You are on page 1of 41

Delft University of Technology

Department of Aerospace Engineering

Experimental simulation of the performance of a


propeller aircraft in wind tunnel
Submitted in partial fulfillment of the course requirements in

AE4115: Experimental Simulation

Submitted by
Sachin Menon (4743571)
V. S. K. Rajampeta (4738799)
Kunal Kanawade (4750721)
Yashasvi Giridhar (4749219)

July 2018
Contents
1 Introduction 1
1.1 Model description and experimental facility . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Tests performed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2 Wind Tunnel Corrections 5


2.1 Solid Model Blockage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1.1 Solid blockage effect: Wings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1.2 Solid blockage effect: Fuselage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2 Wake Blockage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3 Slipstream Interference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.4 Streamline Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3 Results and discussions 11


3.1 Configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.1.1 Effect of horizontal tail wing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.1.2 Effect of propeller . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.2 Blockage Correction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.3 Streamline Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.4 Slipstream Interference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

4 Noise 17
4.1 Generic Investigation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.2 WFHNV Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.3 Unpowered WFHNVP Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.4 Powered WFHNVP Configuration, Zero Angle of Attack . . . . . . . . . . . . . . . . . . . . . 20
4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

5 Scaling Effects 24
5.1 Similarity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
5.2 Boundary layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
5.3 Drag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.4 Lift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.5 Stability and control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

6 Conclusions 28

7 Recommendations 29

A WFHNVP Noise Measurements 30


A.1 n = 220rpm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
A.2 n = 240rpm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
A.3 n = 260rpm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
A.4 n = 280rpm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
A.5 n = 300rpm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
A.6 n = 320rpm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

i
List of Figures
1.1 Dimensions of the Fokker-27 model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Dimensions of cross section in mm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Schematic overview of wind tunnel . . . . . . !
. . . . . . . . . . . . . . . . . . . . . . . . . . . 3
t
2.1 Body shape factor (K) vs. thickness ratio c [1, p. 369] . . . . . . . . . . . . . . . . . . . 6

2.2 τ1 for various tunnel types [1, p. 369] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7


2.3 Maskell’s analysis of wake blockage due to separated flow . . . . . . . . . . . . . . . . . . 8
2.4 Wake blockage effect for WFNVH at V = 50ms−1 . . . . . . . . . . . . . . . . . . . . . . . . 8
3.1 Effect of horizontal tail on the performance curves. . . . . . . . . . . . . . . . . . . . . . . . 11
3.2 Effect of horizontal tail on pitching moment. . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.3 Effect of propeller on the performance curves. . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.4 Effect of propeller on pitching moment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.5 Blockage correction for CL and CD of WFNV configuration . . . . . . . . . . . . . . . . . . 13
3.6 Blockage correction CM for WFNV configuration . . . . . . . . . . . . . . . . . . . . . . . . 13
3.7 Blockage correction for CL and CD for WFNVH configuration . . . . . . . . . . . . . . . . . 13
3.8 Blockage correction for CM for WFNVH configuration . . . . . . . . . . . . . . . . . . . . . 14
3.9 Blockage correction for CL and CD for WFNVHP configuration . . . . . . . . . . . . . . . . . 14
3.10 Blockage correction for CM for WFNVHP configuration . . . . . . . . . . . . . . . . . . . . 14
3.11 Streamline curvature correction for CL and CD for WFNV configuration . . . . . . . . . . . 15
3.12 Streamline curvature correction for CL and CD for WFNVH configuration . . . . . . . . . . 15
3.13 Streamline curvature correction for CL and CD for WFNVHP configuration . . . . . . . . . 15
3.14 Thrust correction for CL and CD due to propeller slipstream interference . . . . . . . . . . . 16
3.15 Thrust Correction for Pitching moment due to propeller slipstream interference . . . . . . . 16
4.1 Microphone placement relative to the propeller. . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.2 SPL as a function of frequency with a powered propeller at α = 0 deg, n = 300rpm . . . 18
4.3 SPL as a function of frequency without an installed propeller at α = 1 deg . . . . . . . . . 18
4.4 Full frequency range of noise measurements in the WFHNV configuration at all angles of
attack α. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.5 Frequency subrange of noise measurements in the WFHNV configuration at all angles of
attack α. Note the color scheme for easy interpretation. . . . . . . . . . . . . . . . . . . . . 19
4.6 Frequency subrange of noise measurements in the WFHNVP configuration at all angles
of attack α. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.7 Magnitudes of integer multiples of the propeller rotational rate, represented as the colored
surface. The white surface represents the µ + 2.5σ values. Where µ is the moving average
and σ the moving standard deviation of the spectrum component magnitudes. . . . . . . . . 21
4.8 WFHNVP configuration, n = 260rpm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.9 WFHNVP configuration, n = 300rpm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.10 WFHNVP configuration, n = 320rpm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.11 Magnitude of the f = 4n components of the WFNVHP configuration for the three micro-
phones. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
5.1 CL2 vs CD for NACA 23012 for two Re [1] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.2 CL2 vs CD for NACA 23012 for two Re [1] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.3 CDT vs log Re[1] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.4 CL vs Re[1] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
A.1 WFHNVP configuration, α = 0 deg, n = 220rpm, microphone 1 . . . . . . . . . . . . . . . . 30
A.2 WFHNVP configuration, α = 0 deg, n = 220rpm, microphone 2 . . . . . . . . . . . . . . . . 30
A.3 WFHNVP configuration, α = 0 deg, n = 220rpm, microphone 3 . . . . . . . . . . . . . . . . 31
A.4 WFHNVP configuration, α = 0 deg, n = 240rpm, microphone 1 . . . . . . . . . . . . . . . . 31

ii
A.5 WFHNVP configuration, α = 0 deg, n = 240rpm, microphone 2 . . . . . . . . . . . . . . . . 31
A.6 WFHNVP configuration, α = 0 deg, n = 240rpm, microphone 3 . . . . . . . . . . . . . . . . 32
A.7 WFHNVP configuration, α = 0 deg, n = 260rpm, microphone 1 . . . . . . . . . . . . . . . . 32
A.8 WFHNVP configuration, α = 0 deg, n = 260rpm, microphone 2 . . . . . . . . . . . . . . . . 32
A.9 WFHNVP configuration, α = 0 deg, n = 260rpm, microphone 3 . . . . . . . . . . . . . . . . 33
A.10 WFHNVP configuration, α = 0 deg, n = 280rpm, microphone 1 . . . . . . . . . . . . . . . . 33
A.11 WFHNVP configuration, α = 0 deg, n = 280rpm, microphone 2 . . . . . . . . . . . . . . . . 33
A.12 WFHNVP configuration, α = 0 deg, n = 280rpm, microphone 3 . . . . . . . . . . . . . . . . 34
A.13 WFHNVP configuration, α = 0 deg, n = 300rpm, microphone 1 . . . . . . . . . . . . . . . . 34
A.14 WFHNVP configuration, α = 0 deg, n = 300rpm, microphone 2 . . . . . . . . . . . . . . . . 34
A.15 WFHNVP configuration, α = 0 deg, n = 300rpm, microphone 3 . . . . . . . . . . . . . . . . 35
A.16 WFHNVP configuration, α = 0 deg, n = 320rpm, microphone 1 . . . . . . . . . . . . . . . . 35
A.17 WFHNVP configuration, α = 0 deg, n = 320rpm, microphone 2 . . . . . . . . . . . . . . . . 35
A.18 WFHNVP configuration, α = 0 deg, n = 320rpm, microphone 3 . . . . . . . . . . . . . . . . 36

List of Tables
1 Specifications of the Fokker-27 model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2 Test matrix for Fokker F27 experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

iii
1 Introduction
This chapter will briefly discuss the model used for the experiment and the experimental facility where
the experiment was conducted.

