You are on page 1of 83

Experimental Aeroacoustic Study of a

Landing Gear in the Unsteady Flow


Induced by a Propeller

by

Rafik Chekiri

A thesis submitted in conformity with the requirements


for the degree of Master of Applied Science
Graduate Department of Aerospace Studies
University of Toronto

April 2014 by Rafik Chekiri

Abstract
Experimental Aeroacoustic Study of a Landing Gear
in the Unsteady Flow Induced by a Propeller
Rafik Chekiri
Master of Applied Science
Graduate Department of Aerospace Studies
University of Toronto
An aeroacoustic study of a two-strut, two-wheel, nacelle-mounted landing gear was
conducted to investigate the effects of an upstream propeller on the radiated noise. The
development of a 1:10.8 scale model based on a Bombardier Q400 aircraft, consisting of a
propeller, motor, nacelle, and landing gear assembly is discussed. Comparisons are made
between cases with and without an actuated upstream propeller. Far-field microphone
measurements out of the airstream are presented to characterize the acoustic effects of
each model component. The main strut and wheels of the model were equipped with
surface-mounted microphones to measure unsteady pressures. It is shown that the noise
signature of the landing gear cannot be observed over the tunnel background noise in the
far-field. Unsteady surface pressures on the main strut show dominant peaks related to
vortex shedding from the drag strut for both steady and unsteady upstream conditions.

ii

Acknowledgements
This work was funded through the Green Aviation Research and Development Network
(GARDN), the Natural Sciences and Engineering Research Council of Canada (NSERC),
Aercoustics Engineering Ltd., and Bombardier Aerospace. Thank you to Dr. Werner
Richarz of Aercoustics Engineering Ltd., as well as Stephen Collavincenzo and Dr. Raymond Wong of Bombardier Aerospace for their technical suggestions and guidance.
I would like to thank my good friend, Cameron Robertson, for bringing this project
to my attention and putting me in contact with my supervisor, Dr. Philippe Lavoie.
Dr. Lavoie has provided the best kind of mentorship and understanding throughout the
innumerable tribulations of this work, for which I am tremendously grateful. I must also
thank Dr. Alis Ekmekci for her prompt refereeing of my thesis and useful feedback.
Joining the Flow Control and Experimental Turbulence group at the Institute for
Aerospace Studies has been an unforgettable experience. The attitude, knowledge, and
generosity of the team made it a wonderful place to work and share ideas. I would
particularly like to thank Nicole Houser, Jason Hearst, Heather Clark, and Derrick Chow
for their insight in solving technical problems and assistance in editing this work. The
friendships from my three years at UTIAS will always be treasured. Thank you to Nicole
Houser for working late nights and early mornings with me, keeping me focussed, teaching
me about life, and being an amazing partner.
Lastly, I must thank my parents for continually insisting on the importance of knowledge and perseverance. Your unceasing support and encouragement, even from a distance, has kept me happy, productive, and sane. I love you both very much.

iii

Contents
1 Introduction
1.1 Aircraft Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2 Background
2.1 Aerodynamic Sources of Sound
2.2 Bluff-body Flows . . . . . . . .
2.2.1 Isolated Cylinders . . . .
2.2.2 Tandem Cylinders . . .
2.3 Landing Gear Studies . . . . . .
2.3.1 Simplified Geometries .
2.3.2 High Fidelity Geometries
2.4 Propeller Flows . . . . . . . . .
2.4.1 Flow Structures . . . . .
2.4.2 Acoustics . . . . . . . .

1
1
2
3

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

4
4
5
5
8
11
12
12
13
13
16

. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
Turbulence Levels
. . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

17
17
19
19
20
21

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

25
25
25
29
29
29
31

5 Results & Analysis


5.1 Far-field Acoustics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.1 Static Configurations . . . . . . . . . . . . . . . . . . . . . . . . .

37
37
37

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

3 Wind Tunnel Characterization


3.1 Facility Overview . . . . . . . . . . .
3.2 Aerodynamic Survey . . . . . . . . .
3.2.1 Experimental Method . . . .
3.2.2 Mean Flow Measurements and
3.3 Background Acoustic Levels . . . . .

.
.
.
.
.
.
.
.
.
.

4 Experimental Method
4.1 Overview . . . . . . . . . . . . . . . . .
4.2 Model Description . . . . . . . . . . .
4.3 Instrumentation . . . . . . . . . . . . .
4.3.1 Far-field Microphones . . . . . .
4.3.2 Surface Microphones . . . . . .
4.3.3 Data Acquisition & Uncertainty

iv

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

5.2

5.1.2 Motor and Propeller Induced Noise


Unsteady Surface Pressures . . . . . . . .
5.2.1 Strut Surface Microphones . . . . .
5.2.2 Wheel Surface Microphones . . . .

6 Conclusions
6.1 Wind Tunnel Characterization . . .
6.2 Model Development . . . . . . . . .
6.3 Landing Gear Testing . . . . . . . .
6.4 Recommendations for Future Work
References

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

39
39
42
44

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

52
52
52
53
53
55

Appendices
A Wind Tunnel Layout

60

B Surface Microphone Calibration

62

C Landing Gear Model Drawings

64

List of Tables
2.1
2.2

Flow regimes of a disturbance-free flow around a circular cylinder. . . . .


Description of tandem cylinder spacing effects. . . . . . . . . . . . . . . .

6
9

4.1
4.2

Experimental test matrix for acoustic measurement tests. . . . . . . . . .


Spacing ratios at the unsteady surface pressure locations. . . . . . . . . .

28
30

vi

List of Figures
1.1

Bombardier Dash 8 Q400 aircraft. . . . . . . . . . . . . . . . . . . . . . .

2.1
2.2
2.3
2.4
2.5
2.6
2.7
2.8
2.9

Regions of disturbed flow surrounding a circular cylinder. . . . . . . . . .


Yawed cylinder schematic geometry and nomenclature. . . . . . . . . . .
Tandem cylinder configuration, definitions, and vortex shedding regimes.
Noise spectra for two 1 inch rods in a tandem configuration. . . . . . . .
Geometry for tandem cylinder configuration with yawed upstream cylinder.
Bombardier Dash-8 Q400 main landing gear assembly. . . . . . . . . . . .
Helical vortex system and slipstream tube generated by a propeller. . . .
Radial distribution of axial and tangential velocities behind a propeller. .
Far-field noise spectra of full-scale and 1/4-scale propellers. . . . . . . . .

5
7
8
9
11
12
14
15
16

3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8

Photograph of the Acoustic Wind Tunnel Facility. . . . . . . . . . . . . .


Anechoic wedge geometry. . . . . . . . . . . . . . . . . . . . . . . . . . .
Measurement grid for aerodynamic survey. . . . . . . . . . . . . . . . . .
Open jet velocity profiles of the Acoustic Wind Tunnel. . . . . . . . . . .
Power spectral density of wind tunnel jet longitudinal velocity fluctuations.
Turbulence intensity in the wind tunnel open jet vs. axial position. . . .
Turbulence intensity in the wind tunnel open jet vs. radial position. . . .
Background narrow-band SPL spectra of the acoustic wind tunnel. . . . .

18
19
20
22
23
23
24
24

4.1
4.2

. . . .
in the
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .

25

4.3
4.4
4.5
4.6
4.7
4.8

Image of landing gear model in wind tunnel test section. . . . . .


Schematic drawing of the complete landing gear model geometry
wind tunnel test section. . . . . . . . . . . . . . . . . . . . . . . .
Propeller wake velocity profiles for 3-blade, 15 7 propeller . . .
Propeller wake velocity profiles for 3-blade and 2-blade propellers.
Schematic drawing of far-field microphone test locations. . . . . .
Schematic drawing of model surface microphone locations. . . . .
Sample calibration curve for a surface microphone sensor. . . . . .
Instrumentation layout for data acquisition . . . . . . . . . . . . .

5.1
5.2
5.3
5.4
5.5

Far-field SPL spectra of static configurations. . . . . . . .


Far-field SPL spectra of motor configurations . . . . . . . .
Far-field SPL spectra of propeller configurations . . . . . .
Slipstream contraction for all test cases. . . . . . . . . . .
Strut unsteady surface pressure spectral levels, M = 0.04

.
.
.
.
.

38
40
41
42
45

vii

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

26
33
33
34
34
35
36

5.6
5.7
5.8
5.9
5.10

Strut unsteady surface pressure spectral levels, M = 0.07 .


Strut unsteady surface pressure spectral levels, M = 0.10 .
Wheel unsteady surface pressure spectral levels, M = 0.04
Wheel unsteady surface pressure spectral levels, M = 0.07
Wheel unsteady surface pressure spectral levels, M = 0.10

.
.
.
.
.

46
47
49
50
51

A.1 Schematic of the Acoustic Wind Tunnel Facility. . . . . . . . . . . . . . .

61

B.1 Calibration curves for all surface microphone sensors. . . . . . . . . . . .

63

viii

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

1 | Introduction
1.1

Aircraft Noise

Communities and regulatory bodies have long been concerned with the noise production of the aviation industry. In the United States, the Federal Aircraft Administration
(FAA) has taken the initiative to reduce aircraft noise through regulations and certifications in accordance with the Federal Aircraft Regulations Part 36 [12]. The approach is
three pronged, with measures taken to control noise sources, set operational restrictions,
and use effective land planning to reduce community noise exposure. Some of these actions have also been adopted by the International Civil Aviation Organization (ICAO).
Recent interest from the aviation industry, driven by these and other regulations from
governmental and international organizations, has motivated research into aircraft noise.
Since the 1970s, there has been a growing interest in aircraft noise studies. In both the
United States and the United Kingdom, early studies were aimed at reducing the noise of
aircraft engines. However, with the advent of high bypass-ratio turbofan engines, focus
was broadened to include airframe noise. Early phases of research into understanding and
predicting airframe noise used total aircraft methods. Aircraft parameters such as wing
area, aspect ratio, and flight velocity were used to approximate the overall noise signature
(sound level and frequency spectrum) [21, 14]. This method of prediction was superseded
by source-component methods which have the advantage of dividing the analysis to
individual components that may separate the noise generation mechanisms. The airframe
elements identified as sound sources include, but are not limited to, the wings, flaps, slats,
stabilizers, wheel-wells, and landing gear [22]. Combining the noise signatures from each
source, the overall airframe noise signature are determined. Although landing gear had
been identified as a primary contributor, early research regarding airframe noise was
directed toward wing components (i.e. slats and flaps) primarily due to the complex
geometries of conventional gear.
Early research concerning landing gear noise wrongly identified it as an exclusively low
frequency phenomenon [8]. This was corrected in the 1990s when wind tunnels capable of
1

Chapter 1. Introduction

testing full scale models became available. Dobrzynski and Buchholz [10] performed tests
concluding that early models lacked the detail required to account for higher frequency
noise apparent in actual landing gear. The tabulated noise levels peaked between 1
and 20 kHz. In addition, these tests attributed low frequency noise to the large gear
components (wheels, struts, etc.) and high frequency noise to the smaller elements (pins,
cables, nuts, etc.). Recent studies have investigated generic landing gear models such as
the four-wheel rudimentary landing gear model introduced by Spalart et al. [42, 34] and
the two-wheel LAGOON model used by Manoha et al. of Airbus [27].
Empirical approaches to the problems of predicting and modeling landing gear noise
remain dominant in industry. Past experimental studies, such as those by Guo [17] and
Dobrzynski et al. [9], have demonstrated that a landing gear noise signature depends on
local flow conditions that can be influenced by the landing gears location on an aircraft
(wing vs. fuselage) or upstream airframe components.

1.2

Objective

The current study aims to identify the aeroacoustic signature specific to a two-strut, twowheel landing gear when the components are in the unsteady slipstream of a propeller.
The motivation of this work stems from community noise concerns related to regional
turboprop aircraft such as the Bombardier Dash-8 Q400, shown in Figure 1.1. This is
the first documented aeroacoustic investigation of the effects of an upstream propeller
on the radiated noise of landing gear. Both far-field acoustic and surface pressure data
will be acquired simultaneously to identify pressure fluctuations which are manifested as
far-field noise.

Figure 1.1: Austrian Arrows Bombardier Dash 8 Q400 aircraft. Taken from [5].

Chapter 1. Introduction

1.3

Thesis Outline

The structure of this document is described below. Chapter 2 reviews some concepts
relevant to the experimental design including the expected flow patterns and noise emissions from both landing gear components and propellers. The work done to characterize
the experimental facility is presented in Chapter 3. This was necessary prior to the
aeroacoustic experiment due to many recent modifications to the wind tunnel. Both
aerodynamic and acoustic testing were performed to quantify the wind tunnel characteristics. Chapter 4 addresses the experimental set-up for far-field acoustic and surface
pressure measurements. The results of the landing gear study for various configurations and flow speeds are reported and discussed in the following chapter. Lastly, some
conclusions and recommendations for future investigations are presented in Chapter 6.

2 | Background
The proposed investigation requires a variety of topics to be reviewed. This section
begins with a brief discussion on the theoretical sources of aerodynamic sound. Canonical
flows of single and tandem cylinder arrangements are presented as a prelude to complex
landing gear geometries. The far-field acoustics of these cylinder flows is presented prior
to a discussion of aeroacoustic studies of landing gear in uniform inflow. Testing of
rudimentary and high-fidelity models are examined with key results addressed. Lastly, a
review of the flow structures of propellers and their expected noise signature is presented.

