe the strength of the sound source, whilst Sound Pressure Level describes the effect it has on its
surroundings, as illustrated in figure 6.3. Equation 6.1 shows how the SWL is computed. It can be seen that
the sound power level, LW , is logarithmically dependent on a power ratio of which P0 = 1pW = 1 · 10−12 W
is the reference power and P is the source power.
P
LW = 10 · log10 [dB] (6.1)
P0
The sound pressure level, Lp , is computed similarly to the sound power level, however in this case a
pressure ratio, rather than a power ratio is used.
prms
Lp = 20 · log10 [dB] (6.2)
p0
In the above equation p is the root mean square sound pressure and p0 = 20µP a is the reference pressure.
As both SWL and SPL are logarithmic functions it can be deduced that the sum of two equal noise sources
results in an increase of noise strength by 3dB independently of their individual strength; so 80dB + 80dB =
83dB as is 40dB + 40dB = 43dB .
To give a better understanding as to how sound scales typical sound pressure and power levels that relate
to well-known sound sources are shown in figure 6.4.
∂2p 1 ∂2p
− =0 (6.3)
∂x2 c2 ∂t2
180 dB 180 dB
Jet airliner (500, 000W )
dB
dB
(relative to 20µP a) 100 dB (relative to 10pW )
100 dB
80 dB Truck at 20 m 80 dB
Small office machine (20 · 10−6 W )
60 dB 60 dB
Conversation at 1 m
40 dB 40 dB
20 dB 20 dB
0 dB Hearing threshold 0 dB
(a) Typical sound pressure levels, Lp . (b) Typical sound power levels, LW .
The above form is valid when there is no ambient flow as in that case the pressure, p, is only a function
of x and t. c in this case is the speed of sound, which can be expressed as c2 = γRT . A solution to the
differential equation shown above is of the following form:
p = f1 (ct − x) + f2 (ct + x) (6.4)
This solution describes characteristics moving left and right with the speed of sound. Note that c describes
the phase velocity and not a transport velocity. In the mean, no material is being transported. When
vibrating sources are used rather than impulses the above solution takes another form, which is shown
below.
p = A1 sin(ωt − kx + φ1 ) + A2 sin(ωt + kx + φ2 ) (6.5)
ω 2π
Where ω is the frequency of the harmonic oscillation, and k = c = λ is the wave number and λ is the
wavelength.
The sound pressure is a simple harmonic in time oscillating around a steady value. It is the linear superpo-
sition of two waves travelling in opposite directions at a constant speed. In order to find this sound pressure
the characteristic impedance of a medium is required, which is computed as: R = ρc = up . This means that
kg kg
for air, with c = 340 m
s and ρ = 1.2 m3 , the characteristic impedance is R = 415 m2 s
The intensity is defined as time averaged power transported across a surface. Power is the product of force
and velocity, P = F · v = p · A · v. Taking the time averaged value of p results in prms and normalising it with
p2rms
I= (6.6)
R
In the case that the signal is a sine wave the above equation can be written in terms of the amplitude of the
pressure signal as:
p2
I= (6.7)
2R
Sound power is the surface integral over a closed surface including the source within it whilst the radiat-
ed power is a function of its boundary conditions. A solid wall doubles the power of a dipole, a corner
quadruples it.
The sound power and sound pressure levels are expressed as shown in equations 6.1 and 6.2 respectively.
W
When it comes to sound power the power can also be interchanged with sound intensity in m2.
In the case for compact subsonic jets the characteristic wavelength is of the order of λ = cDU , where D is
the jet diameter and U the flow velocity. The acoustic power scales with the 8th power of the flow velocity:
hp2 i
∝ M8 (6.9)
(ρ0 c2 )2
Since free turbulence is a less efficient sound generator than the wall bounded turbulence, the amplitude
decay with the distance squared, rD2 , needs to be considered for measurement purposes. This is due to the
fact that the power needs to be distributed over an increasing area, which scales with r2 .
Another case is sound scattering from an edge where the solid body ends. In this case the acoustic power
depends on the direction as seen in the equation below.
hp2 i
5 2 Θ
∝ M · sin (6.10)
(ρ0 c2 )2 2
When looking at the Mach dependence of the acoustic power one can see that the above case is the most
efficient sound generator in the flow field as it scales with M 5 rather than the 6th or 8th power.