1.1 Model description and experimental facility


The model used for the experiment is of Fokker-27. Dimensions of the model and the location support
for the model are shown in the following figure 1.1. The detailed specification of the model are given in
the table 1. An external balance, measuring the force and moment in x,y and z direction of wind tunnel,
mounts the model. Upside down mounting of the Fokker-27 model does not let the suction side of the
wing interact with the model strings. This helps the setup as the suction side is more sensitive than
pressure side.

Figure 1.1: Dimensions of the Fokker-27 model


The experiment has been conducted in the Low Turbulence Tunnel in Delft. This facility is a closed
circuit low speed wind tunnel. A 525 kW DC motor powers the six bladed fan which is used to accelerate
the flow in the facility. Maximum achievable velocity of the flow is 120 m/s and about Reynolds number
3.5 ×106 can be achieved. The dimensions of the cross section are shown in the figure. Owing to the
large contraction ratio the turbulence level varies from 0.015 % at 20 m/s to 0.07 % at 75 m/s.

1
Model Scale 1:20
Wing
Span 1.450 m
Area 0.175 m2
Mean Aerodynamic chord 0.1282 m
Taper 0.4225
Aspect Ratio 12
Dihedral (@ 0.4c) inner wing 0
(@ 0.4c) outer wing 0◦
Sweep angle (@ 0.25c) 2.5◦
Incidence with respect to fuselage center line 3.46◦
Twist inner wing 0◦
Twist outer wing -2◦
Airfoil (root) NACA 642-421 (adapted)
Airfoil (tip) NACA 642-421 (adapted)
Position of wing pivot point (with reference to the
X0 = 34.5 % MAC Z0 = 60.8 % MAC
leading edge of the MAC; Xm -axis)
Fuselage
Length 1.155 m
Height 0.1385 m
Width 0.135 m
Fuselage attachment points (w.r.t.the leading edge
Front : 0.0442 m Back : 0.4064 m
of the MAC)
Horizontal tail
Span 0.490 m
Area 0.0402 m2
Mean Aerodynamic chord 0.0877 m
Aspect Ratio 5.95
Nominal incidence angle 0◦
Tail length 0.534 m
Propeller
Diameter 0.183 m
Number of blades 4
Engines
Manufacturer TASK Corporation
Type 3-phase induction
Power 3.6 kW (each)
Maximum rpm 30 000 rpm
Number 2

Table 1: Specifications of the Fokker-27 model

2
Figure 1.2: Dimensions of cross section in mm
The wind tunnel spans across the whole building, since the motor and fan are on the ground level,
whereas the settling chamber and test section are on the second floor. The flow can be cooled, since the
turning vanes are equipped with a cooling system. The wind tunnel has been numbered to identify the
different parts in the figure below 1.3. The meaning of these numbers are presented below.

Figure 1.3: Schematic overview of wind tunnel


1. The fan is powered by a 525 kW DC motor

2. Turning vanes are used to alter the direction of the flow

3. The flow is accelerated by a six bladed fan

4. A spider web is placed behind the fan for safety purposes

3
5. An expansion screen is used to reduce the flow velocity before the flow interacts with the anti-
turbulence screens

6. 7 Anti-turbulence screens are used to reduce the flow turbulence

7. Storage room for the screens

8. Settling chamber

9. Contraction

10. Test section

11. Balance system

12. Diffuser

13. Security screen

1.2 Tests performed


The test matrix for the whole experiment is shown in table 2. For isolation of thrust the effect of induced
drag by lift had to be removed. Thus an additional run at zero CL was performed for the WFNVHP
configuration at angle of attack α = 4o .

Test Matrix
WFNV
V Angle of attack (α)(in degrees) ∆α
50 m/s -8 to 17 1
WFNVH
V Angle of attack (α)(in degrees) ∆α
50 m/s -8 to 17 1
WFNVHP (alpha sweep)
V α(in degrees) ∆α RPS
50 m/s -7 to 7 1 289
WFNVHP (rps sweep)
V α(in degrees) RPS ∆ RPS
50 m/s 0 220 to 320 10
WFNVHP (rps sweep)(zero lift)
V α(in degrees) RPS ∆ RPS
50 m/s -4 220 to 320 10

Table 2: Test matrix for Fokker F27 experiment

4
2 Wind Tunnel Corrections
The following wind-tunnel wall effects need to be accounted and corrected for while analyzing the aero-
dynamic characteristics of the Fokker F27 model:

• Blockage effects

– Solid model blockage


– Wake blockage

• Slipstream interference

• Streamline curvature

The respective corrections towards the performance coefficient determination are made as shown below.

2.1 Solid Model Blockage


The basic operating principle for applying this correction is to recognize the incompressible nature of the
air flow around the model placed within the wind tunnel. Due to the obstruction thus created, the free-
stream velocity across the model surface increases thus increasing the stresses over the model surface.
The change in velocity due to the presence of the solid model is given as [1, p. 368]:

∆V K1 τ1 (solid volume)
εsolid blockage = = (1)
V C 3/2
where, K1 is the body shape factor, τ1 is a factor determined by the wind-tunnel test-section shape
and the model specification (span-to-tunnel-width ratio). Furthermore, the model blockage effects of a
wing-body combination are additive in nature, that is:

εassembly = εwing + εbody (2)

The computation of these factors individually are carried out as discussed below.

2.1.1 Solid blockage effect: Wings


The wing volume is computed using the data provided within the model specifications (Table 1). The
computation of this volume is as follows:
" #
t
Vwing = S × Cmean × (3)
c
avg

where, S refers to the surface area of the wings and Cmean denotes the mean aerodynamic chord of
the airfoils.Furthermore, the terms t and c are the thickness and chord wing of the airfoil sections
(NACA 642-421 and NACA 642-415) used to compute the averaged thickness-to-chord ratio (over the
wing tip and root). The product of the mean aerodynamic chord along with this average ratio is taken to
be the equivalent rectangular thickness for the wing. The thickness-to-chord ratio at the wing tip and
root are obtained1 to be: 0.0942 and 0.1303 respectively. The average ratio thus turns out to be 0.11227.
Thus, as specified in the model dimensions, the surface area, S is taken to be 0.175 m2 and the
mean aerodynamic chord, Cmean is taken as 0.1285 m. The volume of the wing is thus obtained to be
2.5284 × 10−3 m3 . Further, the figures 2.1 and 2.2 provide the constant values for the body shape factor
1 This ratio was obtained by calculating the area of the airfoil section and dividing it with the square of the chord length, this

is equivalent to the integration of the thickness as it varies along with the chord distance and its ratio with the chord length.