2.1

Aerodynamic Sources of Sound

Lighthills theory of aerodynamic noise generation [26] is widely considered the starting
point for modern aeroacoustics research. His reformulation of the Navier-Stokes equations
resulted in an inhomogeneous form of the wave equation separating acoustic propagation
processes from aerodynamic sources. Assuming a uniform medium at rest, the equations of the propagation of sound due to an externally applied fluctuating stresses using
Einstein tensor notation is
2 Tij
2
2 2

=
,
0
t2
xi xj

(2.1)

where is the fluid density, a0 is the speed of sound in the fluid, and Tij is the Lighthill
stress tensor:
Tij = ui uj + Pij a20 ij ,

(2.2)

where ij is the Kronecker delta. Thus, it can be observed that the sources of acoustic
radiation can be attributed to the three terms in Lighthill stress tensor, corresponding
to momentum flux (ui uj ), the stress tensor of a Stokesian fluid (Pij ), and nonlinear
noise generation mechanisms (a20 ij ). The latter two of these sources are negligible for
low Mach number flows [26, 15]. It can be shown that in this case, the sound produced
4

Chapter 2. Background

by Tij corresponds to a quadrupole field. Lighthills formulation was further developed


by Curle [7] to take into account the presence of solid boundaries. A notable result for
low Mach number flows is that dipole radiation dominates the sound emanating from
quadrupole sources. Ffowcs Williams and Hawkings [13] again extended this theory to
include objects moving in flows, showing that forward motion of a source influences the
radiation pattern with respect to an observer.

2.2

Bluff-body Flows

Landing gear can be decomposed into an assortment of non-streamlined (bluff) bodies.


An understanding of the acoustics of bluff-body flows is beneficial in determining noise
sources of complete landing gear. Flows around all bluff bodies have some similarities
including large regions of flow separation and unsteadiness. The separated flow region is
largely determined by the locations of boundary layer transition and separation, which are
heavily influenced by Reynolds number, body shape, viscosity, and free stream turbulence
levels. Unsteadiness is manifested as fluctuating pressures on the body surface and in
the resulting wake. If pressure fluctuations are periodic, they lead to an Aeolian tone at
the vortex shedding frequency.

2.2.1

Isolated Cylinders

Many components of landing gear can be


characterized as circular cylinders of dif(iii) U > U
ferent diameter, aspect ratio, and alignment. Thus investigating the flow around U

(iv) U < U
D
a circular cylinder provides a good start(i)
(ii)
ing point for landing gear studies. Flow
(iii) U > U
around circular cylinders has been extensively researched in the past, a comprehensive summary of which is provided by Figure 2.1: Regions of disturbed flow surrounding a circular cylinder. Adapted from
Zdravkovich [48, 49]. Surrounding a cylin- Zdravkovich [48].
der, four areas have been identified as areas where the local velocity, U , and pressure, p, vary from the freestream conditions
(U , p ). Figure 2.1 is a schematic representation of these regions. Region (i) is a narrow region of retarded flow, forming unsteady flow structures in the streamwise direction
directly ahead of the cylinder. Region (ii) marks the viscous boundary layer prior to

Chapter 2. Background

separation from the cylinder surface. The separation of the shear layer prescribes the
downstream boundary for region (iii). Region (iii) is an area of displaced and accelerated
flow, the extent of which is affected by the blockage of the cylinder. Lastly, region (iv) is
a wide downstream region of separated flow, identified as the wake.
Perhaps the most influential parameter in defining the state of disturbance-free cylinder flows is the Reynolds number based on the cylinder diameter, D:
ReD =

DU
,

(2.3)

where is the dynamic viscosity of the freestream fluid. The Reynolds number represents
a ratio of inertial to viscous forces. Each cylinder flow state has a subset of flow regimes
in which flow patterns are distinct. These are identified in Table 2.1. Generally, as
ReD increases, the point at which transition to turbulence occurs moves upstream. At
low Reynolds numbers, viscous forces are dominant and the flow around the cylinder is
laminar and attached to the cylinder surface. Inertial forces become more prevalent as
ReD is increased, leading to turbulence in the wake, the shear layer, and eventually the
boundary layer of the cylinder.
Table 2.1: Flow regimes of a disturbance-free flow around a circular cylinder. Adapted
from Zdravkovich [48] and Zawodny [47].
State and Regime
ReD Ranges
Laminar
No Separation
0 to 4-5
Closed Wake
4-5 to 30-48
Periodic Wake
30-48 to 180-200
Transition in Wake
Far-wake
180-200 to 220-250
Near-wake
220-250 to 350-400
Transition in Shear Layer
Lower
350-400 to 1k-2k
Intermediate
1k-2k to 20k-40k
Upper
20k-40k to 100k-200k
Transition in Boundary Layers
Precritical
100k-200k to 300k-340k
Single Bubble
300k-340k to 380k-400k
Two-bubble
380k-400k to 500k-1M
Supercritical
500k-1M to 3.5M-6M
Postcritical
3.5M-6M to (?)

Description
Fully-attached flow
Steady, symmetric, closed wake
Periodically oscillating shear layer
Laminar eddy transition
Irregular eddy transition
Transition wave development
Transition eddy formation
Eddies burst to turbulence
Shear layer transition
Asymmetric turbulent separation
Symmetric boundary layer separation
Disruption of separation bubbles
Transition advances toward stagnation line

Laminar to turbulent transition in the boundary layer, shear layer, and wake are all

Chapter 2. Background

sensitive to disturbances of the free-stream. These disturbances include but are not limited to freestream turbulence, surface roughness, transverse and streamwise oscillations,
and wall blockage. The resulting flow structures in each flow regime can be significantly
modified due to disturbances, typically initiating transition at lower Reynolds numbers.
A detailed review of each flow state and regime as well as influencing parameters can be
found in [48].
The Strouhal number based on cylinder diameter, StD , is a dimensionless quantity
used to characterize the flow around a cylinder and is defined as
StD =

fD
,
U

(2.4)

where f is the cylinder vortex shedding frequency. The shedding of vortices results
in unsteady pressure fluctuations, which are manifested in the acoustic far-field as an
Aeolian tone [48]. For 300 < ReD < 2 105 , Bearman [2] found the Strouhal number
remained approximately constant at a value of StD 0.2. For higher Reynolds numbers
(3 105 < ReD < 7 105 ), StD was found to shift in behaviour from a sharp tone to a
broader spectral peak, indicating the presence of turbulent flow.
A schematic depicting a yawed cylinder geometry and definitions is shown in Figure 2.2. In the
case of a yawed circular cylinder with yaw angle ,
the cross-section with respect to the flow becomes
elliptical and freestream velocity, U , has compo
nents both along the span, Ur , and normal to the
cylinder axis, Un . Sears [37] simplified the NavierStokes equations for an infinite cylinder to show
that the projected flow on the normal plane is de
L
scribed by the same equations as the unyawed case,
and is independent of the spanwise flow. This separation of variables is known as the Independence
U
Principle. It should be noted that the assumption

of a laminar boundary layer and the omission of end


Ur
Un
D
effects (which are present for cylinders with finite
Figure 2.2:
Yawed cylinder aspect ratio, L/D) make this an idealization. Howschematic geometry and nomencla- ever, it has been proven to be accurate for numerous
ture.
cases. It follows that an equivalent Reynolds number, Ren , and Strouhal number, Stn , based on the

Chapter 2. Background
normal component of the velocity can be defined by
Ren =

DU cos

and

Stn =

fD
.
U cos

(2.5)

The validity of the Independence Principle is dependent on the flow state as defined
in Table 2.1. The equivalent Reynolds and Strouhal numbers for vortex shedding were
shown to be valid in the Laminar state for < 45 by Hanson [20]. Shirakashi et al. [38]
found that in the Transition in Shear Layer state, the Independence Principle holds for
. 50 . For each of these cases, the observed shedding frequency occurred at St 0.20.
In his review of flow around circular cylinders, Zdravkovich reports that the Independence
Principle cannot correlate the onset of the Transition in Boundary Layer state [49].

2.2.2

Tandem Cylinders

Parallel Tandem Cylinders


Analyzing the flow around landing gear geometries requires some understanding of

flow interactions between bluff bodies. A


canonical case for this flow interaction is
the flow around tandem cylinders, with
one cylinder behind the other, as defined
in Figure 2.3. The separation between
the cylinders is typically defined using the
spacing ratio, S/D, equal to the ratio
of distance between cylinders to cylinder
diameter. In the case of non-identical
cylinders, the spacing ratio is typically
defined using the upstream cylinder diameter. A review of the interactions of
tandem cylinders based on the separation
is presented in Table 2.2, as described
by Zdravkovich [49]. The spacing ratio Figure 2.3: Tandem cylinder configuration,
regimes are affected by the flow Reynolds definitions, and vortex shedding regimes.
number.

(a)

(b)

(c)

The effect of a tandem arrangement on Strouhal number has been examined in the
Transition in Shear Layer state and results in three distinct ranges as indicated in Ta-

Chapter 2. Background
Table 2.2: Description of tandem cylinder spacing effects.
Spacing Ratio, S/D
1 < S/D
1.1-1.3 < S/D
3.0

< S/D

3.8

< S/D
S/D

Description
Forward cylinder shear layers
< 1.1-1.3
envelope downstream cylinder
Reattachment of forward shear
< 3.5-3.8
layers on downstream cylinder
Intermittent vortex shedding
< 4.0
behind forward cylinder
Synchronized vortex shedding
< 5-6
of both cylinders
Independent (uncoupled) vortex
> 5-6
shedding

St
0.24-0.28

0.12-0.15

0.17-0.19

ble 2.2. The first flow regime (a) involves two closely spaced cylinders producing a
narrow wake, resulting in a high St (0.24-0.28). For larger cylinder spacings (b), the
upstream cylinder shear layer reattaches to the downstream cylinder, yielding a low St
behind the downstream cylinder (0.12-0.15). Finally for large enough spacing ratios (c),
both cylinders behave in vortex street regime exhibiting a slowly rising Strouhal number
(0.17-0.19).
Essentially, the differences in these
S
shedding patterns can be attributed to
the difference in the local flow state of
each cylinder. The upstream cylinder has
disturbance-free oncoming flow and thus
behaves as an isolated cylinder. The downstream cylinder contends with a velocity
deficit due to the wake and the turbulent
shear layer of the upstream cylinder. As
the spacing between the cylinders is increased, the effect of the tandem arrange- Figure 2.4: Noise spectra for two 1 inch rods
ment decreases.
in a tandem configuration (M = 0.13).
In the Transition in Boundary Layer Dashed line represents measurements taken
state, any changes are only relevant to with turbulence generated from an upstream
grid. Figure adapted from Hutcheson and
the upstream cylinder, as the downstream Brooks [23].
cylinder is fully in its turbulent wake. Okajima [30] examined two tandem cylinder spacing ratios for ReD < 4.5 105 . His results
show approximately constant values of Strouhal number of StD = 0.14 for S/D = 3 and
StD = 0.19 for S/D = 5.

10

Chapter 2. Background

The acoustic emissions of tandem parallel cylinder correspond to the vortex shedding frequencies and thus the corresponding Strouhal numbers. Some studies have been
conducted in order to quantify the far-field aeroacoustic signature of tandem cylinder
configurations. Hutcheson and Brooks [23] examined single and multiple rod configurations for 3.8 105 < ReD < 105 . Considering the tandem configuration, spacing ratios
of 1,2,3, and 4.5 were investigated for D = 1 inch. The acoustic spectra for each of
these spacing ratios for smooth cylinders at a Mach number, M = U /a0 , of 0.13 is
presented in Figure 2.4. In this figure, acoustic emissions are shown in terms of sound
pressure levels, SPL, defined as
SPL = 10 log10

p2rms
p2ref


,

(2.6)

where prms is the root mean square sound pressure being measured and pref is a reference pressure (typically 20 Pa). The shedding frequency for a smooth 1 inch diameter
cylinder at this Mach number was determined in these experiments to be approximately
300 Hz (StD = 0.19).
For the tandem cylinder case with the minimum spacing, a single tone was observed
at a frequency nearly double the expected shedding frequency of an isolated rod. For
spacing ratios of 2 and 3, the observed primary tone is lower than the expected single
cylinder frequency by approximately 20%. Each of these cases emit additional tones at
the second and third harmonics of the dominant tone frequency. For the largest spacing
(S/D = 4.5), a single tone approaching the single cylinder frequency is observed. The
corresponding Strouhal numbers for the observed tones in Figure 2.4, correspond well to
the regimes defined by Zdravkovich [49].

Yawed Tandem Cylinders


Wilkins and Hall [46] have investigated tandem yawed cylinders in the configuration
shown in Figure 2.5. They observed that this arrangement of cylinders resulted in a
broadened peak at the shedding frequency for S > 4.5D. It is suggested that the turbulence produced by the upstream cylinder disrupts organized vortex shedding leading
to highly three-dimensional flow. This three-dimensionality also resulted in decreasing
the regularity of vortex shedding from the rear cylinder. The study examined yaw angles = 10 , 20 , and 30 , finding that the yaw angle was of greater influence to vortex
shedding than the cylinder spacing.

11

Chapter 2. Background

Sn

Ur

Un
D

Figure 2.5: Geometry for tandem cylinder configuration with yawed upstream cylinder.

2.3

Landing Gear Studies

The use of cylinder studies in engineering applications of landing gear noise has its
limitations due to the inherently three-dimensional nature of a landing gear flow field.
Wind tunnel tests on landing gear geometries of various scales and fidelities have been
conducted to examine the aerodynamic and aeroacoustic behaviour of representative
models. A depiction of a typical landing gear and its components are shown in Figure 2.6.
Wind tunnel models may include any combination of struts, torque links, bogies, wheels,
and finer details such as hydraulic hoses and brakes. Flow measurement techniques such
as hot-wire anemometry and Particle Image Velocimetry (PIV) are used to find regions of
fluctuating velocity, indicative of noise sources. Unsteady pressure transducers located
on model surfaces have also aided in the detection of sound production regions. The
radiated noise is measured using microphones positioned around the model. Microphones
located at different angles allow for acoustic directivity patterns to be determined. Phased
microphone arrays are commonly used to localize noise sources through a processes known
as beamforming. Beamforming results in a series of spatial maps depicting the sound
levels for given frequency ranges. The data from these studies has been used to validate
computational aeroacoustic codes as well as provide references for noise prediction tools
that can be used in aircraft design. These studies are briefly summarized below.