When measuring sound in aerodynamic situations, it has to be remembered, that sound waves can be
convected with the flow. In the measurements this manifests itself as a Doppler effect. The Doppler effect
describes the perceived change in frequency or wavelength. When a wave from a single source carried
with velocity w with its emitted frequency ω, or corresponding period T = 2π
ω , then the front end of the wave
travels a distance cT during that period, whereas the tail end has already travelled the distance wT in that
time. The perceived wavelength is then shortened by wT . The Doppler-shifted frequency is then:
ω
ωshifted = (6.11)
1 − wc
Maximum sound pressure level SP Lmax ≥ 130dB due to boundary layer noise
Low self-noise level Lnoise < 20dB for high frequency signals
High sensitivity S > 10 mV
Pa sufficient signal to noise ratio
Broad frequency range frange ≤ 100Hz and ≥ 40kHz for scale models
shown in table 6.1. In most cases B&K 1/2”, 1/4” or 1/8” are used, which are very expensive. When including
the cable and the pre-amplifier the cost of the set is about e2000. Due to the cost of these microphones
these are only used for very accurate measurements where fewer microphones in an array are used, for
large arrays cheaper microphones, such as the WM-61 B102A electret microphone seen in table 6.2, are
usually opted for. These microphones are commonly used in arrays with ∼ 500 installed microphones.
SP Lmax ≥ 120dB
Lnoise 20dB
45 mV dBV
S 94 dB or −27 P a
frange 20Hz ≤ frange ≤ 40kHz
e ∼3
When acquiring noise data the acquisition parameters shown in table 6.3 are typical for noise measure-
ments. This results to ∼ 1Gb raw time data for 1 data point (dpn).
Sound levels are quantified with the help of frequency spectra, power density as a function of the frequency,
similar to the transfer function in the control theory. The difference being that SPL’s are integrated over
octaves in order to get a result. Octave filters are bandpass filters, with a bandwidth of 71% of the center
frequency.
!x ξ!
2
Figure 6.5: Microphone array and scan plane with monopole source.
noise experienced are in different phases. In the delay and sum algorithm these signals are corrected such
that identical phases are obtained and these corrected signals are summed. This procedure of delaying
and summing the signals is done for every scan point and the source location is found by having the highest
sum as this must be the shortest distance to the noise source. A graphical representation how this is done
is shown in figure 6.6.
In order to compute the array output as a function of the source position several steps need to be taken.
First of all Green’s function, which is an expression describing the response to an acoustic monopole, needs
to be determined.
2πif
− c ||~ ~
x−ξ||
~ =e
G(~x, ξ) (6.12)
~
4π||~x − ξ||
" !
! "
Difference in
amplitude
and phase
" !
! "
Microphones
Figure 6.6: Microphone array and scan plane with monopole source.
When Green’s function is determined the array output can be found as shown below.
N
~ = 1 X pn
a(ξ)
~
N n=1 G(~x2 , ξ)
N (6.13)
1 X h 2πif ||~x−ξ||
~
i
~
= pn e c 4π||~x − ξ||
N n=1 | {z | {z }
} Amplitude correction
Phase correction
In the above equation pn is the pressure of the nth microphone in the array
pN (f )
~
G(~x1 , ξ1 )
..
g= are required and the minimisation problem is posed: J = ||p − ag||2 The solution of this
.
G(~xN , ξ~N )
minimisation problem is shown below.
~ = g? p
a(ξ)
||g||2
PN
gn? pn
= Pn=1
N (6.14)
n=1 |gn |2
N 2πif ~
1 X e− c ||~x−ξ||
= p n
||g||2 n=1 ~
4π||~x − ξ||
Comparing equations 6.13 and 6.14 one can see that for the array output for the delay and sum algorithm
~ term, whilst for the conventional beam forming case it is inversely
is linearly dependent on the 4π||~x − ξ||
As a rule of thumb:
425 ds
resolution = · (6.15)
f DA
Where f is the source frequency, ds is the distance from the source to the array and DA is the array
diameter.
When first designing microphone arrays for the mapping of sound the microphones were placed randomly
as one thought this would prevent aliasing of the signal and result in better array gains. This theory was put
to the test at DNW where two different arrays were tested, one being completely random whilst the other
array still had some order in the placement of microphones, see figure 6.8. Both arrays consisted of 50
microphones and were fitted to an array of 2x2 m. In order to test the performance of these arrays a noise
source with a frequency of 2000 Hz was located 6 m from the centre of the array (0, 0, 6). In the perfect
case only one peak at should be visualised in the acoustic image located at the centre of the array, however
when looking at figure 6.9 one can see this is not the case. Especially the acoustic image of array 1 does
not perform that well. One can see that apart from the main lobe, located at (0, 0) there are several side
lobes. These side lobes are not physical and are due to the aliasing of the noise. From this figure it can
be seen that random arrays do not necessarily outperform arrays with some order as the acoustic image of
array 2 performs a lot better as the array gain has be increased from 8.5 dB to 12.5 dB.