5
!
t
Figure 2.1: Body shape factor (K) vs. thickness ratio c [1, p. 369]

(K ) and τ1 . The body shape factor is easy to trace for the average thickness-to-chord ratio (=0.1127) as
0.978. The factor τ1 is determined based on two other parameters —the ratio of tunnel width-to-height
( HB ) and the model span-to-tunnel-breadth ( 2b B ).
These parameters are calculated based on the model and tunnel dimensions specified earlier in section
1: B = 1.80 m, 2b = 1.45 m and H = 1.25 m. Based on these values, the parameters HB and 2b B are
computed as 1.44 and 0.805.The factor τ1 is then obtained as 0.885 from the figure 2.2.
Finally, in order to determine the solid blockage effect due to wing as given in equation 1, the term
C is evaluated as the area of the wind tunnel section (in favor of simpler computation the boundary layer
thickness is neglected for the free-stream flow). This wind tunnel section area is computed to be 2.07m2 .
The blockage effect εwing , is then given as:

0.978 × 0.885 × 2.5284 × 10−3


εwing = = 0.735 × 10−3 (4)
(2.07)3/2

2.1.2 Solid blockage effect: Fuselage


The fuselage blockage effect is computed similar to the wing effect as done previously. The fuselage is
assumed to be a streamlined body of revolution (similar to a cylinder) with its volume being expressed as:

Vfuselage = π (req )2 l (5)

where, the equivalent radius (req ) is taken to be the half of the average height and width of the fuselage,
whereas the length of the fuselage is represented by l. This volume is thus obtained to be 16.964 × 10−3 m3 .
Furthermore, the factors K and τ1 are identified for body of revolution (the straight line plot in figure
2.1) and for circular (figure 2.2) geometry. For obtaining the body shape factor the figure 2.1 is referred;
the x-axis corresponding to the graph plotted is evaluated using the parameter dl which is determined
to be 0.1184 (based on model dimensions). Thus, the body shape factor is obtained as 0.9154. Similarly,

6
Figure 2.2: τ1 for various tunnel types [1, p. 369]

the factor τ1 is obtained for a body of revolution ( 2b


B = 0) and whose cross-section remains the same as
B
before ( H = 1.44), giving τ1 = 0.862.
The blockage effect is then determined to be:

0.9154 × 0.862 × 16.964 × 10−3


εfuselage = = 4.494 × 10−3 (6)
(2.07)3/2
Thus the effective solid blockage is estimated to be the sum of the two effects:

εassembly = εwing + εbody = 5.229 × 10−3 (7)

thus, the change in freestream velocity (∆V ) is then:

∆V
εassembly = = 0.005229 (8)
V
thus, the corrected velocity is given as Vc = V (1 + ε ).

2.2 Wake Blockage


Similar to the solid model blockage discussed previously, the basic idea of this blockage effect is the
obstruction of the freestream flow area. The obstruction however, is caused due to the axially symmetric
expansion of the wake downstream of the aircraft model which "induces a pressure gradient in the stream-
wise direction which produces a drag increment on the body" [1, p. 370]. This phenomenon is modeled
using doubly-infinite source-sink system placed vertically as well as horizontally at the extremities of the
wind tunnel in each direction. A factor of 12 is used to determine the velocity induced by the sink image
system as it is assumed that the induced velocity is split in half in both the upstream as well as the
downstream direction. The final expression is obtained as follows:

∆V S
εwake = = CD0 (9)
V 4C
or in terms of the dynamic pressures:

qwake S
= 1+ CD0 (10)
q 2C

7
(a) Experimental values (WFNVH at V = 50ms−1 ) (b) Maskell’s method
Figure 2.3: Maskell’s analysis of wake blockage due to separated flow

0.3 1.6

0.25 1.5

0.2 1.4

\frac{q_c}{q_u}
0.15 1.3
wake

0.1 1.2

0.05 1.1

0 1

-0.05 0.9
-5 0 5 10 15 20 -5 0 5 10 15 20

qc
(a) εwb vs. α (b) qu vs. α
Figure 2.4: Wake blockage effect for WFNVH at V = 50ms−1

where, CDu is the uncorrected drag coefficient, whereas, S and C have the same meaning as in the solid
blockage discussion. Furthermore, Maskell identified the following formulation for the wake blockage for
the separated flow:
S 5S
εwake = CD0 + (CDu − CDi − CD0 ) (11)
4C 4C
qc S 5S
= 1+ CD0 + (CDu − CDi − CD0 ) (12)
qu 2C 2C
where, the coefficients CD0 and CDi are introduced as a means to analyze the drag over a lifting wing as
shown in figure 2.3. As can be seen in this figure, the coefficients help determine the drag contribution
from three sources: the zero-lift drag (CD0 ), the induced drag due to the wake (CDi ) and the drag due to
separation (CDs ). The value for CD0 is obtained as a constant for the zero (in experimental terms, minimum
CL2 when lifting begins (CL ≥ 0)).
The current examination of the wake blockage effect is done for the WFNVH configuration of the
model at a freestream velocity of 50ms−1 . The zero-lift drag is then identified as 0.04896, the induced
drag is determined as the slope of the linear shaded area marked in light blue in the figure 2.3a. This
gives, CDi = 0.02636. The values εwb and qquc are shown in figure 2.4. Similarly, for the configurations
WFNV and WFNVHP CDi is found to be 0.031806 and 0.02791 respectively.

2.3 Slipstream Interference


When propellers are running, the flow around the propellers produce slipstream which interferes with
the flow over the wing. The air in the streamtube of the propellers have higher velocity than that of
outside. The effect of the interference depends on the thrust by the propellers. In the experiment, thrust

8
was calculated in indirect way by considering the difference between the forward force in propellers ’on’
condition and ’off’ condition. The difference between the thrusts of powered (WFNVHP) and unpowered
(WFNVHP) conditions would determine the thrust purely due to the geometry. The thrust coefficient can
be calculated as follows:
4(CDof f − CDon )S
Tc = (13)
πDp2
where, CDof f is the is the coefficient of drag in the unpowered configuration, CDon is the is the coefficient
of drag in the powered configuration, Dp is the diameter of the propeller and S is the wing area. The
kinetic pressure difference in both the configurations can be determined as the ratio of velocity by the
interference (Uc ) and free stream velocity, which is as follows:
 √
Uc Dp ci  1 + Tc + 1 1/2 1/2
= 1 + np k √ Tc (14)
U b MAC 2 1 + Tc

where, np is number of propellers, k is shape factor (0.6) [2], b is wing span 1, ci is local chord at the
location and MAC is mean aerodynamic chord. Here the local chord is calculated assuming the wing to
be tapered with taper ratio of λ and is given as :
 2 
c ( y ) = cr 1 + ( λ − 1 ) y (15)
b
In the above equation we estimate the MAC by following relation which is then used to to calculate the
cr ,root chord of the model.
Z b
2 2
MAC = (c (y))2 dy (16)
b 0
and root chord of the model would be as follows:
s
MAC S
cr = (17)
2 2 + 2 (λ − 1) + b6 (λ − 1)2
b b