Chapter 2. Background

12

Early studies investigating airframe


noise performed in the 1970s indicated
Main gear pin
that landing gear are important contribuAuxiliary
actuator
tors to overall noise levels. The first study
Stabilizer
brace
Main actuator
to isolate landing gear for noise measureU
ments was conducted by Heller and DoDrag strut
Fairings
brzynski [22], using a rudimentary twowheel landing gear consisting of only a
Main (Shock)
strut, axle, and wheels. Landing gear noise
strut
Proximity
was identified as a predominantly low fresensors
quency phenomenon. This was later corrected after Dobrzynski and Buchholzs
Figure 2.6: Bombardier Dash 8 Q400 main full-scale testing of Airbus A320 and A340
landing gear assembly. Figure adapted from landing gear in 1997 [10], which showed a
www.smartcockpit.com [6].
broadband noise signature spanning from
200 Hz - 10 kHz.
Despite the fact that it is necessary to include the fine geometric details of a landing
gear to acquire a thorough description of its resulting sound field, many conclusions can
be drawn by investigating simplified geometries.

2.3.1

Simplified Geometries

Testing of simplified landing gear geometries is commonplace due to the difficulties associated with replicating fine details down to model scale sizes. Some important results
have been reported from testing of these geometries. In the tests of Dobrzynski and
Buchholz [10] as well as those of Guo et al. [18], it was shown that the addition of details
onto a rudimentary model manifests itself as a broadband increase in sound levels with
the greatest gains in the high frequency range. Flow structures developed due to wake
body interactions between closely spaced components have been identified as important
airframe noise contributors. These regions include the gaps between wheels [25, 32] and
the junctions of the strut and axle [32].

2.3.2

High Fidelity Geometries

The full-scale landing gear tests of Dobrzynski and Buchholz were the first to demonstrate
that landing gear noise is fundamentally broadband in nature [10]. These noise levels
generally had a sixth power scaling relative to flow velocity. In the middle frequencies

Chapter 2. Background

13

of the two wheel gear emissions, the behaviour differed which was attributed to velocityindependent noise generation between the wheels. In 2006, Guo et al. [18] tested a
high-fidelity model-scale Boeing 737 main landing gear, identifying dominant sources
downstream of the main strut and front wheels. Small components were removed for some
tests showing broadband decrease in acoustic power (most noticeably at high frequencies).
Turbulence generated from the landing gear assembly was proposed as a primary
source of high frequency noise as seventh-power velocity scaling was the best fit in that
region of the acoustic spectrum. Other high-fidelity tests on Boeing landing gear by
Ravetta et al. [33] showed that the removal of the drag brace produced a reduction in
overall sound levels. Ringshia et al. [35] revisited this work and showed high turbulence in
the wake region of the drag brace and that the vortex formed in the strut wake impinges
on the front face of the rear drag brace which may be a significant noise source.
Experimental investigations of models of various fidelities by Guo et al. [19] showed
that the noise spectrum of landing gear can be decomposed into three frequency ranges.
Each range represents contributions from the wheels (low frequencies), struts (mid frequencies), and small details such as cut-outs and hoses (high frequencies). In a prediction
model based on the Strouhal numbers of major components, Guo et al. were able to introduce a complexity factor to account for the effects of small details, with satisfactory
results.

2.4

Propeller Flows

The current work aims to investigate the effects of an upstream propeller on the flow
over a landing gear. Since Guo et al. [17] showed that the local flow conditions affect the
radiated noise, a review of the flow structures and acoustics related to propeller flows are
discussed below.

2.4.1

Flow Structures

A propeller consists of equally spaced blades rotated by the torque of an engine. The
direct result of the rotating blades is a pressure increase behind the propeller and a
decrease of pressure ahead of it. This pressure difference is a measure of propeller thrust.
In the wake region aft of an actuating propeller, vortices are shed downstream in almost
helical paths and represent the propeller slipstream. Vortices are shed both by the trailing
edge of each blade as well as the blade tips.
In his work regarding the propeller-wing aerodynamic interference, Veldhuis provides

14

Chapter 2. Background

an excellent summary of the characteristics of propeller slipstream flow [44]. Results


concerning the axial and swirling velocity profiles as well as slipstream contraction are
presented below. Figure 2.7 is a schematic of the major flow structures associated with
the propeller slipstream.

blade vortex sheet


rolled up
vortex system

root vortex

slipstream tube

tip vortex

Figure 2.7: Helical vortex system and slipstream tube generated by a propeller. Figure
adapted from Veldhuis [44].
A propeller is functionally used to provide thrust in most applications and thus one
would expect the axial velocity behind a propeller to be greater than the freestream
value. Just behind the propeller, the axial flow velocity is equal to U + Ua , the sum of
freestream velocity and the propeller-induced velocity. Far downstream of the propeller,
the axial flow velocity in the propeller slipstream tube reaches U + 2Ua .
A typical velocity profile aft of a propeller is shown in Figure 2.8 for both the axial
and tangential velocities, Ua and Ut , respectively. The induced axial velocity is shown
to reach a maximum near 34 R, where R is the propeller radius, typically at a location
of maximum blade loading. The tangential (or swirl) velocity distribution is dependent
on the loading conditions of the blades and is susceptible to change for different advance
ratios [44]. The swirl angle of a blade vortex sheet, sw , can be computed using
1

sw = tan

Ut
Ua + U


.

(2.7)

Since the axial velocity is a function of the streamwise position and the tangential velocity
is not, the swirl angle changes along the length of the slipstream. Thus, the incident flow

15

Chapter 2. Background

angle on any downstream components should be calculated based on their location behind
a propeller.

Ua

Ut

0.05

Ua / U , U t / U

Figure 2.8: Typical radial distribution of axial velocity, Ua , and tangential velocity, Ut ,
directly behind a propeller. Figure adapted from Veldhuis [44].

The slipstream of a propeller contracts to maintain a constant mass flow as velocity


is increased through the propeller. The diameter of the slipstream tube downstream of
the propeller should be computed to identify which regions are affected by the propeller.
Veldhuis [44] presents an approximation of the contraction ratio,
v
1+a

Rs u
u

,
=t
R
1+a
1 + R2x+x2

(2.8)

where Rs is the radius of the slipstream at location x, and a


is the axial inflow factor
representing the velocity at the propeller plane.
A dimensionless quantity frequently used to compare propeller operational settings
and performance is the advance ratio, J, defined as
J=

U
,
nDp

(2.9)

where n is the rotational rate of the propeller and Dp is the propeller diameter. This
quantity indicates the ratio between the distance the propeller moves forward per revolution and Dp . The advance ratio is a measure of effective pitch and may be used to
determine the incoming angle of the fluid relative to a propeller blade.

16

Chapter 2. Background

2.4.2

Acoustics

Increased interest in the aviation industry with respect to turboprop aircraft and other
open rotor technologies, has resulted in research concerning propeller noise reduction.
Trebble et al. [43] have performed some aeroacoustic studies of both model- and fullscale propellers. The authors report both near- and far-field acoustic measurements for
various operating conditions, modifying rotational rates, blade-angle settings, and wind
tunnel flow speeds. In order to compare the full-scale and model-scale datasets, the
propeller blade tip Mach numbers, Mtip , were matched for the two cases.
q
2
2
+ U
Utip

vtip
=
(2.10)
a
a
Presented in Figure 2.9 is a noise spectra comparison between full-scale and 1/4scale models normalized to 1600 RPM = 26.7 Hz. This corresponds to a blade passing
frequency of 106.7 Hz (four-bladed propeller), which can be observed as the location of
the first, and largest tonal peak. Good correlation exists between the observed value of
sound pressure levels of the tonal harmonics in the spectra for both propellers. Though
a discrepancy exists in the broadband levels, this is attributed to contamination from
the wind tunnel background noise. The number of measurable harmonic tones was found
to increase with the propeller speed. The sound pressure level of the fundamental tone
also increased with propeller rotational speed and propeller blade-setting angle (increased
thrust and torque). When the propellers were lightly loaded, they exhibited a rise in the
high frequency content of the broadband noise levels.
Mtip =

120
dB
SPL

Full-scale

scale MEASURED CORRECTED

SPL

100

m=1 116.8
m=2 106.4
m=3 98.9
m=4 91.5

115.0
106.1
100.9
95.5

118.9
103.9
100.3
96.9

80
Full-scale
60

-scale
0

0.2

0.4
0.6
0.8
1.0
1.2
1.4
Equivalent full-scale frequency (kHz)

1.6

Figure 2.9: Far-field noise spectra comparison between full-scale and 1/4-scale propeller
models normalized to 1600 RPM. Figure adapted from Trebble et al. [43].

3 | Wind Tunnel Characterization


The following sections detail the work done to characterize the flow and acoustic properties of the acoustic wind tunnel at the University of Toronto Institute for Aerospace
Studies (UTIAS). These tests were completed in order to quantify the performance of the
wind tunnel, as multiple modifications have taken place since its previous experimental
use. The wind tunnel was constructed in 1975 and has since been used for various studies ranging from investigations of Tollmien-Schlichting instabilities to marine propeller
performance. Since 2011, changes include:
repair of the motor windings,
adjustments of the fan blades,
installation of a variable frequency drive (VFD) to adjust the motor rotational rate,
installation of anechoic wedges in the test chamber,
application of acoustic treatments to the tunnel contraction and nozzle, and
construction of a downstream silencer.

3.1

Facility Overview

The experiments presented in the current work were conducted at UTIAS in the Acoustic
Wind Tunnel Facility, a photograph of which is shown in Figure 3.1. This facility features
an open-circuit, open jet, suction-type wind tunnel. The nozzle diameter, DN , of the
open jet is 0.70 m, and the nozzle-to-collector distance is 2.78 m. The wind tunnel can
be operated at a maximum flow speed of U 60 m/s.
The wind tunnel is powered by a 150 h.p. motor driving a nine-bladed axial fan
with a maximum RPM of 1800. Air is drawn from a geodesic dome 55 m in diameter
to mitigate atmospheric effects on the inlet conditions. The flow is then conditioned
through an aluminum hexagonal honeycomb of 1.59 mm cell width by 50 mm deep, a
17

Chapter 3. Wind Tunnel Characterization

18

Figure 3.1: Photograph of the Acoustic Wind Tunnel Facility (right) and the adjacent
geodesic dome (left).

120 mesh (120 strands per inch) screen and four 54 mesh screens for turbulence reduction
as described by Ball [1]. The conditioning devices lead into a 10:1 contraction which is
immediately followed by a square to circular transition and the nozzle to the open jet
test section. Flow continues from the jet collector into the diffuser toward the motor
and fan. On either side of the motor and fan, silencers attenuate noise from propagating
into the tunnel test section where measurements are taken or outside near the facility.
A plenum with catcher screens protects the fan blades from debris flowing through the
tunnel. Figure A.1 in Appendix A shows a schematic of the entire Acoustic Wind Tunnel
Facility.
The issues associated with wall reflections in closed-section wind tunnels for aeroacoustic measurements can be resolved by placing the test section in an anechoic environment. In this case, open jet wind tunnels are preferable to closed section facilities since an
acoustically treated room can be easily built around an open test section. The open jet
of the Acoustic Wind Tunnel Facility is located within a chamber treated with fibreglass
wedges. A schematic of the stepped wedge geometry used in the anechoic chamber is
shown in Figure 3.2. Each wedge block is rotated 90 to its adjacent wedges. The wedges
can effectively absorb wavelengths greater than four times the height of the wedge. Assuming the speed of sound in air, a0 , is constant (343 m/s at standard temperature and
pressure), the frequency, f , of a sinusoid and its wavelength, , are related by
a0 = f ,

(3.1)

and given a quarter-wavelength equal to the height of the wedge, 0.610 m (24 inches),
the resulting theoretical cutoff frequency is equal to approximately 140 Hz. Thus it is
advised that experiments conducted in the UTIAS acoustic wind tunnel be designed to
examine frequencies greater than 140 Hz since investigations of lower frequencies will
require reflected sound waves to be taken into account.

19

Chapter 3. Wind Tunnel Characterization

23

30

46

61

5.1
61

122

Figure 3.2: Schematic of the anechoic wedge geometry. Dimensions are in centimeters.