Figure 6.8: Two different microphone arrays with varying randomness of microphone placement.
(a) Acoustic image of random microphone array 1. (b) Acoustic image of random microphone array 2.
Nowadays the arrays used at DNW are mostly ordered and the microphones are distributed as seen in
figure 6.10. Note that when using an ordered microphone distribution that different length scales between
microphones should be taken into account as without this aliasing can occur. When looking at the radial
distribution used by DNW one can see that the microphones are placed in circles with a higher microphone
density at the center. This allows for a high frequency range and a dynamic range of 12 − 13 dB
of the different setups are discussed as well as how to tackle some of the problems when testing in these
conditions.
Open jet noise measurements Open jet wind tunnels are favourable for noise testing wind tunnel models
as the testing hall around the Open Jet provides a semi anechoic environment that results in better quality
of the measurements. Having enough room in the hall also allows for far-field microphone measurements
to be used in parallel with the microphone array, however there are some drawbacks too. One of the most
problematic drawbacks of using open jets for noise measurements is the fact that the shear layer between
the hall and the jet causes serious disturbances. Not only does it refract noise as a lens does, see figure
6.11, it also causes a strong loss of coherence for frequencies > 20 kHz. This coherence loss is dependent
on several factors such as the frequency of the noise, the flow speed and the distance from the nozzle.
To correct for the refraction of the noise as it passes through the shear layer an Amiet correction needs to
be performed.
The shear layer is not the only factor that makes noise testing hard. In most cases for open jet facilities
the arrays are placed further away from the model with respect to closed test sections, which limits the
resolution for frequencies > 15 kHz.
The last problem with testing in open jet sections is the lesser aerodynamic quality when comparing tests
in closed sections. Thus it is harder to obtain the correct flight condition when testing aircraft link the noise
measurements to the final flight condition.
Closed test section noise measurements One of the advantages of doing noise testing in closed sec-
tions are that significant data up to ≈ 50 kHz can be acquired without too much loss of coherence for these
high frequencies. This is mainly due to the fact that the microphone arrays can be placed closer to the
model and that refraction of noise through the shear waves becomes less of an issue. Also, by placing
the array closer to the model higher resolutions can be achieved and noise sources can be determined
more accurately. Furthermore, there is no loss of aerodynamic simulation quality, which means that the
noise measurements and balance measurements can be performed simultaneously, which is cheaper than
performing two tests.
Wind
tunnel
Shear layer
Distorted
signals
Array
However, there are some drawbacks to this type of noise testing too. Since testing in a closed section
typically means testing in a hard walled environment issues like standing waves, acoustic reflections are a
real issue and have to be accounted for. Not only that, but the noise generated by the turbulent boundary
layer can be a real nuisance. This is due to the fact that microphones measure pressure differences and the
highly turbulent flow continually causes pressure differences that are recorded by the microphones as seen
in figure 6.12. Since the arrays have to be placed in the closed section they are often put in the walls of the
test section. Having these arrays inside the test section means that the echoing of the noise has to be dealt
with. And even though a closed jet wind tunnel does not have a shear layer such as the open jet, it does
cope with other problems causing loss of coherence. The main issue with this kind of noise testing is the
noise generated with the turbulent boundary layer. This boundary layer noise can be seen as generating
Flow
Array
random white-noise that distorts the real signal. When looking at only one microphone one could say that
the measured noise of the nth microphone is the sum of the signal noise and the boundary layer noise as
shown below.
pn = qn + εn (6.16)
|{z} |{z} |{z}
measured noise signal noise boundary layer noise
One of the solutions to tackle this problem is to exclude the diagonal of the main cross-power matrix as this
provides clean source plots with improved resolution. This effect can be seen in figure 6.13. One drawback
of removing the diagonal of the main cross-power matrix is that it may lead to non-physical negative ‘source
powers’ in the acoustic source plots.
(b) Noise map including diagonal terms. (c) Noise map without diagonal terms.
Figure 6.13: Acoustic maps of the F-100 model with and without boundary layer noise.