The corrected coefficient of the lift ∆CL is calculated with the above calculated velocity ratio and can be
written as : h U 2 i
∆CL = CL −1 (18)
Uc
The drag and moment corrections are as follows:

∆CL2
∆CD = (19)
πAR

∆CM = −0.25∆CL2 (20)


To summarize above correction method, thrust coefficient is estimated from the data of drag coefficients of
the two configurations and then it is used to calculate the velocity ratio. This velocity ratio is then used
to calculate the corrected lift with the help of above given equation 18.
From the above thrust correction for the WHNVP configuration, the lift coefficient is corrected to
reduce from the higher uncorrected value. This tells us that the lift was over-predicted due to the effect
of slipstream interference.

9
2.4 Streamline Curvature
As a result of circulation, lift is generated. This leads to curving of the streamlines. Inside the wind
tunnel, the curved streamlines are constrained by the walls of the tunnel. Due to presence of walls, the
streamlines are straightened. Due to which an increase in the lift is observed. It can be assumed that
the airfoil is a vortex at quarter chord point which leads to the up-wash at the wing. This causes the
resultant angle-of-attack to be more than the actual angle-of-attack. The resultant angle-of-attack can
be described as follows :
S
∆α = δ0 CLunc (21)
C
where δ is 0.132, C is 2.07 m2 and S is 0.175 m2 This effective angle-of-attack can be used to calculate
the corrected drag and lift coefficient. They can be written as follows:

CLcorr = CLunc cos∆α − CDunc sin∆α ≈ CLunc (22)

CDcorr = CLunc sin∆α + CDunc cos∆α ≈ CDunc + CLunc ∆α (23)


This tells us that the lift coefficient is not majorly affected by the change in angle-of-attack whereas the
drag coefficient is changed. The corrected Cm is given as follows:

∆Cm = −0.25∆CL (24)

10
3 Results and discussions
3.1 Configurations
In this section the performance curves of different configurations are compared and the differences are
discussed. We look at the effect of adding the horizontal tail wing and the propeller to the aircraft.

3.1.1 Effect of horizontal tail wing


The lift, drag and pitching moment polar for WFNV and WFNVH configurations are shown in figure 3.1
and 3.2 respectively. The tail acts as an additional lifting surface. At lower angles of attack, the tail
counteracts the lift produced by the main wing and total lift is lesser with the tail. At higher angles of
attack the tail compounds the lift and thus the maximum lift is increased.
Looking at the drag polar, the addition of tail increases the overall drag as expected. The pressure
drag and skin friction drag of the tail adds to the drag of the whole scaled model without the tail.
The pitching momentum of an aircraft gives a good indication of longitudinal stability of the aircraft.
A positive pitching moment coefficient tends to pitch the nose up and the opposite happens for negative
pitching moment. The configuration without the tail has a positive slope moment polar and the aircraft is
unstable. Adding the tail makes the moment polar with a negative slope increasing the stability of the
aircraft. The aircraft returns to the original stable position when an instability is introduced. The effect
of a disturbance is negated by the tail.

WFNV 0.40 WFNV


1.25 WFNVH WFNVH
0.35
1.00
0.30
Coefficient of Drag (CD)
Coefficient of Lift (CL)

0.75
0.25
0.50
0.20
0.25
0.15
0.00
0.10
−0.25
0.05
−0.50
−5 0 5 10 15 −5 0 5 10 15
Angle of attack (α) Angle of attack (α)

(a) Lift polar for WFNV and WFNVH. (b) Drag polar for WFNV and WFNVH.
Figure 3.1: Effect of horizontal tail on the performance curves.

0.2

0.1
Coefficient of Moment (Cm)

0.0

−0.1

−0.2

−0.3

−0.4 WFNV
WFNVH
−5 0 5 10 15
Angle of attack (α)

Figure 3.2: Effect of horizontal tail on pitching moment.

11
3.1.2 Effect of propeller
Similar to previous section, the effect of propeller is studied by comparing the performance curves of
configuration with and without propeller. The lift curve for WFNVH and WFNVHP is shown in figure 3.3a.
The addition of propeller increases the lift slightly. This can be attributed to the increase in effective
angle of attack of the wing in the wake of the propeller.
Figure 3.3b shows the drag polar for the two configurations. The addition of propeller decreases the
effective drag measured by the balances. Thus the drag coefficient for the configuration with propeller on
is much lesser than that without it.

WFNVH 0.40 WFNVH


1.25 WFNVHP WFNVHP
0.35
1.00
0.30

Coefficient of Drag (CD)


Coefficient of Lift (CL)

0.75 0.25

0.50 0.20

0.25 0.15

0.10
0.00
0.05
−0.25
0.00
−0.50
−5 0 5 10 15 −5 0 5 10 15
Angle of attack (α) Angle of attack (α)

(a) Lift polar for WFNVH and WFNVHP. (b) Drag polar for WFNVH and WFNVHP.
Figure 3.3: Effect of propeller on the performance curves.
The pitching moment coefficient for the two configuration is shown in figure 3.4. The configuration
with propeller on is seen to be less stable than without it. The presence of propeller increases the lift on
the wings and the tail thus adding to the positive pitching moment.

0.2
WFNVH
WFNVHP
0.1
Coefficient of Moment (Cm)

0.0

−0.1

−0.2

−0.3

−0.4

−5 0 5 10 15
Angle of attack (α)

Figure 3.4: Effect of propeller on pitching moment.

3.2 Blockage Correction


The given plots 3.5a, 3.5b and 3.6 describe the effect of wake and solid blockage correction on CL , CD and CM
respectively in WFNV, WFNVH and WFNVHP configurations. It is evident that in the experimental con-
ditions clearly the coefficients are over-predicted due to increase in velocity due to blockage.