3.2
3.2.1

Aerodynamic Survey
Experimental Method

A survey was conducted in the open jet of the wind tunnel test section to examine
the spread of the jet shear layer, the shape of the jet potential core (region with nearly
uniform flow velocity), and the turbulence levels within the jet. Two linear traverses were
placed under the test section allowing for movement in the x and r directions as defined
in Figure 3.3. The traverses were operated using MATLAB through a Motion Group 4channel motion controller via a serial cable. A pitot-static tube and single hot-wire probe
were near the jet centerline and were traversed to the grid locations shown in Figure 3.3.
Pressure measurements were acquired using an MKS Instruments Type 120AD Baratron
100 Torr differential pressure transducer, which has an accuracy of 0.12% of the reading.
Pressure measurements from a pitot-static tube correspond to the dynamic pressure, q
and can be related to the velocity, U , through the incompressible form of Bernoullis
equation:
1 2
= q
p0 p = U
2

(3.2)

The hot-wire used was a 5 m copper-plated tungsten wire with a sensing length
of 1 mm. Data were taken through a DISA (Dantec) 55M10 Constant Temperature
Anemometer (CTA) operating at an overheat ratio of 0.6. Prior to each test run, the
measurement hot-wire was calibrated in-situ against a pitot-static tube. This was done

20

Chapter 3. Wind Tunnel Characterization

placing the pitot-static tube and hot-wire probes about the tunnel centreline, 1 cm apart
and equating measurements from the CTA with the measured dynamic pressures for
a range of 26 flow speeds between 5 m/s and 55 m/s. A J-type thermocouple was
mounted in the tunnel collector to provide a temperature correction to hot-wire data.
Temperature fluctuations for the duration of each test were typically 2 C. Hot-wire
data and pressure transducer data were acquired at a rate of 4 kHz for 30 seconds using
a National Instruments PCI-6361 A/D data acquisition system at each measurement
location. The overall uncertainty in the mean velocity profiles was estimated to be less
than 1%.
1.0 typ.

4.0 typ.
x
48.0
r

80.0

10.0 typ.
120.0
278.1

Figure 3.3: Measurement grid for aerodynamic survey. Dimensions are in centimeters.

3.2.2

Mean Flow Measurements and Turbulence Levels

The mean flow velocity profiles at the measured downstream stations for Mach numbers
of 0.10 (U = 33 m/s) and 0.15 (U = 51 m/s) are presented in Figures 3.4a and 3.4b,
respectively. For M = 0.10, the accelerated flow near the lip of the nozzle presents
an uneven velocity profile until approximately x/DN = 1. However, the flow remains
uniform over the majority of the jet width beginning at x/DN 0.3. A similar trend
is observed for the survey conducted at M = 0.15. The jet shear layer appears to
grow identically for both test cases, reaching the extent of the measurement grid at
x/DN = 0.71.
The power spectral density of turbulent fluctuations along the jet axis for M = 0.10
and M = 0.15 have been estimated from hot-wire data and are presented in Figure 3.5.
Both Figures 3.5a and 3.5b show a peak in the spectra for x/DN > 0.99 occuring at
a Strouhal number based on DN between 0.7 and 0.8. This is identical to the peak in
the turbulent power spectral density shown by Ball for this facility in 1983. Michel and
Froebel [29] identify a broad hump at St = 1 as a perturbation inherent to free jets,

Chapter 3. Wind Tunnel Characterization

21

where turbulent eddies in the shear layer create near-field pressure gradients which cause
fluctuations in the flow. This spectral peak is shown to increase with distance from the
jet nozzle.
Turbulence intensity, defined as
T u = u0 /U

(3.3)

where u0 is the root mean square of the velocity fluctuations, was determined for M =
0.10 and 0.15. The power spectral density of the turbulent fluctuations were integrated
above 10 Hz to compute u0 . Turbulence intensity along the jet axis is presented in Figure
3.6. In the potential core, the turbulence intensity remains under 0.5% for x/DN less
than 1.4. The radial variations in turbulence intensity for M = 0.10 are presented in
Figure 3.7. Though irregular variations are evident near the nozzle, a consistent profile
is achieved by x/DN = 0.57. By x/DN = 0.99 the region of T u 0.1% is restricted to
0.6DN .

3.3

Background Acoustic Levels

A Bruel & Kjaer 4192 free field condenser microphone was placed 0.7 m downstream of
the nozzle and 2.1 m away from the jet centreline to measure background noise levels
at different jet exit velocities. The background sound pressure level (SPL) narrowband
spectra for a range of jet exit velocities between M = 0.04 and 0.16 are presented in
Figure 3.8. The measurements were taken at a sampling rate of 32 kHz for 180 seconds
using a National Instrument PCIe-6259 A/D data acquisition card. Flow velocities were
determined using a pitot-static tube located in the jet collector. The broad peaks occuring
at the low frequencies of the spectra can be attributed to jet shear layer impinging on
the wind tunnel collector and are similarly located to the peaks found in the turbulence
spectra. Strouhal scaling of the sound pressure level data results in a collapse of the
high frequency data and a shift of the low frequency peak to lower Strouhal numbers as
velocity is increased. Excessive background noise due to recirculating flows induced by
the jet may be due to the omission of wind screens on far-field microphones [41]. Between
100 Hz and 1 kHz the broadband background noise of this acoustic wind tunnel drops at
a rate of 20 dB per decade. This result is comparable to those presented by Mathew at
the University of Florida Aeroacoustic Flow Facility for speeds between 18 and 46 m/s
[28] .

r/DN

-0.5

-0.25

0.25

0.5

20

-0.5

-0.25

0.25

0.5

40

60

40

20

40

0.14

20

60

0.14

40

20

40

0.28

20

60

0.28

40

20

40

0.43

20

0.43

60

40

20

40

0.57

20

0.57

60

40

20

40

veloc ity, m/s

60

20

40

60

veloc ity, m/s

0.85

(b) M = 0.15

40

0.71

20

40

0.85

(a) M = 0.10

20

0.71

20

40

0.99

20

0.99

60

40

20

40

1.14

20

1.14

60

40

20

40

1.28

20

1.28

60

40

20

40

1.42

20

1.42

60

40

20

40

1.56

20

1.56

60

40

Figure 3.4: Open jet velocity profiles of the Acoustic Wind Tunnel for Mach numbers of (a) 0.10 and (b) 0.15.

20

x/DN = 0.00

r/DN

x/DN = 0.00

20

40

1.70

20

1.70

60

40

Chapter 3. Wind Tunnel Characterization


22

23

Chapter 3. Wind Tunnel Characterization

Frequency, Hz

10

10

-3

10 1

10 2

10 3

10

10 1

-2

10 2

10 3

x/D N
x/D N
x/D N
x/D N
x/D N
x/D N
x/D N

10 -3
x/D N
x/D N
x/D N
x/D N
x/D N
x/D N
x/D N

10 -4

= 0.00
= 0.28
= 0.57
= 0.85
= 1.13
= 1.42
= 1.70

Power Spectral Density

Power Spectral Density

Frequency, Hz
-2

10 -5

10 -4

10

= 0.00
= 0.28
= 0.57
= 0.85
= 1.13
= 1.42
= 1.70

-5

10 -6

10 -6

10 -7
10

-7

10

-1

10

10

10

-1

10

10

St = fDN/U

St = fDN/U

(a) M = 0.10

(b) M = 0.15

Figure 3.5: Power spectral density of wind tunnel jet longitudinal velocity fluctuations,
measured on-axis, for (a) M = 0.10 and (b) M = 0.15.

0.02
0.015

M = 0.10
M = 0.15

0.01

Tu = u/U

0.005

0.2

0.4

0.6

0.8

1.2

1.4

x/DN

Figure 3.6: Turbulence intensity in the wind tunnel open jet vs. axial position, measured
on-axis, for M = 0.10 and M = 0.15.

24

Chapter 3. Wind Tunnel Characterization

10

-1

Tu = u/U

x/D N
x/D N
x/D N
x/D N
x/D N

= 0.00
= 0.28
= 0.57
= 0.99
= 1.42

10 -2

10

-3

-0.6 -0.5 -0.4 -0.3 -0.2 -0.1

0.1 0.2 0.3 0.4 0.5 0.6

r/DN

Figure 3.7: Turbulence intensity in the wind tunnel open jet vs. radial position for
M = 0.10.

120

120
110

90

100

SPL, dB re 20 Pa

SPL, dB re 20 Pa

100

80
70
60

90
80
70
60

50

50

40

40

30
0
10

10

10

Frequency, Hz

(a)

10

M = 0.04
M = 0.07
M = 0.10
M = 0.13
M = 0.16

110

M = 0.04
M = 0.07
M = 0.10
M = 0.13
M = 0.16

10

30
-2
10

10

-1

10

10

St = fDN/U

(b)

Figure 3.8: Background narrow-band sound pressure level spectra of the acoustic wind
tunnel plotted against (a) frequency in hertz and (b) Strouhal number based on nozzle
diameter.

4 | Experimental Method
4.1

Overview

The purpose of the current work is to determine the baseline aeroacoustic behaviour for
a configuration involving a model landing gear in the unsteady wake of a propeller. The
landing gear geometry considered includes an inclined drag strut ahead of the main strut
with wheels. Of primary interest to this study are the measurements of the far-field
acoustics and the unsteady surface pressures. When acquired simultaneously, this data
can aid in deducing fluctuations which result in far-field sound emissions from the model.
Coherence and cross-correlations between the far-field and surface microphones can provide insight into noise generation as shown by Siddon [39]. Experiments attempting to
identify the contributions of various geometries and configurations to the surface and
far-field (acoustic) pressures are described below.

4.2

Model Description

An image of the tested model is shown in


Figure 4.1. The model geometry is based
on the main landing gear of the Bombardier Dash-8 Q400, a regional turboprop
aircraft, at a scale of 1:10.8. This scaling
ensures that an upstream model propeller
and the landing gear can be located entirely within the jet potential core of the
wind tunnel. The model consists of an
electric motor driving a 15 inches (38.1 cm) Figure 4.1: Image of landing gear model in
diameter propeller, a tubular nacelle, a wind tunnel test section.
main strut, a drag strut, an axle, and two
25

26

Chapter 4. Experimental Method


27.0

42.0

70.0
38.1

28

8.0

1.0

1.4

6.0 28.7
9.0

7.8
1.8

Figure 4.2: Schematic drawing of the complete landing gear model geometry with an
upstream propeller in the wind tunnel test section. The propeller is rotated 45 between
images for clarity. A two-bladed propeller was used for tests at M = 0.10, whereas a
three-bladed propeller was used for M = 0.04 and 0.07. Dimensions are in centimeters
wheels. The landing gear components are primarily constructed of aluminum. The tread
of each wheel and a removable strip on the main strut were made of ABS plastic to allow
for the complicated cut-outs that accommodate the surface sensors. Although small-scale
features such as hoses, cables, and cut-outs are not replicated, it is possible to predict
the additional noise of such small features by introducing a complexity factor in future
prediction models [16]. A schematic drawing of the landing gear geometry with respect
to the open jet nozzle is shown in Figure 4.2.
A Power 46 Brushless Outrunner Motor was used to turn the propeller along using a
60-Amp Pro Switch-Mode BEC Brushless Electronic Speed Controller, both from E-flite.
Power for the motor was provided by a TRC Electronics SP-750-12 (12V) single output
power supply. Throttle was controlled through USB communication with an Arduino
microcontroller. An optical tachometer was constructed using an phototransistor switch
(OPB606A) to count propeller revolutions and was placed within the nacelle, adjacent to
the motor. The back of the propeller hub was covered completely with black tape save for
one reflective strip, such that the phototransistor switch would provide a signal once per
revolution. An Eagle Tree Systems eLogger V4 was also used to monitor power consumption of the system, throttle levels, temperature, and motor RPM (through a brushless
motor sensor). The constructed tachometer and the brushless motor RPM sensor were
found to be within 50 RPM for all tests. This corresponds to a maximum relative error
of 2% for the lowest RPM case and an absolute uncertainty of 0.83 Hz. Since the averaging method of the eLogger was unknown, the RPM of the optical tachometer is used
as the reference value and verified against the provided eLogger data recorder software
to ensure a consistent reading.
Table 4.1 summarizes the model geometries, configuration names, and flow speeds for

27

Chapter 4. Experimental Method

all test cases. Configuration elements were incrementally added to the model assembly
to give additional information regarding the contributions of each isolated component to
the total sound emissions.
The motor speed in each case was adjusted to achieve maximum thrust while maintaining motor current at safe levels. Two propellers of diameter Dp = 38.1 cm, were
used for the tests described in this section. At M = 0.04 and 0.07, a Master Airscrew
3-blade 15 7 propeller was used in order to achieve a blade passing frequency (BPF)
of the propeller exceeding the cutoff frequency of the anechoic chamber (140 Hz). At
M = 0.10, the 3-blade propeller was unable to produce a net positive thrust, so it was
replaced with a Master Airscrew 2-blade 15 8 propeller. At this wind tunnel speed,
the 2-blade propeller was able to operate at a suitable RPM such that the BPF exceed
the chamber frequency cut-off limit. A radial survey of the slipstream for each propeller
operating condition was conducted in the wind tunnel using a pitot-static tube located
at a position 0.25Dp downstream of the actuation plane as was done by Beveridge [4].
An MKS Type 120AD Baratron differential pressure transducer was used to determine
the dynamic pressure at each survey location and thus the velocity. Figure 4.3 shows the
time-averaged velocity profile behind the three-blade propeller at the maximum operating condition for all test speeds. The values of the advance ratio, J, for the propeller
operating at each freestream velocity are marked in the legend of Figure 4.3. Thrust
values, T , were calculated using
Z

T = 2


Ua U Ua2 dr,

(4.1)

r=r0

which is a result from the momentum equation. Integration was done using the trapezoidal method and each thrust coefficient, CT , was computed using

CT =

T
.
n2 Dp4

(4.2)

The computed thrust coefficients for the three-blade propeller at M = 0.04, 0.07, and
0.10 are 0.049, 0.013, and -0.007, respectively. The negative value is an indication that
the generated thrust could not overcome the drag of the propeller blades in this case.
The two-blade was also tested at M = 0.10 and was shown to have a thrust coefficient
of 0.013 operating at maximum throttle. Figure 4.4 compares the velocity profiles of the
two-blade and three-blade propellers for M = 0.10.