Where χ(tn ) is the microphone signal received at time tn , σ(τe ) is the source signal emitted at time τe and
F (τe ) is the transfer function for a moving monopole that relates the two signals. Using one microphone the
source signal can be determined if the transfer function is known as shown below:
χn (tn )
σ(τe ) = ≡ σ(τe ) (6.19)
F (τe )
When an array with a multitude of microphones is used the delay and sum algorithm can be used to solve
for the source signals as:
N N
1 X 1 X χn (tn )
σ(τe ) = σn (τe ) = ≡ σ(τe ) (6.20)
N n=1 N n=1 F (τe )
In figure 6.14 it can be seen that using this moving source processing technique delivers much better results
than ignoring this phenomenon.
(a) Conventional beam forming on moving source. (b) Moving source processing.
Another good example on moving source processing is measuring the noise of a helicopter. Tests by DNW
have shown that both the frequency of the rotors as well as the flight condition affect the measured noise.
When looking at figure 6.15 one can see that increasing the rotor frequency increases the rotor noise as
higher frequencies lead to higher tip velocities of the main rotor that at some point will become supersonic
causing loud shocks to occur. This is especially visible for the 4kHz case where the shock location can be
deduced by the relatively high noise region. As the frequency increases the relative noise of the tail rotor
goes down as the main rotor becomes the more dominant noise source. Not only does the frequency of the
rotor have an effect on the noise characteristics, the flight condition proves to be an important factor as well.
It can be seen that when descending the main rotor contributes the most noise. This is due to the fact that
Figure 6.15: Most dominant sound sources for different frequencies and flight scenarios.
the vortices created by the rotor travel downwards and when descending the main rotor interferes with the
previously generated vortices. This is also why relatively no noise is generated by the rotor during climb.
In figures 6.14 and 6.15 the noise measurements for the moving turbine and helicopter are still performed
inside a wind tunnel, however many moving source measurements are performed outside. One of these
tests are fly-over tests in which an aircraft flies over a microphone array in order to find out what aircraft
components cause the most amount of noise. Some typical results are shown in figure 6.16.
Nose gear
Engine inlet Slats
(a) Measure aircraft weight 2 times. (b) Measure helicopter weight 2 times.
Designed experiment: When one carefully designs an experiment different less measurements are re-
quired with often the same amount of precision. In this example the solution would be to measure the
added weight of the models and the difference, see figure 7.2, as this way the weight of the models is still
measured twice, but omits the need for two more tests saving money and time. Due to the fact that each
model is still weighed twice the precision of √σ2 remains unchanged, with the proof being: Assume S is the
sum of the aircraft and helicopter models and D is the difference
S +D S −D
S =A+H
→A= & H= (7.1)
D =A−H 2 2
It can be seen from above that A averages S + D and H averages S − D so in both cases n = 2.
From this example it can clearly be seen that designing experiments can reduce the costs and testing time
involved in testing.
EA1 = 50 − 30 = 20
EA2 = 62 − 40 = 22 (7.2)
50 + 62 30 + 40
EA = − = 21
2 2
EB1 = 40 − 30 = 10
EB2 = 62 − 50 = 12 (7.3)
40 + 62 30 + 50
EB = − = 11
2 2
Factor B
Low - Low -
30 50 30,32 50,54
- + - +
Low High Low High
Factor A Factor A
(a) Two-level factorial. (b) One factor at a time.
It can be seen that with only four data points the required precision is found. However, it can be seen from
the figure that one measurement takes place when the other factor has been varied as well and it is likely
that this has an effect on the main effect. In other words, there is an interference effect of factor B on factor
A. In this example the effect can be computed as follows:
EA2 − EA1 22 − 20
EAB = = =1 (7.4)
2 2
OFAT: When the same precision is wanted with OFAT all points need to be measured twice, which leads to
six data points.
EA1 = 50 − 30 = 20
EA2 = 54 − 32 = 22 (7.5)
54 − 32 50 − 30
EA = − = 21
2 2
EB1 = 38 − 30 = 8
EB2 = 42 − 32 = 10 (7.6)
38 − 30 30 + 50
EB = − =9
2 2
Since there is no interaction between factors A and B there are no interaction effects to be deduced when
using OFAT.
From the above example it can be seen that the main and interaction effects take the form of:
±y1 ± y2 ± · · · ± yn
ζ=
n/2
(7.7)
2 2 2
=± y1 ± y2 ± · · · ± yn
n n n
The partial derivative term from the above equation can also be found for the main and interaction effects
as shown below.