12
1.2 Blockage Correction 0.35 Blockage Correction
Uncorrected Values Uncorrected Values
1.0 0.30

0.8

Coefficient of Drag (CD)


0.25
Coefficient of Lift (CL)

0.6
0.20
0.4

0.2 0.15

0.0 0.10

0.2
0.05
−5 0 5 10 15 −5 0 5 10 15
Angle of attack (α) Angle of attack (α)

(a) Correction for Coefficient of lift (b) Correction for Coefficient of drag
Figure 3.5: Blockage correction for CL and CD of WFNV configuration

Blockage Correction
Uncorrected Values
0.1
Coefficient of Moment (Cm)

0.0

−0.1

−0.2

−5 0 5 10 15
Angle of attack (α)

Figure 3.6: Blockage correction CM for WFNV configuration

Blockage Correction 0.40 Blockage Correction


1.25 Uncorrected Values Uncorrected Values
0.35
1.00
0.30
Coefficient of Drag (CD)
Coefficient of Lift (CL)

0.75
0.25
0.50
0.20
0.25
0.15
0.00
0.10
−0.25
0.05
−0.50
−5 0 5 10 15 −5 0 5 10 15
Angle of attack (α) Angle of attack (α)

(a) Correction for Coefficient of lift (b) Correction for Coefficient of drag
Figure 3.7: Blockage correction for CL and CD for WFNVH configuration

13
0.2
Blockage Correction
Uncorrected Values
0.1

Coefficient of Moment (Cm)


0.0

0.1

0.2

0.3

0.4

−5 0 5 10 15
Angle of attack (α)

Figure 3.8: Blockage correction for CM for WFNVH configuration

1.2 0.03
corrected corrected
1.0 uncorrected uncorrected
0.02
0.8

0.6
0.01
CD

0.4
CL

0.2 0.00

0.0
−0.01
−0.2

−0.4
−6 −4 −2 0 2 4 6 −6 −4 −2 0 2 4 6
α α

(a) Correction for Coefficient of lift (b) Correction for Coefficient of drag
Figure 3.9: Blockage correction for CL and CD for WFNVHP configuration

0.16 corrected
uncorrected
0.14

0.12

0.10
Cm

0.08

0.06

0.04

−6 −4 −2 0 2 4 6
α

Figure 3.10: Blockage correction for CM for WFNVHP configuration

3.3 Streamline Curvature


From the plots for CL andCD it is evident that after correcting for the streamline curvature the value of
CL is not affected much whereas significant change in CD is observed in the plots. Since the correction
for Cm depends on change in CL , it also follows the same trend as CL of no significant change in the
value after the correction. Comparing the drag polar with blockage correction and streamline curvature, it
is evident that streamline curvature has much more influence on the drag than blockage correction. The
streamline curvature increases the effective angle of attack which increases the total drag.

14
1.2 Streamline Correction 0.35 Streamline Correction
Uncorrected Val es Uncorrected Values
1.0
0.30
0.8

Coefficient of Drag (CD)


Coefficient of Lift (CL)

0.25
0.6
0.20
0.4

0.2 0.15

0.0 0.10

−0.2
0.05
−5 0 5 10 15 −5 0 5 10 15
Angle of attack (α) Angle of attack (α)

(a) Correction for Coefficient of lift (b) Correction for Coefficient of drag
Figure 3.11: Streamline curvature correction for CL and CD for WFNV configuration

Streamline Correction Streamline Correction


1.25 Uncorrected Values
0.40 Uncorrected Values

1.00 0.35
Coefficient of Drag (CD)

0.30
Coefficient of Lift (CL)

0.75

0.50 0.25

0.20
0.25
0.15
0.00
0.10
−0.25
0.05
−0.50
−5 0 5 10 15 −5 0 5 10 15
Angle of attack (α) Angle of attack (α)

(a) Correction for Coefficient of lift (b) Correction for Coefficient of drag
Figure 3.12: Streamline curvature correction for CL and CD for WFNVH configuration

1.2
Streamline Correction Streamline Correction
0.05
Uncorrected Values Uncorrected Val es
1.0
0.04
0.8
Coefficient of Drag (CD)

0.03
Coefficient of Lift (CL)

0.6
0.02
0.4

0.2 0.01

0.0 0.00

0.2 −0.01

0.4
−6 −4 −2 0 2 4 6 −6 −4 −2 0 2 4 6
Angle of attack (α) Angle of attack (α)

(a) Correction for Coefficient of lift (b) Correction for Coefficient of drag
Figure 3.13: Streamline curvature correction for CL and CD for WFNVHP configuration

3.4 Slipstream Interference


From the results of the corrections for the drag coefficient it can be seen that the drag coefficients do not
have significant effect. Since the lift is reduced after the slipstream interference correction, effectively the

15
pitching moment increases. The correction in drag coefficient is mainly due to change in induced drag.
But the change in magnitude of induced drag is too small to have a significant effect. To conclude above
discussion, it can be said that the presence of propellers do not have much of effect on drag coefficient
whereas the pitching moment and coefficient of lift have considerable effect.

1.2
corrected corrected
1.0 uncorrected uncorrected
0.02
0.8

0.6 0.01

Cd
0.4
Cl

0.2 0.00

0.0
−0.01
−0.2

−0.4
−6 −4 −2 0 2 4 6 −6 −4 −2 0 2 4 6
α α

(a) Correction for Coefficient of lift (b) Correction for Coefficient of drag
Figure 3.14: Thrust correction for CL and CD due to propeller slipstream interference

corrected
uncorrected
0.14

0.12

0.10
Cm

0.08

0.06

0.04
−6 −4 −2 0 2 4 6
α

Figure 3.15: Thrust Correction for Pitching moment due to propeller slipstream interference

16
4 Noise
Three microphones were installed on the upper wall of the windtunnel. Mic 1 is located upstream of the
propeller, Mic 2 is in the vicinity of the propeller (in the propeller plane) and Mic 3 is downstream of the
propeller. The placement of the microphones is shown in figure 4.1.

Figure 4.1: Microphone placement relative to the propeller.


Sound data was collected in four distinct configurations at a sampling rate of 51200Hz. These
configurations are:

• WFHNVP, zero angle of attack α, with rotational rates from nl = 220rpm to nu = 320rpm in steps
of ∆u = 20rpm.

• WFHNVP, zero lift L, with rotational rates from nl = 220rpm to nu = 320rpm in steps of ∆u =
20rpm.

• WFHNVP, the propeller not being powered. The angles of attack range from αl = −5 deg to
αu = 7 deg in steps of ∆α = 3 deg.

• WFHNV, angles of attack ranging from αl = −8 deg to αu = 17 deg in steps of ∆α = 3 deg.

This section will first analyse the data globally. After the main characteristics are identified, the data
will be postprocessed to allow for some generalisations with respect to generated noise by the aircraft
and the propellers.

4.1 Generic Investigation


10s of data were obtained at a sampling rate of fs = 51200Hz, yielding N = 512000 samples. Performing
a Fourier transform in the form of a single sided spectrum will yield a maximum resolvable frequency of
25600Hz according to the Nyquist criterion, slightly above the standard upper limit for human hearing.
That is, given a discrete time series of pressure
measurements pi , the discrete Fourier transform

produces P̃ (fi ). The magnitude P (fi ) = P̃ (fi ) indicates the intensity of each mode. As a quantity that
is more easily interpretable, and somewhat corresponds to the manner in which the human ear perceives
the scaling of sound level, the sound pressure level is defined in dB as:
" #  
P ( fi ) 2 P ( fi )
SPL = 10log10 = 20log10 (25)
p0 p0

Here we’ve used the fact that F {ag (x )} = aF {g (x )}. Using the presented equations, the sound
pressure level is non-dimensionalised with respect to the lower threshold of human hearing p0 = 20µPa.