28

Chapter 4. Experimental Method

Table 4.1: Experimental test matrix for acoustic measurement tests. *All tests conducted
at M = 0.10 used a 2-blade propeller.
Config.
Name

Speed
(M )

Main Strut Drag Rotating


& Wheels Strut
Motor

S-1

0.04, 0.07, 0.10*

S-2

0.04, 0.07, 0.10*

S-3

0.04, 0.07, 0.10*

M-1

0.04, 0.07, 0.10*

M-2

0.04, 0.07, 0.10*

M-3

0.04, 0.07, 0.10*

P-1

0.04, 0.07, 0.10*

P-2

0.04, 0.07, 0.10*

P-3

0.04, 0.07, 0.10*

Propeller

X
X
X
X

X
X

Schematic

Chapter 4. Experimental Method

4.3
4.3.1

29

Instrumentation
Far-field Microphones

Far-field acoustic spectra were acquired using two Bruel & Kjaer 1/2 4192 free-field
condenser microphones, mounted on a linear traverse in the anechoic chamber. Farfield measurements were taken along the wind tunnel centerline at locations F1 to F5,
and also 15 above the model nacelle at locations G1 to G5, as indicated in Figure 4.5.
The farthest upstream measurement locations possible due to space restrictions were
positions F1 and G1, which are 12 cm behind the propeller plane. Position F3 is collinear
with the model main strut when it is installed. To be considered in the acoustic farfield, measurements should be taken at least one wavelength, , from the source. For
these experiments, the far-field acoustic measurement plane is situated 2.10 m from the
tunnel centreline. Assuming the speed of sound, a = 343 m/s, this results in a minimum
frequency of 163 Hz. Any frequencies below this value cannot be considered to be in the
acoustic far-field.
Pistonphone Calibration
A Bruel & Kjaer Type 4428 pistonphone was used to calibrate the far-field microphones.
The pistonphone produces a 250 Hz tone at a known amplitude that is corrected to
the local atmospheric pressure conditions. Each far-field microphone was separately
calibrated with the pistonphone and their sensitivities recorded. The sensitivities for
the microphones measuring at locations F1 - F5 and G1 - G5 were determined to be
10.63 mV/Pa and 12.02 mV/Pa, respectively.

4.3.2

Surface Microphones

Unsteady Surface Pressure Sensor Locations


Measuring the unsteady surface pressures that manifest themselves in the acoustic farfield as sound is paramount in this investigation. The small size of the landing gear
model allowed for only a small number of surface locations to be investigated. Studies
on tandem cylinders have demonstrated that downstream components experience much
higher pressure fluctuations than those experienced by upstream ones [24]. The primary
cause of this was determined to be the impingement of the shear layer of the upstream
component onto the downstream one. Thus, the effects of the drag strut on the oncoming
propeller flow were of primary interest in this study. Sensors with a high dynamic range

30

Chapter 4. Experimental Method

were selected to enable the measurement of the large pressure fluctuations anticipated
with this configuration.
Unsteady surface pressure measurements were acquired on the model at the locations
indicated in Figure 4.6. The spacing ratios, S/D, at each microphone location based
on the upstream drag strut diameter and normal distance from the yawed cylinder are
listed in Table 4.2. Knowles Electronics EK-26899-P03 microphones were affixed with
epoxy into cut-outs on the wheel tread and the main strut to provide this pressure
fluctuation data. There were a total of 8 sensor locations: 4 along the span of the main
strut facing the oncoming flow, and 2 on each wheel, one directly facing oncoming flow
and another 45 below it on the wheel tread surface. These locations were strategically
chosen to investigate the presence of flow structures associated with the propeller and
their interactions with the drag strut. Note that the positions of the wheel microphones
are adjustable by rotation of the wheel. However, only the stated locations were used
in this study. Microphone wires have been routed through the model and out of the
airstream, so as not to influence the flow-field and acoustics.
Table 4.2: Spacing ratios, Sn /D, at the unsteady surface pressure sensor locations based
on drag strut diameter and normal distance between cylinders.
Microphone
Location
A
B
C
D

Spacing Ratio,
Sn /D
7.92
6.89
5.72
4.54

Calibration of Surface Sensors


Characterization of the surface (measurement) microphones was conducted in-situ with
the use of a function generator, speaker, and calibrated free-field (reference) microphone.
The reference microphone was placed adjacent the surface sensor with their diaphragms
parallel to one another and 0.2 cm apart. The centre of the speaker cone was aligned with
the centre of the two microphones, approximately 5 cm away. The function generator was
used to perform a logarithmic sweep through a frequency range from 100 Hz to 8 kHz at
2 Vpp . Data were then acquired from both sensors in order to calibrate the measurement
microphone relative to the reference. This method is similar to the linear system approach
presented by Snarski [40]. Details of the calibration procedure used in the flow control
laboratory can be found in the internal report by Zhao [50]. The sensitivity of the surface
sensor, Sm , was computed using

Chapter 4. Experimental Method

Sm =

Sref Aref |Gxy |


Am Gxx

31

(4.3)

where Sref is the sensitivity of the reference microphone, Aref is the gain of the reference
microphone, Gxx is the estimated power spectrum of the reference microphone, Am is the
gain of the measurement microphone, and Gxy is the estimated cross spectrum of the two
signals.
A representative calibration curve from the surface sensor calibration procedure is
shown in Figure 4.7. A nearly flat region in the sensitivity was observable for all measurement microphones in the range of 200 Hz to 1 kHz. The average value of the sensitivity
over this frequency interval was selected as the nominal sensitivity for all acoustic tests.

4.3.3

Data Acquisition & Uncertainty

In addition to microphone measurements, flow temperature and speed were collected


during each test run to determine the flow Mach number and fluid properties, such as
viscosity. Atmospheric pressure was obtained prior to each test from the website of Environment Canada [11]. Flow speeds were measured via a pitot-static tube in the wind
tunnel collector connected to an MKS Type 120AD Baratron differential pressure transducer operating in 10 Torr mode, since the flow speeds of these experiments were less than
that of the aerodynamic survey described in Section 3.2.1. Temperature measurements
were made using a J-type thermocouple positioned in the wind tunnel collector.
Surface microphone signals were passed through a circuit designed to increase the
dynamic range of sensors. The unsteady pressure signals were then passed through
amplifiers with nominal gains of 20 dB and built-in bandpass filters of 20 Hz - 10 kHz. All
unsteady signals were passed through 10 kHz antialiasing filters and were AC-coupled to
the data acquisition unit to remove any DC bias. A schematic of the instrumentation
layout is provided in Figure 4.8.
All data were sampled for 180 seconds at 32,878 Hz using a 16-bit National Instruments PCIe-6259 A/D data acquisition card, yielding a total number of samples per test
run of N = 5.982 106 . In order to resolve the tonal peaks associated with the propeller
noise emissions, a frequency resolution of f = 2 Hz was achieved by dividing the data
into Nfft = 16, 384 samples per block. The number of records averaged was nrec =361.
The unsteady pressure measurements at the model surface and in the far-field are
subject to both bias and random error. The root mean square error of spectral data is
expressed by

32

Chapter 4. Experimental Method

q
 = 2r + 2b .

(4.4)

For each sensor, the normalized random error in the autospectrum, Gxx , of the measured
data can be calculated using the following equation from Bendat and Piersol [3]:
r =

1
.
nrec

(4.5)

The normalized bias error is estimated using the equation below, also from [3]:
b =

b[Gxx ]
f 2 G00xx

,
Gxx
24 Gxx

(4.6)

where b[Gxx ] is the bias in the autospectral estimate. Since this equation is difficult to
compute for a random system and the bias error is expected to be small relative to the
random component, the bias error is neglected. This yields a total error, , equal to
5.26% for the spectral estimates of the measured data.

33

Chapter 4. Experimental Method

0.6
M = 0.10, J = 0.52
M = 0.07, J = 0.54
M = 0.04, J = 0.34

0.5

rp / Dp

0.4

0.3
0.2

0.1
0
0.4

0.6

0.8

1.2

1.4

1.6

1.8

Ua / U

Figure 4.3: Propeller wake time-averaged velocity profiles for the 3-blade, Master
Airscrew 15 7 propeller at M = 0.04, 0.07 and 0.10, at maximum throttle.

0.6
3-blade, J = 0.52
2-blade, J = 0.54

0.5

rp / Dp

0.4

0.3
0.2

0.1
0
0.4

0.6

0.8

1.2

Ua / U

Figure 4.4: Propeller wake time-averaged velocity profiles for the 3-blade, Master
Airscrew 157 propeller and the 2-blade, Master Airscrew 158 propeller at M = 0.10
and maximum throttle.

34

Chapter 4. Experimental Method

G5

15
G4
G3

58

G2

OW
FL

G1
21
6

F5
F4
F3

20

F2
20

F1
20
10

Figure 4.5: Schematic drawing of far-field microphone test locations. Dimensions are in
centimeters.

A
18.7

16.2

13.7

11.1
B

28.7
31.9
C

R1

L1

R2

L2

Figure 4.6: Schematic drawing of model surface microphone locations. Drag strut omitted
for clarity. Dimensions are in centimeters.

35

Chapter 4. Experimental Method

8.5

Mic rophone A

S ens itivity, mV /P a

8.4
8.3
8.2
8.1
8
7.9
7.8
7.7
7.6
7.5

3 4 5

F requenc y, kHz
Figure 4.7: Sample calibration curve for a surface microphone sensor (Microphone A).
Solid line represents the mean and dashed line represents one standard deviation of the
mean between 200 Hz and 1 kHz, respectively.

36

Chapter 4. Experimental Method

Larson-Davis
2200C Microphone
Power Supply

Antialiasing
Filter

Aeroacoustics
Laboratory
Computer

National Instruments
PCIe-6259 A/D
Data Acquisition Card

Arduino Duemilanova
Microcontroller

20 dB Signal
Amplifier

Bruel & Kjaer


Type 4192, 1/2
Condenser Microphones

High SPL
Circuit

Phototransistor
Switch

E-flite 60 Amp
Electronic Speed
Controller

EK-26899-P03
Surface Microphone
Sensors

OPB606A Reflective
Object Sensor

E-flite Power 46
Brushless Outrunner
Motor

Eagle Tree Systems


eLogger v4

Brushless Motor
RPM Sensor

The Motion Group


MMC-43-3C
Motor Controller

ESC Temperature
Sensor

Streamwise Traverse

Figure 4.8: Instrumentation layout for data acquisition .

5 | Results & Analysis


5.1

Far-field Acoustics

As indicated in Table 4.1, far-field acoustics were measured for nine different configurations tested at three wind tunnel Mach numbers. In the following sections, spectra from
measurements taken at location F3, as shown in Figure 4.5, are presented.

5.1.1

Static Configurations

Figure 5.1 shows the far-field SPL spectra for all the configurations at each wind tunnel
flow speed. Background far-field measurements of the empty test section are also plotted
for comparison. Without the propeller or motor actuating, the levels for each configuration are nearly indistinguishable from one another as shown in Figure 5.1a. Figure 5.1b
focuses on the range of 100 Hz to 2000 Hz, where sharp tonal peaks are observed near
1 kHz. It is suspected that these tones result from whistling across small gaps around the
test chamber door. Despite numerous efforts to completely seal the door, narrow peaks
in the range of 800 Hz to 1.3 kHz occasionally are apparent in the far-field spectra. In the
case of M = 0.10, two small broad peaks rising 4 dB over background levels are observed
centred at 0.8 kHz and 1.1 kHz respectively for S-1 and S-3. Further investigations are
needed to identify the cause of these peaks as they do not directly correspond to vortex
shedding from any model component at St = 0.2 or any of the peaks observed in the
unsteady surface pressure spectra.
For M = 0.04 to 0.10, the expected equivalent Reynolds number range for the drag
strut and main strut is 5k < Ren < 30k. This flow state is characterized by transition
in the shear layer [48]. Thus, the expected equivalent Strouhal number range of the
tandem cylinder system at microphone locations A, B, and C is 0.17 < Stn < 0.19. For
microphone location D, the spacing ratio of Sn /D = 4.54 yields an expected value of
0.12 < Stn < 0.15. Vortex shedding behind an isolated main strut is expected to occur
at StD = 0.2. No peaks corresponding to the expected vortex shedding are observed in
the far-field acoustic spectra.
37

38

Chapter 5. Results & Analysis

110
S -1- M = 0.04
S -2- M = 0.04
S -3- M = 0.04
B kgd - M = 0.04
S -1- M = 0.07
S -2- M = 0.07
S -3- M = 0.07
B kgd- M = 0.07
S -1- M = 0.10
S -2- M = 0.10
S -3- M = 0.10
B kgd- M = 0.10

S P L , dB re 20 P a

100
90
80
70
60
50
40
30
0
10

10

10

10

10

F requency , Hz
(a)

S -1- M = 0.04
S -2- M = 0.04
S -3- M = 0.04
B kgd - M = 0.04
S -1- M = 0.07
S -2- M = 0.07
S -3- M = 0.07
B kgd- M = 0.07
S -1- M = 0.10
S -2- M = 0.10
S -3- M = 0.10
B kgd- M = 0.10

S P L , dB re 20 P a

70

60

50

40

30
500

1000

2000

F requency , Hz
(b)

Figure 5.1: Far-field SPL spectra of configurations S-1, S-2, and S-3. for frequency ranges
of (a) 1 Hz-20 kHz and (b) 100 Hz-2 kHz

Chapter 5. Results & Analysis

5.1.2

39

Motor and Propeller Induced Noise

The far-field acoustic SPL spectra for the M and P configurations are shown in Figures
5.2 and 5.3. The spectra typically show behaviour identical to the background noise
of the tunnel, with tones of the motor fundamental frequency, fm , and propeller blade
passing frequency, BPF. Considering when the motor is spinning and the propeller is
absent (M configurations), the dominant peak of all far-field spectra occurs at the second
harmonic of the motor fundamental frequency, fm . The peak expected at fm only exceeds
the background levels for M = 0.10 and can be seen as a small peak at 122 Hz in
Figure 5.2b. For all values of M , the strength of the harmonics of fm are greatest
in configuration M-2, when the main strut is cantilevered from the nacelle, without the
support of the drag strut. This may be indicative of structural vibrations of the model
support at the motor fundamental frequency. Upon inclusion of the drag strut (M-3),
the main strut is more rigidly supported, and modes 3-7 of fm are attenuated as shown
in Figure 5.2b. Reductions of 2-3 dB SPL are evident from the fourth harmonic onward
until the broadband background noise overcomes the peaks in the spectrum.
When the propeller is added (configurations P-1, P-2, and P-3), the dominant peak
is the BPF for all test cases. For M = 0.04 and M = 0.07, the propeller is 3-bladed
and thus the BPF = 3fm , whereas for M = 0.10, the propeller has only two blades
so the BPF = 2fm . Contrary to the spectra of the M configurations (Figure 5.2b), the
strengths of the upper fm harmonics are not reduced when the drag strut supports the
main strut. Natural harmonics of the BPF are stronger than their neighbouring BPF
subharmonics. For M = 0.10, the two strongest tones of the far-field acoustic spectra
occur at BPF and 2BPF, which is in contrast to the lower speed cases where the strongest
tone is found at the BPF but the next largest peak is a subharmonic (2/3BPF = 2fm ).
For the complete landing gear configuration, the isolated motor noise at the BPF is at
least 15 dB less than when the propeller is present. The inclusion of a rotating propeller
increases the levels of the expected tonal peak at fm , which was not prominent in the M
configurations. A broadband increase in levels of approximately 3 dB for all frequencies
greater than the BPF is also observable.