2
∂ζ n 2 4
= ± = 2 (7.9)
∂yi 2 n
Since all effects y1 . . . yn takes the same form it can be seen that all standard deviations are the same so
σyi = σy . substituting this into the variance formula the following is found.
4 4 4
σζ2 = σ 2
y + σ 2
y + · · · + σy2
n2 n2 n2
4
=n σy2 (7.10)
n2
4
= σy2
n
Thus the variance formula can be finally written as:
2
σζ = √ σy (7.11)
n
Note that the same variance formula applies to all factorial effects, including main effects and interactions.
Run Variable
Number A B C D E F ABCDEF Block
1 - - - - - - + 1
2 - - - - - + - 2
3 - - - - + - - 2
4 - - - - + + + 1
5 - - - + - - - 2
6 - - - + - + + 1
7 - - + + - + 1
8 - - - + + + - 2
.. .. .. .. .. .. .. .. ..
. . . . . . . . .
63 + + + + + - - 2
64 + + + + + + + 1
F = + − + + −···+ (7.12)
(7.13)
F · F = F2 = I = + + + + +···+ (7.14)
It can be seen in table 7.2 the ABCDEF column is such a unit vector, thus:
I = ABCDEF (7.15)
F =F ·I
= F · ABCDEF
= ABCDEF 2 (7.16)
= ABCDEI
= ABCDE
The duplicate columns can be explained since F = ABCDE as proven above. This shows that the amount
of effects are reduced from 2k − 1 = 63 to 2k−1 − 1 = 31. Other aliases are found in a similar fashion such
as AB · I = AB · ABCDEF → AB = A2 B 2 CDEF = I 2 CDEF = CDEF . In general, n-way interactions
are aliased with m-way interactions, where n + m = nvariables . So in the case of a 6 variable experiment the
main effects are aliased with the 5-way interactions, and the 2-way interactions with the 4-way interactions,
or, in extreme cases 3-way interactions will be aliased with 3-way interactions.
Thus it is not possible to find all 63 effect by only measuring 32 times. Though not all effects can be
determined the most dominant ones are, so a half factorial design might be of use to discern the 31 most
dominant effects. These dominant terms are shown in the Pareto chart in figure 7.5. From both the pareto
chart and the nominal plots from figure 7.6 one can see that the most dominant effects remain visible,
however the less dominant effects are filtered out when reducing the amount of measurements. When
using only 16 measurements, or a quarter factorial experiment, one can still see the most dominant effects,
such as the angle of attack, A, or the flap size, E.
Figure 7.6: Normal plots for full factorial, half factorial and quarter factorial designs.
In this representation the forces are represented by F and the moments by M . The measured strain gauge
voltage is denoted as U . This matrix representation corresponds to the schematic representation as shown
in figure 8.1.
The coefficients ki in the balance matrix are the balance coefficients and are obtained from a calibration
procedure. In this procedure, the loads are applied a single one at a time, and it is assumed that the
combined loads do not affect the coefficients in the application. When a more refined approach is opted for,
the combination loads can not be neglected. In this case the balance equation is including the quadratic
(interaction) terms as well:
R1
..
Fx
.
Fy
a11 a12 · · · a16 a111 · · · a166 R6
Fz . 2
= .
Mx . R1
(8.2)
R1 R2
a61 a62 · · · a66 a611 · · · a666
My
.
Mz ..
R62
In this quadratic equation the balance coefficients are denoted as aij and the balance signals as R. (This is
probably a historical notation because the Wheatstone-bridge output voltage U is a function of the change
of the resistance R of the sensitive element).
It is obvious, that the 2nd order calibration to obtain the balance coefficients is more time consuming than
the first order one. There are a number of recent developments represented in the literature about different
strategies of balance calibration in order to have an efficient procedure and still maintain a low uncertainty
level. The balance uncertainty is usually given by the expression:
6
K X Fj
UCALn = ± Fnmax an + bn (8.3)
1000 Fjmax
j=1
j 6= n
This expression is used to link the inaccuracy of the individual measured balance signal to its full scale
range.
How the balance uncertainty depends on the choice of the calibration procedure is discussed in the pro-
vided literature [11, 17–19]. However procedures as one factor at a time (OFAT) or a more statistics based
approach, such as modern design of experiment (MDOE), do have different effects on the uncertainty, see
figure 8.2.