17
In the two configurations where a propeller is running, the frequency can be non-dimensionalised using
the propeller rotational frequency as f /n. For the other two cases the frequency is rid of its dimensionality
using the Strouhal number St = f L/U.
Taking the WFHNVP configuration at zero angle of attack α, n = 300rpm, and the WFHNV configu-
ration at the angle of attack α = 1 deg as representative samples, the following graphs are produced:

Figure 4.2: SPL as a function of frequency with. a powered propeller at α = 0 deg, n = 300rpm

Figure 4.3: SPL as a function of frequency without an installed propeller at α = 1 deg


The plots on the left show the low-frequency components while the plots on the right show the general
trend for the higher frequencies. One can see a general downward trend in sound energy: the lower
frequency components contain the bulk of the sound energy.
Figure 4.2 clearly shows some dominating frequencies at integer multiples from the propeller frequency
n, more specifically at 4n, which is the blade passing frequency. There are various large components
at 25Hz, 50Hz, 100Hz and diminishing at 150Hz and 200Hz. These components are influenced by the
propeller frequency and the microphone position. The 25Hz component stays reasonably constant while
the 50Hz and 100Hz increase with propeller frequency n and as the microphones lie further downstream.
The components lie too far away from multiples of 50Hz to be caused by the mains frequency. As
these components are also present without propeller these components are interpreted as caused by
aerodynamic effects of the aircraft, the propeller amplifying them through aerodynamic interaction.

18
4.2 WFHNV Configuration
The Fourier components P (fi ) across the frequency spectrum feature too much variation to be graphically
analysed (as visible in figures 4.3 and 4.2). Taking a moving average loses some of the data fidelity, but
reveals some interesting effects. The variance was investigated as well, but only showed that variance
follows the same general trend as the average sound pressure level: higher at low frequencies and
decreasing as the frequency is increased.

Figure 4.4: Full frequency range of noise measurements in the WFHNV configuration at all angles of
attack α.
These values are averaged over 0.5% of the frequency range, equal to ±100Hz, hence the peak
locations appear smaller than they are. Over the complete frequency range the sound level increase
slightly per frequency component. The more interesting frequency range lies at 10 < St < 20. Taking a
better look yields the following picture.

Figure 4.5: Frequency subrange of noise measurements in the WFHNV configuration at all angles of
attack α. Note the color scheme for easy interpretation.
Here we come to the interesting conclusion that aerodynamic noise produced by the aircraft differs
in frequency for negative angles of attack (at St ≈ 14 and St ≈ 18) and for positive angles of attack. As
expected small angles of attack produces the least amount of noise, as no severe unsteadiness is present.
As the angle of attack becomes more negative local separation causes the magnitude of these frequencies

19
to increase. Interestingly enough the frequencies increase as a positive angle of attack α becomes larger.
As the rough range over which the amplitudes increase is 10dB, these are significant changes.
We will see in the next section that the peaks near St = 17 are due to the aerodynamic interaction
with the propeller support.

4.3 Unpowered WFHNVP Configuration


The configuration including the unpowered propeller is very similar to the one presented in section 4.2.
The difference lies in the presence of the noise present due to the freely spinning propeller. The overall
frequency spectrum is very similar to the one in figure 4.4. Upon re-investigating the frequency spectrum
between 10 < St < 20, as in figure 4.5, we obtain the following results:

(a) Measurements from microphone 1 (b) Measurements from microphone 3


Figure 4.6: Frequency subrange of noise measurements in the WFHNVP configuration at all angles of
attack α.
The difference here is quite obvious. The inclusion of a propeller causes the components near St = 17
to dominate. This is the reason for assuming that the components near St = 17 in figure 4.5 are caused
by the propeller supports. Further investigation may yield more definitive conclusions.
The overall difference in noise between microphone 1 and 3 is roughly 2dB, but near the peak at
St = 17 the difference is 6dB. Given that the human ear perceives a difference in 10dB as twice as loud,
this a significant presence of sound directionality.

4.4 Powered WFHNVP Configuration, Zero Angle of Attack


From the initial investigation it is clear that the modes that are multiples of the propeller rotational
rate are the dominant, a reasonable expectation. The results presented in this chapter are based on the
investigation of the frequency spectra from all microphones at all propeller rotational rates, as can be
found in appendix A. Picking the non-dimensionalising unit rpm is a bit unnatural, choosing a value in
terms of rps would make more sense. But it appears that the waveform produced by the blades is more
interpretable in terms of rpm.
Given the propeller modes f /n, one can reasonably expect that the blade passing frequency (4n,
given that there are four blades per propeller) will be somewhat dominant. Higher frequency components
that are multiples of 4n shape the audible waveform due to the blades. But aerodynamic interaction with
the aircraft body, the windtunnel, irregularities in the blades and the flow-field generated by preceding
blades can cause more interesting waveforms. Extracting these modes for further investigation yields the
following figure.

20
Figure 4.7: Magnitudes of integer multiples of the propeller rotational rate, represented as the colored
surface. The white surface represents the µ + 2.5σ values. Where µ is the moving average and σ the
moving standard deviation of the spectrum component magnitudes.
The figure is quite telling. One can see that at low rotational rates n the overall background noise
is at the same level as the modes caused by the propeller. However, as the rotational rate is increased
the integer multiples of the blade passing frequency become statistically significant. Although slightly
unscientific, playing the audio samples confirm that at n = 220 and n = 240 the propeller noise is
barely audible, drowned out by the wind-tunnel noise. At n = 260 and n = 280 the propeller noise
becomes distinct from the background noise. Furthermore, as we’re looking at frequencies larger than
f /nprop > 20 one can see that the components become much less loud than the dominating ones, and
they start to recede into the statistical background.
A second marked increase is present at n = 300rpm. One finds that the waveform becomes much
more complex: the components other than the 4n ones lift themselves from the background noise as well.
This is also clearly visible in the following figures:

Figure 4.8: WFHNVP configuration, n = 260rpm

21
Figure 4.9: WFHNVP configuration, n = 300rpm

Figure 4.10: WFHNVP configuration, n = 320rpm


From these figures one reaches the same conclusions as one would from figure 4.7, but now within
the context of the entire frequency spectrum. In figure 4.8, at n = 260rpm, only a few components are
dominant. Figure 4.9 shows that as the rotational rate is increased the audible waveform produced by
the propeller becomes more complex, although the components that are not integer multiples of 4n are
not as dominant as the multiples of 4n. Finally, as the rotational rate is increased to its tested maximum
of n = 320rpm, figure 4.10 shows that the non-integer multiples of 4n become relatively more dominant.
One can only guess what is happening at the propeller without a more detailed investigation. There might
be some aeroelastic effects, the propellers may be partially stalled or the propellers might be causing a
complex interaction with the wings.
An attempt was made to determine if any data could be obtained regarding the directionality of the
noise caused by the propeller. Extracting the components at f = 4n yields the following figure.