5.2

Unsteady Surface Pressures

For each case outlined in Table 4.1 that included the main strut and wheels (S-2, S3, M-2, M-3, P-2, and P-3), unsteady surface pressure measurements were acquired
simultaneously with the far-field acoustic data as described in Section 4.3.2.

40

Chapter 5. Results & Analysis

110
M-1- M = 0.04
M-2- M = 0.04
M-3- M = 0.04
B kgd- M = 0.04
M-1- M = 0.07
M-2- M = 0.07
M-3- M = 0.07
B kgd- M = 0.07
M-1- M = 0.10
M-2- M = 0.10
M-3- M = 0.10
B kgd- M = 0.10

S P L , dB re 20 P a

100
90
80
70
60
50
40
30
0
10

10

10

10

10

F requency , Hz
(a)

M-1- M = 0.04
M-2- M = 0.04
M-3- M = 0.04
B kgd- M = 0.04
M-1- M = 0.07
M-2- M = 0.07
M-3- M = 0.07
B kgd- M = 0.07
M-1- M = 0.10
M-2- M = 0.10
M-3- M = 0.10
B kgd- M = 0.10

S P L , dB re 20 P a

70

60

50

40

30
500

1000

2000

F requency , Hz
(b)

Figure 5.2: Far-field SPL spectra of configurations M-1, M-2, and M-3 for frequency
ranges of (a) 1 Hz-20 kHz and (b) 100 Hz-2 kHz.

41

Chapter 5. Results & Analysis

110
P -1 - M = 0.04
P -2 - M = 0.04
P -3 - M = 0.04
B kgd - M = 0.04
P -2 - M = 0.07
P -2 - M = 0.07
P -3 - M = 0.07
B kgd - M = 0.07
P -1 - M = 0.10
P -2 - M = 0.10
P -3 - M = 0.10
B kgd - M = 0.10

S P L , dB re 20 P a

100
90
80
70
60
50
40
30 0
10

10

10

10

10

F requency , Hz
(a)

90
P -1 - M = 0.04
P -2 - M = 0.04
P -3 - M = 0.04
B kgd - M = 0.04
P -2 - M = 0.07
P -2 - M = 0.07
P -3 - M = 0.07
B kgd - M = 0.07
P -1 - M = 0.10
P -2 - M = 0.10
P -3 - M = 0.10
B kgd - M = 0.10

S P L , dB re 20 P a

80
70
60
50
40
30
500

1000

2000

F requency , Hz
(b)

Figure 5.3: Far-field SPL spectra of configurations P-1, P-2, and P-3 for frequency ranges
of (a) 1 Hz-20 kHz and (b) 100 Hz-2 kHz.

42

Chapter 5. Results & Analysis

0.2

Rs/Rp

0.4
0.6
0.8

0.5

1.0

x/Rp

1.5

2.0

2.5

a = 0.5 (M = 0.04)

a = 0.2 (M = 0.07)

a = 0.1 (M = 0.10)

1.0
D

1.2

Figure 5.4: Slipstream contraction for all test cases with strut microphone locations
indicated. Computed using (2.8) for inflow ratios based on Figures 4.3 and 4.4.

5.2.1

Strut Surface Microphones

Figures 5.5, 5.6, and 5.7 show the surface SPL spectra for each strut microphone at each
wind tunnel flow speed. Each subfigure investigates the case with and without the drag
strut for a given propeller mode.
The expected radius of the slipstream relative to the main strut microphones based
on (2.8) are shown in Figure 5.4. For all test conditions, microphones A, B, and C are
expected to be within the propeller slipstream. However, microphone D approaches the
limit of the slipstream for M = 0.04. Thus, the strut is expected to experience both
the influence of propeller tip vortex when and strong axial shear at that location when
the flow is unimpeded.
Isolated Main Strut
For the isolated main strut, unsteady surface pressure spectra at microphone locations A
- D are presented as the dashed lines on the left side of in Figures 5.5, 5.6, and 5.7. The
Strouhal number presented in these cases is based on the nominal freestream velocity,
U , and the main strut diameter. For all tests omitting the propeller (i.e. S-2 and M-2),

Chapter 5. Results & Analysis

43

a broad hump spans the lower frequencies and is typically capped by the vortex shedding
frequency of the main strut (St 0.2). For M = 0.04 and 0.07, at frequencies greater
than 1 kHz, the pressure signal quality deteriorates as electrical noise dominates. The
steep roll-off above 10 kHz is due to the antialiasing filter employed.
The spinning motor (Configuration M-2) increases the tonal content of the spectra
with a strong peak at the motor fundamental frequency, fm , reaching SPL values between 95-105 dBSPL. Harmonics of fm are also evident with the tonal spectral levels
monotonically decreasing. Upon the introduction of the motor, the peak associated with
the vortex shedding frequency increases by 10-15 dBSPL for M = 0.04 and 0.07. At
M = 0.10, peak levels at St = 0.2 remain unchanged. This may be due to broadband
noise produced by the motor or possibly vibrations emanating throughout the model and
support structure. Further investigation is necessary to identify the cause of this level
increase. Microphone A is most heavily influenced by the motor, exhibiting a broadening
of the peak at the cylinder vortex shedding frequency, which is indicative of increased
turbulence levels and less coherent vortex shedding.
Adding the propeller to the isolated main strut (configuration P-2) causes the BPF
harmonics to rise above the neighbouring tones (harmonics of fm ). The BPF tones are
most prominent on microphones C and D, which are closest to the plane of the propeller
tip. More than 20 harmonics of fm are observed on microphones C and D for M = 0.07
in this case.
Tandem Configuration
The unsteady surface pressure spectra at microphone locations A - D upon inclusion
of the drag strut (configurations S-3, M-3 and P-3) are shown on the right hand side of
Figures 5.5, 5.6, and 5.7. In these cases, the indicated Strouhal number is calculated based
on the normal component of the velocity, Un = U cos , and the drag strut diameter.
The turbulent wake of the drag strut is manifested as a broadband SPL increase over the
entire measured bandwidth of each of the surface unsteady pressure sensors. An increase
by 20-30 dB relative to the case of the isolated main strut is observed and is of a generally
consistent shape between configurations for a given flow speed.
Peaks observed at the vortex shedding frequency of the drag strut (St 0.2) are
broader, and for configurations S-3 and M-3 are only observed at microphone locations C
and D. This is consistent with the work of Wilkins [45], which shows a weakening of the
spectral peaks associated with vortex shedding of a tandem yawed cylinder arrangement
at high spacing ratios (S/D > 4.5) for Re = 56k. He suggests that as the spacing ratio is
increased, the shear layer of the upstream cylinder no longer impinges on the downstream

Chapter 5. Results & Analysis

44

cylinder at which point coherent vortex shedding ceases. The spacing ratio where this
behaviour occurs in the current study is 5.72 < Sn /D < 6.89 for 6.4k< Ren <19.6k.
In addition to the primary vortex shedding peak, the inclusion of the drag strut
yields a secondary peak at twice the shedding frequency, St 0.4, in configurations
S-3 and M-3. This peak has been widely observed at subcritical Reynolds numbers for
measurements not far off the wake centreline [36] and corresponds to the impingement of
both shear layers of the upstream cylinder on the main strut. These secondary peaks are
observed on all surface pressure spectra which exhibit the peak at the shedding frequency
for configurations S-3 and M-3. This behaviour is primarily observed on microphones C
and D.
Typically, the unsteady surface pressure spectra of microphones A - D do not exhibit
dramatic changes between cases with and without the propeller (i.e. M-3 and P-3). For
M = 0.04 and 0.07, the influence of the propeller is observed on the surface unsteady
pressure spectra of all microphones as shown on the right side of Figures 5.5c and 5.6c.
In these cases, the peak associated with vortex shedding appears to shift along the span
of the main strut. At microphone locations A and B, a peak is observed near Stn = 0.25.
This peak reduces in strength and shifts toward Stn = 0.2 at microphone locations A and
B. This may be an indication of categorically different vortex shedding regimes along the
drag strut. Further investigation should be undertaken to examine the flow structures
associated with the spectral peaks and their spanwise distributions.
The BPF tone and its first two harmonics peak over the background levels in configuration P-3 and are easily distinguishable from the peaks related to vortex shedding.
The peaks at the BPF are comparable in magnitude to the vortex shedding peak of the
tandem configuration. However, in comparison to the BPF tone in the isolated strut case
(P-2), the magnitude is decreased by approximately 5 dB. This may be due to a shielding
effect of the drag strut, breaking the structures associated with the blade vortex sheet
prior to them impinging on the main strut.

5.2.2

Wheel Surface Microphones

Figures 5.8, 5.9, and 5.10 show the unsteady wheel surface pressures at locations L1, L2,
R1, and R2 for all test configurations. Since the wheels are not in the slipstream of the
propeller, or the wake of the drag strut, the changes in unsteady pressures from case
to case are minimal. Each wheel surface microphone has levels that are nearly identical
to those of the sensor located symmetrically across the x z plane (e.g. L1 and R1).
Compared to the levels of unsteadiness experienced by the main strut, the wheels do not

45

Chapter 5. Results & Analysis

Stn

St
10 -2

10 -1

10 0

10 1

10 -2

140

10 0

10 1

140
130
A
B
C
D

120
110
100
90
80
70
60

S P L , dB re 20 P a

130

S P L , dB re 20 P a

10 -1

S -2

50
40
10

A
B
C
D

120
110
100
90
80
70
60

S -3

50
10

10

10

10

40

10

10

F requenc y, Hz

10

10

10

F requenc y, Hz

(a) Static
Stn

St
10 -2

10 -1

10 0

10 1

10 -2

140

10 0

10 1

140
130
A
B
C
D

120
110
100
90
80
70
60

fm

M-2

50
40
10

10

10

S P L , dB re 20 P a

130

S P L , dB re 20 P a

10 -1

A
B
C
D

120
110
100
90
80
70
60

10

10

40

fm

M-3

50
10

10

F requenc y, Hz

10

10

10

F requenc y, Hz

(b) Motor
Stn

St
10 -3

10 -2

10 -1

10 0

10 1

10 -3

140

10 -1

10 0

10 1

140
130
A
B
C
D

120
110
100
90
80
70
60

B PF

P -2

50
40
10

10

10

S P L , dB re 20 P a

130

S P L , dB re 20 P a

10 -2

A
B
C
D

120
110
100
90
80
70

P -3

50
10

10

B PF

60

40
10

F requenc y, Hz

10

10

10

10

F requenc y, Hz

(c) Propeller

Figure 5.5: Strut unsteady surface pressure spectral levels, M = 0.04

46

Chapter 5. Results & Analysis

Stn

St
10 -3

10 -2

10 -1

10 0

10 1

10 -3

140

10 -1

10 0

10 1

140

130

130

A
B
C
D

120
110

S P L , dB re 20 P a

S P L , dB re 20 P a

10 -2

100
90
80
70
60
50
10

110
100
90
80
70
60
50

S -2

40

A
B
C
D

120

10

10

10

10

S -3

40

10

10

F requenc y, Hz

10

10

10

F requenc y, Hz

(a) Static
Stn

St
10 -3

10 -2

10 -1

10 0

10 1

10 -3

140

10 -1

10 0

10 1

140

130

130

A
B
C
D

120
110

S P L , dB re 20 P a

S P L , dB re 20 P a

10 -2

100
90
80
70

fm

60
50

M-2

40
10

10

10

A
B
C
D

120
110
100
90
80
70

50

10

10

M-3

40

fm

60

10

10

F requenc y, Hz

10

10

10

F requenc y, Hz

(b) Motor
St
10 -3

10 -2

10 -1

10 0

10 1

10 -3

140

Stn

10 -1

10 0

140
A
B
C
D

120
110
100
90
80
70
60
50

B PF

P -2

40
10

10

10

130

S P L , dB re 20 P a

130

S P L , dB re 20 P a

10 -2

A
B
C
D

120
110
100
90
80
70

B PF

60
50

10

10

P -3

40
10

F requenc y, Hz

10

10

10

10

F requenc y, Hz

(c) Propeller

Figure 5.6: Strut unsteady surface pressure spectral levels, M = 0.07

47

Chapter 5. Results & Analysis

St
10 -3

10 -2

10 -1

10 0

10 -3

140
A
B
C
D

120

Stn

10 -1

10 0

110
100
90
80
70
60
50

130

S P L , dB re 20 P a

130

S P L , dB re 20 P a

10 -2

140

10

110
100
90
80

10

10

10

10

60

S -3

40

A
B
C
D

70

50

S -2

40

120

10

10

F requenc y, Hz

10

10

10

F requenc y, Hz

(a) Static
St
10 -3

10 -2

10 -1

10 0

10 -3

140
A
B
C
D

120

Stn

10 -1

10 0

110
100
90
80
70

fm

60
50

M-2

40
10

10

10

130

S P L , dB re 20 P a

130

S P L , dB re 20 P a

10 -2

140

120
110
100
90
80

10

M-3

40

fm

60
50

10

A
B
C
D

70

10

10

F requenc y, Hz

10

10

10

F requenc y, Hz

(b) Motor
St
10 -3

10 -2

10 -1

10 0

10 -3

140

Stn

10 -1

10 0

140
A
B
C
D

120
110
100
90
80
70

B PF

60
50
10

120
110
100
90
80

10

10

10

10

A
B
C
D

70
60
50

P -2

40

130

S P L , dB re 20 P a

130

S P L , dB re 20 P a

10 -2

B PF

P -3

40
10

F requenc y, Hz

10

10

10

10

F requenc y, Hz

(c) Propeller

Figure 5.7: Strut unsteady surface pressure spectral levels, M = 0.10

Chapter 5. Results & Analysis

48

appear to play an important role in the flow interactions with the propeller or drag strut.
Reorienting the wheels such that the surface microphones are in a separated region of
the flow may give indications of wake interactions. Future studies involving landing gear
in the unsteady wake of a propeller should concentrate efforts on the interaction of the
slipstream and strut(s).

49

Chapter 5. Results & Analysis

S P L , dB re 20 P a

140
L1
L2
R1
R2

120
100
80
60
40 0
10

10 1

10 2

10 3

10 4

F requenc y, Hz

(a) Static
10

-2

10

-1

B PF

10

10

S P L , dB re 20 P a

140
L1
L2
R1
R2

120

fm
100
80
60
40 0
10

10 1

10 2

10 3

10 4

F requenc y, Hz

(b) Motor
10

-2

10

-1

B PF

10

10

S P L , dB re 20 P a

140
L1
L2
R1
R2

120

B PF
100
80
60
40 0
10

10 1

10 2

10 3

10 4

F requenc y, Hz

(c) Propeller

Figure 5.8: Wheel unsteady surface pressure spectral levels, M = 0.04

50

Chapter 5. Results & Analysis

S P L , dB re 20 P a

140
L1
L2
R1
R2

120
100
80
60
40 0
10

10 1

10 2

10 3

10 4

F requenc y, Hz

10

-2

10

-1

(a) Static
B PF

10

10

S P L , dB re 20 P a

140
L1
L2
R1
R2

120

fm
100
80
60
40 0
10

10 1

10 2

10 3

10 4

F requenc y, Hz

10

-2

10

-1

(b) Motor
B PF

10

10

S P L , dB re 20 P a

140
L1
L2
R1
R2

120

B PF
100
80
60
40 0
10

10 1

10 2

10 3

10 4

F requenc y, Hz

(c) Propeller

Figure 5.9: Wheel unsteady surface pressure spectral levels, M = 0.07

51

Chapter 5. Results & Analysis

S P L , dB re 20 P a

140
L1
L2
R1
R2

120
100
80
60
40

10

10

10

10

10

F requenc y, Hz

10

S P L , dB re 20 P a

140

-2

10

-1

(a) Static
B PF 0
10

10

L1
L2
R1
R2

120

fm
100
80
60
40

10 0

10 1

10 2

10 3

10 4

F requenc y, Hz

10

S P L , dB re 20 P a

140

-2

10

-1

(b) Motor
B PF 0
10

120

10

L1
L2
R1
R2

B PF

100
80
60
40

10 0

10 1

10 2

10 3

10 4

F requenc y, Hz

(c) Propeller

Figure 5.10: Wheel unsteady surface pressure spectral levels, M = 0.10

6 | Conclusions
6.1

Wind Tunnel Characterization

Prior to conducting the aeroacoustic testing described in the current work, it was necessary to characterize the aerodynamic and acoustic properties of the Acoustic Wind
Tunnel Facility at UTIAS due to many recent modifications. Irregularities in the mean
flow profile were observed to be sufficiently reduced by x/DN = 0.3 for test velocities
of 33 and 55 m/s. The maximum attainable flow velocity in the current configuration
was found to be approximately 60 m/s. Turbulence intensity in the potential core was
measured at less than 1% for x/DN 1.4. As pertaining to the tunnel acoustics, far-field
noise spectra were presented showing levels comparable to similarly sized wind tunnels.
This study provides a basis for future wind tunnel testing in the facility and an indication
of the aerodynamic and acoustic limits that are imposed for such testing.

6.2

Model Development

The development of the model constituted a significant portion of this project. The use of
model-scale parts required thrust test data based on slipstream measurements in order to
determine axial inflow factors. The model set-up allowed for tests including combinations
of the nacelle, main strut, wheels, drag strut, and propeller. Configurations including the
main strut and wheels were equipped with unsteady surface pressure sensors to measure
the interactions with the drag strut and propeller. Descriptions of the sensor calibration
procedure and the data acquisition instrumentation are given followed by a discussion of
the uncertainties in the presented spectral estimates.
52

Chapter 6. Conclusions

6.3

53

Landing Gear Testing

The first ever far-field noise and surface pressure measurements taken for a two-strut,
two-wheeled landing gear with an actuated upstream propeller have been compared and
discussed for various configurations at Mach numbers of 0.04, 0.07, and 0.10. The propeller and model support were shown to dominate the far-field acoustic signature for
most test cases.
Unsteady surface pressure measurements on the landing gear main strut and wheels
were also recorded. For an isolated main strut, the pressure fluctuations on the main strut
were observed to exhibit the strongest fluctuations at the blade passing frequency and
its harmonics. When the drag strut was added, these pressure fluctuations were masked
by a broadband increase in levels due to the wake of the upstream yawed cylinder. Two
peaks corresponding to vortex shedding from the drag strut were observed above the
broadband levels for the main strut surface microphones located nearest to the edge of
the propeller slipstream.
The recorded sound pressure level spectra can be used as an indication of what to
expect for full-scale testing thereby enabling the development of more accurate noise
prediction tools. The results of this study also provide a new benchmark case for computational aeroacoustic codes that are used in design and analysis of aircraft landing
gear.

6.4

Recommendations for Future Work

The current work represents the first attempt at investigating noise from landing gear
in a propeller slipstream. In order to continue the analysis of these configurations and
improve the quality of the data, some suggestions for future work are relayed below.
The structure of the motor support and the landing gear model should be decoupled for further testing. Efforts were made to reduce to the vibrations propagating
through the structure for the current work. However, oscillations at motor frequency, fm , were evident in all surface pressure measurements. Though the model
did not appear to move by visual inspection, vibrations could be felt by placing a
hand on the wheel while the motor was spinning.
In order to more adequately represent the full-scale configuration, acquisition of
an accurately scaled down propeller should be considered. An investigation using
higher fidelity components as such would give valuable insight to the drawbacks

Chapter 6. Conclusions

54

of using remote control model parts. Operating at lower Reynolds numbers than
actual flight conditions may invoke difficulties with respect to matching both aerodynamic and acoustic parameters for propellers and may constitute an additional
experimental program.
Future studies should concentrate efforts on strut interactions as the influence of the
drag strut appears to dominate the surface pressure spectra across the span of the
main strut. A study examining the observed spanwise variation in vortex shedding
is of particular interest; It is advised that hot-wire or PIV investigations of the
tandem yawed cylinder configuration be performed to identify the flow structures
surrounding the cylinders and quantify the local flow conditions. These localized
variations of the flow field may have a significant effect on prediction models as
shown by Guo [17].
Techniques such as those of Siddon [39] and Pan [31] could be used to characterize
the acoustics of the fluid-surface interactions. These cross-spectral analysis methods
of the far-field and surface pressure signals may identify the regions of greatest
influence to the far-field acoustics, allowing for more focussed efforts to reduce
noise emissions.

References
[1] N. G. Ball. Acoustic Excitation of Tollmien-Schlichting Waves in a Laminar Boundary Layer. Ph.D. dissertation, University of Toronto, 1987.
[2] P. W. Bearman. On vortex shedding from a circular cylinder in the critical Reynolds
number rgime. Journal of Fluid Mechanics, 37(3):577585, 1969.
[3] J. S. Bendat and A. G. Piersol. Engineering Applications of Correlation and Spectral
Analysis. John Wiley and Sons, New York, 2nd edition, 1993.
[4] J. L. Beveridge. The Radial Distribution of Propeller Thrust from Model Wake Measurements. Technical report, Department of the Navy, Hydromechanics Laboratory,
1958.
[5] Bombardier Aerospace. Q400 Customer Images. http://media.bombardiercms.
com/q400/medias/gallery/q400customerImages/high/06.jpg.
[6] Bombardier Aerospace. Smart Cockpit: Bombardier Dash-8-400 Landing Gear.
http://www.smartcockpit.com/docs/Q400-Landing\_Gear\_1.pdf.
[7] N. Curle. The influence of solid boundaries upon aerodynamic sound. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences,
231(1187):505514, Sept. 1955.
[8] W. M. Dobrzynski. Almost 40 years of airframe noise research: What did we achieve?
Journal of Aircraft, 47(2):353367, Mar. 2010.
[9] W. M. Dobrzynski, M. Pott-Pollenske, D. Foot, and M. Goodwin. Landing gears
aerodynamic interaction noise. In European Congress on Computational Methods in
Applied Sciences and Engineering 2004, pages 115135, Jyvskyl, Finland, 2004.
Multi-Science.
55

References

56

[10] W. M. D. Dobrzynski and H. Buchholz. Full-scale noise testing on Airbus landing gears in the German Dutch Wind Tunnel. In 3rd AIAA/CEAS Aeroacoustics
Conference, Atlanta, Georgia, 1997. AIAA Paper 19971597, American Institute of
Aeronautics and Astronautics.
[11] Environment Canada. Toronto, ON - 7 Day Forecast - Environment Canada. http:
//www.smartcockpit.com/docs/Q400-Landing\_Gear\_1.pdf.
[12] Federal Aviation Administration. Part 36: Noise Standards: Aircraft Type and
Airworthiness Certification. In Code of Federal Regulations, Title 14: Aeronautics
and Space. U.S. Government Printing Office, Washington, D.C., 2011.
[13] J. E. Ffowcs Williams and D. L. Hawkings. Sound generation by turbulence and
surfaces in arbitrary motion. Philosophical Transactions of the Royal Society of
London. Series A, Mathematical and Physical Sciences, 264(1151):321342, May
1969.
[14] J. S. Gibson. Non-engine Aerodynamic Noise Investigation of a Large Aircraft.
Technical Report October, NASA Langley, Hampton, Virginia, 1974.
[15] M. E. Goldstein. Aeroacoustics. McGraw-Hill International Book Company, New
York, 1976.
[16] Y. Guo. A statistical model for landing gear noise prediction. Journal of Sound and
Vibration, 282(1-2):6187, Apr. 2005.
[17] Y. Guo. Effects of local flow variations on landing gear noise prediction and analysis.
Journal of Aircraft, 47(2):383391, Mar. 2010.
[18] Y. Guo, R. W. Stoker, and K. J. Yamamoto. Experimental study on aircraft landing
gear noise. Journal of Aircraft, 43(2):306317, Mar. 2006.
[19] Y. P. Guo, K. J. Yamamoto, and R. W. Stoker. An empirical model for landing gear
noise prediction. In 10th AIAA/CEAS Aeroacoustics Conference, Manchester, U.K.,
2004. AIAA Paper 20042888, American Institute of Aeronautics and Astronautics.
[20] A. Hanson. Vortex shedding from yawed cylinders. AIAA Journal, 4(4):738740,
1966.
[21] G. J. Healy. Measurement and Analysis of Aircraft Far-Field Aerodynamic Noise.
Technical Report December, NASA Langley, Hampton, Virginia, 1974.