22
Figure 4.11: Magnitude of the f = 4n components of the WFNVHP configuration for the three
microphones.
The same type of graph was plotted for components other than those at f = 4n, but amount to the
same conclusion. One might be inclined to conclude that at lower rotational rates the propeller noise tend
to radiate downstream more loudly than upstream and vice versa at higher rotational rates. However, the
small amount of measurements makes definitive statements unwise.

4.5 Conclusions
As an educational exercise, it is clear that performing sound measurements is quite complicated. The
dominating background noise caused by the windtunnel itself makes a global investigation (in terms of
Overall Sound Pressure Level (OASPL)) unreasonable. The conclusions reached therein might be valid,
but acoustic interactions between the model and the windtunnel would remain uninvestigated. It would be
helpful to perform future investigations after a baseline noise spectrum of the tunnel itself is established.
The most obvious result arising from an analysis of sound is that one can establish its perceived level by
humans. This can help in determining the operating range of aircraft components without being perceived
as annoying. It may also guide an engineering process into determining which aircraft components have
to be redesigned in order to reduce noise
Various different types of analysis have to be performed to extract data from the noise. Non-
dimensionalisation helps with regard to the investigation of the source of noise in the case of a powered
propeller. A statistical approach is more likely to yield results with a small number of microphones due
to the noise present in the data, it also helps in analysing data with a varying test variable (e.g. angle of
attack, rotational rate) if the data itself is too noise to be interpreted directly. In the case that an array
of microphones is present, one might be able to deduce the sources of the various frequency components,
this would help in distinguishing noise produced by a component from noise produced due to interactions
with other components.
Lastly, the investigation of noise can clearly show the onset of aerodynamic effects that require further
study. The same results may be obtained by a wide variety of other investigative methods (e.g. PIV, wake-
rake, LDA, etc.) but if one is in the early stages of design where experimental studies are performed, the
installation of a couple of microphones is a cheap way to enhance the obtained data. This may indicate
if aeroelastic effects, separation or unwanted vortices are present.

23
5 Scaling Effects
An integral part of flow measurement in wind tunnels is the post processing of data and scaling them into
actual flight conditions. To simulate the real flight condition in a wind tunnel we mainly use dimensionless
numbers - Reynolds Number, Mach number, Strouhal number and many more. A combination of these
numbers are used to simulate the real flight condition by keeping them as equal as possible to what the
actual flight experience. But due to limitations and errors due to various effects addressed before, the
wind tunnel conditions may not exactly capture the real flight conditions. These are affected by scaling
effects. The sheer difficulty in finding a correlation between real flight tests and wind tunnels makes
these effects a headache for aerodynamicists and experimentalists equally. Lack of these data arises
cause of different reasons. One owing to the fact that aviation companies keep these data to themselves
and two the difficulty in repeatability of the experiments both in real flight test and wind tunnel. Although
optimization of the design is achieved using wind tunnel testing of a model, we are not assured these
improvements in real flight conditions due to these scaling effects. Thus this remains an important area
to consider while wind tunnel testing.

5.1 Similarity
• Geometric similarity- In wind-tunnel testing, the impracticality to test the actual aircraft left us
with scaling the aircraft to a 1:20 model which is being tested. For this we scale all the dimensions
by the same scale ratio, thus achieving geometric similarity. But problem arises when complicated
details present in aircraft like, hinges, surface roughness, inspection doors and a dozens of other
surface features. These things affect the flow on an actual aircraft and changes flow characteristics
which are not captured in model testing.

• Kinematic Similarity- The kinematic quantities of the flow should be proportional at all points on
the aircraft. This is achieved by keeping the velocities at all points on the model and actual aircraft
proportional.

• Dynamic similarity- This is the similarity we are most interested in for this particular experiment.
If the first two types of similarities are satisfied this automatically ensures dynamic similarities.

For satisfying the similarities of the flow, the most important dimensionless numbers to look out for
are Reynolds number, Mach number and Strouhal number. Strouhal number becomes important when
there are unsteady variation or forces. Thus we can neglect the effects of Strouhal number on scaling.
For a propeller driven aircrafts flying at low speeds, Mach number also has a small effect. The most
important parameter to consider while coping for errors while scaling is the Reynolds number effects. In
the subsequent sections we address different aircraft performance parameters dependence on Reynolds
number.

5.2 Boundary layer


Boundary layer grows thicker as Reynolds number decreases. Wind tunnel tests usually has a lower
Re than in real flight tests. Thus proper scaling of lift, drag, power, stability factors need to be done
before manufacturing the aircraft. Basically four sources of direct scaling effects on boundary layer are
identified[3].

• Conventional scale effects, which considers the decrease in boundary layer thickness as Re increases.
The reduction in BL thickness towards the trailing edge helps the flow tp withstand high pressure
gradient preventing separation.

24
• Bubble dominated effects deals with laminar bubble formation on the suction side of the airfoil. The
size of bubble decreases with increasing Re and thus is a favourable scale effect. But this too leads
to underestimation of lift and drag of the aircraft.
• Slot flow dominated scale effects are not relevant for our experiment and will not be considered.
• Transition of flow due to scaling effects on low sweep wings can be considered a favourable
factor. The transition location moves further upstream with increasing Reynolds number. Thus
an ameliorating condition is that, if separation doesn’t happen on the model at low Re, then the
chances of separation on the real aircraft is really low. Early transition can be brought about in
the model by using tripping surfaces such as grit, wire, thread, triangles etc. But the exact location
of transition on the aircraft is difficult to quantify.
The scaling of separation in wind tunnel testing is difficult as separation occurs sooner in models. Therefore
there are no standard practices to delay the separation in models to match with real flight[1].

Figure 5.1: CL2 vs CD for NACA 23012 for two Re [1]

5.3 Drag
Although there are more important performance parameters to be considered while addressing scaling
effects, the fuel efficiency which depends on the drag in a fairly important parameter. The minimum drag
of the aircraft depends on numerous small parts on the aircraft which cannot be accurately represented
on a model. But an engineer assumes it to be compensated by rise in Reynolds number[1].
The rate of change of drag with lift also displays some interesting variation with Reynolds number.
The rate of change is represented by span efficiency factor e given by,
1
e=
dCD /dCL2 πAR
where AR is the aspect ration of the wing. If we plot CL2 vs CD we get the curve in figure 5.2.
An interesting fact is the independence of the slope of the curve with Reynolds number, although the
curve shift with increasing Re which doesn’t have a good rule[1]. Another option to estimate CD is to run
the model tests at different flight speeds to produce a logarithmic curve between CD0,min vs Re as shown
in figure 5.3 and extrapolate to find approximate full scale minimum drag coefficient.
Although these methods exist, its still difficult to scale drag coefficients due to presence of lot of
components. The drag dependence with lift exists for all parts of aircraft such as fuselage, horizontal tale,
pitot tube, nacelles etc. The scaling of drag follows from experience and testing existing flight design
and correlate with wind tunnel test. A final optimized design is not just minimum drag, but depends on
stability, complexity, weight, function etc.

25
Figure 5.2: CL2 vs CD for NACA 23012 for two Re [1]

Figure 5.3: CDT vs log Re[1]

5.4 Lift
Effect of Reynolds number on lift is really strong and demands close attention. As Reynolds number
increase the lift curve becomes more and more linear and the stall becomes more steeper. The lift curve
obtained from model testing extends to larger angles of attacks linearly and the C lmax and angle of
maximum lift also increases. Even though this is a positive scaling effect, danger lies in steepness of stall.
From figure 5.4 we can observe the increase in stall angle and the steepness.
There are methods based on calculating ∆CL at the required Reynolds number and modifying the
lift curve to find stall angle of full scale aircraft’s. But this requires information of Reynolds number
characteristics of airfoils. Because most designers use custom modified airfoils a universal relation is hard
to find.

26
Figure 5.4: CL vs Re[1]

5.5 Stability and control


The longitudinal and lateral stability of the aircraft in real flight matches with that of model wind tunnel
test around Reynolds number of order of 106 . Usually the static longitudinal stability is slightly more
in full scale than in model. Contrary to this the directional stability is seriously affected by Reynolds
number scaling. It is very much less in flight compared to that predicted in model.

27
6 Conclusions
A wind tunnel experiment was performed on a 1:20 scaled model of Fokker F27 aircraft. From the
experiment, important performance parameters of an aircraft such as lift coefficient, drag coefficient, pitching
moment coefficient, propeller thrust, etc was obtained. These parameters were corrected for blockage
effects, streamline curvature and slipstream interference due to effect of propeller. This task helped
in understanding the importance of individual corrections on estimating the performance of an actual
aircraft. few corrections such as buoyancy effects, strut interference, horse-shoe vortex in the wall were
not considered to keep the study simple. Although the corrections are correlated, a linear addition of the
individual corrections was done to calculate the final corrections to the parameters. The lift coefficient
was lowered after applying the streamline curvature corrections. The wind tunnel walls straighten the
streamlines of the flow, which effectively lead to camber being more than that of the actual model. The
drag coefficient is not affected significantly and the coefficient of pitching moment is under-predicted.
The corrected values are not representative of the actual performance of the full scale model because the
experiment was not performed at relevant Reynolds number, mach number and Strouhal number. Thus
proper scaling effects should be considered while correlating these parameters to the full scale aircraft.
Thrust calculations can be done by comparing the data from the model with thrust and without thrust
which is calculated at angle of attack being zero. It can be concluded that the increase in the lift for the
model is caused by the thrust and also leads to change in coefficient of moment. Along with performance
measurements, acoustic data was also obtained with and without propeller using three microphones in
the vicinity of the propeller.

28
7 Recommendations
The current experiment reveals important information about the benefits and limitations of performing
wind tunnel experiments to study the performance of the aircraft. In applying corrections to the acquired
performance data several assumptions were taken regarding the volume of the model, and estimation of
thrust.
The experiment was performed at Reynolds number much lower than what is experienced by the actual
aircraft in flight. Relevant similarity parameters should be matched to acquire important performance
parameters. In the case of this aircraft, matching Reynolds number and mach number would be important.
matching Reynolds number helps in estimating the separation location without using tripping strips on
the wing. The effect of mach number on the performance coefficients is significant for mach number above
0.3. Since most aircraft fly close to a mach number of 0.8, this seems pretty important to investigate. This
can be achieved by performing the experiment at right combination of velocity and temperature.
In the experiment, while applying for blockage corrections the volume of the model was approximated
which may lead to certain degree of errors in applying blockage corrections. Also the strings which were
attached to the model were never considered as a part of any corrections. The blockage effects due to
them can be considered as part of blockage corrections.
In the measurement of noise, it was observed that there was significant background noise in the wind
tunnel caused by aerodynamic interaction between wind tunnel walls and the model. It would be helpful
to perform future investigations after a baseline noise spectrum of the tunnel itself is established.

29
A WFHNVP Noise Measurements
The following figures were obtained from the WFHNVP configuration from three different microphones at
all tested propeller rotational rates.

A.1 n = 220rpm

Figure A.1: WFHNVP configuration, α = 0 deg, n = 220rpm, microphone 1

Figure A.2: WFHNVP configuration, α = 0 deg, n = 220rpm, microphone 2

30
Figure A.3: WFHNVP configuration, α = 0 deg, n = 220rpm, microphone 3
A.2 n = 240rpm

Figure A.4: WFHNVP configuration, α = 0 deg, n = 240rpm, microphone 1

Figure A.5: WFHNVP configuration, α = 0 deg, n = 240rpm, microphone 2

31
Figure A.6: WFHNVP configuration, α = 0 deg, n = 240rpm, microphone 3
A.3 n = 260rpm

Figure A.7: WFHNVP configuration, α = 0 deg, n = 260rpm, microphone 1

Figure A.8: WFHNVP configuration, α = 0 deg, n = 260rpm, microphone 2

32
Figure A.9: WFHNVP configuration, α = 0 deg, n = 260rpm, microphone 3
A.4 n = 280rpm

Figure A.10: WFHNVP configuration, α = 0 deg, n = 280rpm, microphone 1

Figure A.11: WFHNVP configuration, α = 0 deg, n = 280rpm, microphone 2

33
Figure A.12: WFHNVP configuration, α = 0 deg, n = 280rpm, microphone 3
A.5 n = 300rpm

Figure A.13: WFHNVP configuration, α = 0 deg, n = 300rpm, microphone 1

Figure A.14: WFHNVP configuration, α = 0 deg, n = 300rpm, microphone 2

34
Figure A.15: WFHNVP configuration, α = 0 deg, n = 300rpm, microphone 3
A.6 n = 320rpm

Figure A.16: WFHNVP configuration, α = 0 deg, n = 320rpm, microphone 1

Figure A.17: WFHNVP configuration, α = 0 deg, n = 320rpm, microphone 2

35
Figure A.18: WFHNVP configuration, α = 0 deg, n = 320rpm, microphone 3

36
References
[1] J.B. Barlow, W.H. Rae, and A. Pope. Low-Speed Wind Tunnel Testing. Aerospace engineering, me-
chanical engineering. Wiley, 1999. isbn: 9780471557746. url: https://books.google.nl/books?
id=LBMoAQAAMAAJ.
[2] W. Kuhn D. Eckert G.H. Hegen. “DNW’s method to correct for support and wall interference effects
on low speed measurements with large propeller powered transport aircraft model”. In: ICAS (2006).
[3] Ralf Rudnik and Eric Germain. “Re-No. Scaling effects on the EUROLIFT high lift configurations”. In:
45th AIAA Aerospace Sciences Meeting and Exhibit. 2007, p. 752.

37

You might also like