References

57

[22] H. H. D. Heller and W. M. D. Dobrzynski. Sound radiation from aircraft wheelwell/landing-gear configurations. Journal of Aircraft, 14(8):768774, 1977.
[23] F. V. Hutcheson and T. F. Brooks. Noise radiation from single and multiple rod
configurations. In 12th AIAA/CEAS Aeroacoustics Conference, number May, Cambridge, Massachusetts, 2006. AIAA Paper 20062629, American Institute of Aeronautics and Astronautics.
[24] L. N. Jenkins, D. H. Neuhart, C. B. McGinley, M. M. Choudhari, and M. R. Khorrami. Measurements of unsteady wake interference between tandem cylinders. In
36th AIAA Fluid Dynamics Conference, Monterey, California, 2006. AIAA Paper
20063202, American Institute of Aeronautics and Astronautics.
[25] B. S. Lazos. Mean flow features around the inline wheels of four-wheel landing gear.
AIAA Journal, 40(2):193198, Feb. 2002.
[26] M. J. Lighthill. On sound generated aerodynamically. I. General theory. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences,
211(1107):564587, Mar. 1952.
[27] E. Manoha, J. Bult, and B. Caruelle. LAGOON: an experimental database for
the validation of CFD/CAA methods for landing gear noise prediction. In 14th
AIAA/CEAS Aeroacoustics Conference, number May, Vancouver, British Columbia,
2008. AIAA Paper 20082816, American Institute of Aeronautics and Astronautics.
[28] J. Mathew. Design, Fabrication, and Characterization of an Anechoic Wind Tunnel
Facility. Ph.D. dissertation, University of Florida, 2006.
[29] U. Michel and E. Froebel. Lower limit for the velocity fluctuation level in wind
tunnels. Experiments in Fluids, 6(1):4954, 1988.
[30] A. Okajima. Flows around two tandem circular cylinders at very high Reynolds
numbers. Bulletin of the Japan Society of Mechanical Engineers, 22(166):504511,
1979.
[31] Y. S. Pan. Cross-correlation methods for studying near- and farfield noise characteristics of flow-surface interactions. The Journal of the Acoustical Society of America,
58:586594, 1975.
[32] A. R. Quayle, A. P. Dowling, H. Babinsky, W. R. Graham, and Y. Liu. Phased
array measurements from landing gear models. In 13th AIAA/CEAS Aeroacoustics

References

58

Conference, Rome, Italy, May 2007. AIAA Paper 20073463, American Institute of
Aeronautics and Astronautics.
[33] P. A. Ravetta, R. A. Burdisso, and W. F. Ng. Wind tunnel aeroacoustic measurements of a 26%- scale 777 main landing gear model. In 10th AIAA/CEAS Aeroacoustics Conference, Manchester, U.K., 2004. AIAA Paper 20042885, American
Institute of Aeronautics and Astronautics.
[34] R. Reger, F. Liu, and L. Cattafesta. An experimental study of the rudimentary landing gear. In 51st AIAA Aerospace Sciences Meeting, number January, Grapevine,
Texas, 2013. AIAA Paper 20130464, American Institute of Aeronautics and Astronautics.
[35] A. K. Ringshia, P. A. Ravetta, W. F. Ng, and R. A. Burdisso. Aerodynamic measurements of the 777 main landing gear model. In 12th AIAA/CEAS Aeroacoustics
Conference, number May, Cambridge, Massachusetts, 2006. AIAA Paper 20062625,
American Institute of Aeronautics and Astronautics.
[36] A. Roshko. Experiments on the flow past a circular cylinder at very high Reynolds
number. J. Fluid Mech, pages 345356, 1961.
[37] W. R. Sears. The boundary layer of yawed cylinders. Journal of the Aeronautical
Sciences, 15(1):4952, 1948.
[38] M. Shirakashi, S. Ueno, Y. Ishida, and S. Wakiya. Vortex excited oscilation of a
circular cylinder in a uniform ulow. Bulletin of the Japan Society of Mechanical
Engineers, 27(228):11201126, 1984.
[39] T. E. Siddon. Surface dipole strength by cross-correlation method. The Journal of
the Acoustical Society of America, 53:619633, 1973.
[40] S. R. Snarski. Relation Between the Fluctuating Wall Pressure and the Turbulent
Structure Of a Boundary Layer on a Cylinder In Axial Flow. Technical Report
August, Naval Undesea Warfare Center, New London, Connecticut, 1993.
[41] P. T. Soderman and C. S. Allen. Microphone Measurements In and Out of Airstream.
In T. J. Mueller, editor, Aeroacoustic Measurements, pages 161. Springer-Verlag,
Berlin, Germany, 2002.
[42] P. R. Spalart, M. L. Shur, M. K. Strelets, and A. K. Travin. Towards noise prediction for rudimentary landing gear. In Procedia IUTAM, volume 1, pages 283292,
Southampton, U. K., 2010.

References

59

[43] W. J. G. Trebble, J. Williams, and R. P. Donnelly. Propeller noise at model- and


full-scale. Journal of Aircraft, 20(1):3441, 1983.
[44] L. L. M. Veldhuis. Propeller Wing Aerodynamic Interference. Ph.D. dissertation,
Technische Universiteit Delft, 2005.
[45] S. J. Wilkins. Vortex-Shedding in Yawed-Tandem Circular Cylinder Arrangements.
Ph.D. dissertatioin, University of New Brunswick, 2011.
[46] S. J. Wilkins and J. W. Hall. Experimental investigation of a tandem cylinder
system with a yawed upstream cylinder. Journal of Pressure Vessel Technology,
136:01130218, Feb. 2014.
[47] N. S. Zawodny. Aeroacoustic Characterization of Scaled Canonical Nose Landing
Gear Configurations. Ph.D. dissertation, University of Florida, 2013.
[48] M. Zdravkovich. Flow Around Circular Cylinders, Volume 1: Fundamentals. Flow
Around Circular Cylinders: A Comprehensive Guide Through Flow Phenomena, Experiments, Applications, Mathematical Models, and Computer Simulations. Oxford
University Press, New York, New York, 1997.
[49] M. M. Zdravkovich. Flow Around Circular Cylinders, Volume 2: Applications. Flow
Around Circular Cylinders: A Comprehensive Guide Through Flow Phenomena, Experiments, Applications, Mathematical Models, and Computer Simulations. Oxford
Universiy Press, New York, New York, 2003.
[50] W. Zhao. Microphone Calibration. Technical report, University of Toronto, Flow
Control and Experimental Turbulence Laboratory, Toronto, Ontario, 2013.

A | Wind Tunnel Layout

60

0.48

2.44

2.59 x 2.59

2.87
2.78

Nozzle

0.62

0.86

4.57

Jet Collector

Diffuser

0.91

1.22

Upstream Silencer

1.52

6.10

2.44

1.40

2.25

Plenum

Figure A.1: Schematic of the Acoustic Wind Tunnel Facility. All dimensions in meters.

0.61 0.91 0.28

0.91

Anechoic Chamber

0.70

Contraction

1.02

Fan and Motor

3.96

6.05

1.33

Downstream Silencer

3.05
1.83

Appendix A. Wind Tunnel Layout


61

B | Surface Microphone Calibration

62

3 4 5

F requenc y, kHz

3 4 5

Mic rophone L 1

F requenc y, kHz

7.6

7.8

8.2

8.4

8.5

8.6

8.7

8.8

8.9

9.1

9.2

9.3

9.4

9.5

3 4 5

F requenc y, kHz

3 4 5

Mic rophone L 2

F requenc y, kHz

Mic rophone B

7.5

7.6

7.7

7.8

7.9

8.1

8.2

8.3

8.4

8.5

7.5

7.6

7.7

7.8

7.9

8.1

8.2

8.3

8.4

8.5

3 4 5

F requenc y, kHz

3 4 5

Mic rophone R 1

F requenc y, kHz

Mic rophone C

7.5

7.6

7.7

7.8

7.9

8.1

8.2

8.3

8.4

8.5

8.1

8.2

8.3

8.4

8.5

8.6

8.7

8.8

8.9

3 4 5

F requenc y, kHz

3 4 5

Mic rophone R 2

F requenc y, kHz

Mic rophone D

Figure B.1: Calibration curve for all surface microphone sensors. Solid lines represent the mean and dashed lines represent two
standard deviations from the mean between 200 Hz and 1 kHz, respectively.

8.5

8.6

8.7

8.8

8.9

9.1

9.2

9.3

9.4

9.5

7.5

7.6

7.7

7.8

7.9

8.1

8.2

8.3

S ens itivity, mV /P a
S ens itivity, mV /P a

8.4

Mic rophone A
S ens itivity, mV /P a
S ens itivity, mV /P a

S ens itivity, mV /P a

S ens itivity, mV /P a

S ens itivity, mV /P a

S ens itivity, mV /P a

8.5

Appendix B. Surface Microphone Calibration


63

C | Landing Gear Model Drawings

64

ITEM NO.

1
2
3
4
5
6
7

PART NUMBER / DWG. NUMBER


UTIASAA-RLGM-010
UTIASAA-RLGM-020
UTIASAA-RLGM-030
B18.3.1M - 6 x 1.0 x 70 Hex SHCS -- 24NHX
B18.3.1M - 4 x 0.7 x 20 Hex SHCS -- 20NHX
B18.2.4.1M - Hex nut, Style 1, M6 x 1 --D-N
B18.2.4.1M - Hex nut, Style 1, M4 x 0.7 --D-N

SECTION A-A

DRAWN

NAME

DATE

RC

05/03/2013

CHECKED
UNLESS OTHERWISE SPECIFIED:
DIMENSIONS ARE IN INCHES

TOLERANCES:
.XX
.02

THIS DRAWING WAS PRODUCED USING:


SOFTWARE: SOLIDWORKS

1/16

REMOVE ALL BURRS AND SHARP EDGES.

NEXT ASSY

1
1
2
1
1
1
1

USED ON

INSTITUTE FOR AEROSPACE STUDIES


UNIVERSITY OF TORONTO
4925 DUFFERIN STREET
TORONTO, ONTARIO
M3H 5T6

TITLE:

RUDIMENTARY DASH-8 MAIN LANDING


GEAR WIND TUNNEL MODEL ASSEMBLY

FRACTIONS
1

VERSION: 2012

FILENAME: UTIASAA-RLGM-000

QTY.

DO NOT SCALE DRAWING

.XXX
.005

DESCRIPTION
MAIN STRUT ASSEMBLY
DRAG STRUT
WHEEL ASSEMBLY
M6 x 1.0 x 70 SHCS
M4 x 0.7 x 20 SHCS
M6 x 1HEX NUT
M4 x 0.7 HEX NUT

Appendix C. Landing Gear Model Drawings

PROPRIETARY AND CONFIDENTIAL


THE INFORMATION CONTAINED IN
THIS DRAWING IS THE SOLE
PROPERTY OF THE UNIVESITY OF
TORONTO. ANY REPRODUCTION IN
PART OR AS A WHOLE WITHOUT
THE WRITTEN PERMISSION OF THE
UNIVERSITY OF TORONTO IS
PROHIBITED.

SIZE DWG. NO.

UTIASAA-RLGM-000

SCALE: 1:3

WT: 987.84

REV

SHEET 1 OF 2

65

0.551

74

0.3

14.565

2.126

52
9.2

12.795

0.709

3.543

1.063

3.071
DRAWN

NAME

DATE

RC

05/03/2013

CHECKED
UNLESS OTHERWISE SPECIFIED:
DIMENSIONS ARE IN INCHES

TOLERANCES:
.XX
.02

THIS DRAWING WAS PRODUCED USING:


SOFTWARE: SOLIDWORKS

.005

TITLE:

RUDIMENTARY DASH-8 MAIN LANDING


GEAR WIND TUNNEL MODEL ASSEMBLY

FRACTIONS
1

1/16

REMOVE ALL BURRS AND SHARP EDGES.

NEXT ASSY

INSTITUTE FOR AEROSPACE STUDIES


UNIVERSITY OF TORONTO
4925 DUFFERIN STREET
TORONTO, ONTARIO
M3H 5T6
DO NOT SCALE DRAWING

.XXX

VERSION: 2012

FILENAME: UTIASAA-RLGM-000

Appendix C. Landing Gear Model Drawings

USED ON

PROPRIETARY AND CONFIDENTIAL


THE INFORMATION CONTAINED IN
THIS DRAWING IS THE SOLE
PROPERTY OF THE UNIVESITY OF
TORONTO. ANY REPRODUCTION IN
PART OR AS A WHOLE WITHOUT
THE WRITTEN PERMISSION OF THE
UNIVERSITY OF TORONTO IS
PROHIBITED.

SIZE DWG. NO.

UTIASAA-RLGM-000

SCALE: 1:3

WT: 987.84

REV

SHEET 2 OF 2

66

0.630

0.551

1.260

2
3
13.110

C
1.063

2.126

ITEM NO.

PART NUMBER/DWG. NUMBER

DESCRIPTION

QTY.

1
2
3

UTIASAA-RLGM-011
UTIASAA-RLGM-012
EK-26899-P03
B18.6.7M - M2.5 x 0.45 x 4 Type I
Cross Recessed FHMS --4N

MAIN STRUT
MAIN STRUT MICROPHONE INSERT
SURFACE MICROPHONE (KNOWLES)

1
1
4

M2.5 X 0.45 X 4 FHMS

DRAWN

NAME

DATE

RC

05/03/2013

CHECKED
UNLESS OTHERWISE SPECIFIED:
DIMENSIONS ARE IN INCHES

TOLERANCES:
.XX
.02

THIS DRAWING WAS PRODUCED USING:


SOFTWARE: SOLIDWORKS

1/16

REMOVE ALL BURRS AND SHARP EDGES.

NEXT ASSY

TITLE:

RUDIMENTARY LANDING GEAR WIND


TUNNEL MODEL MAIN STRUT ASSEMBLY

FRACTIONS
1

VERSION: 2012

FILENAME: UTIASAA-RLGM-010

.005

INSTITUTE FOR AEROSPACE STUDIES


UNIVERSITY OF TORONTO
4925 DUFFERIN STREET
TORONTO, ONTARIO
M3H 5T6
DO NOT SCALE DRAWING

.XXX

USED ON

Appendix C. Landing Gear Model Drawings

PROPRIETARY AND CONFIDENTIAL


THE INFORMATION CONTAINED IN
THIS DRAWING IS THE SOLE
PROPERTY OF THE UNIVESITY OF
TORONTO. ANY REPRODUCTION IN
PART OR AS A WHOLE WITHOUT
THE WRITTEN PERMISSION OF THE
UNIVERSITY OF TORONTO IS
PROHIBITED.

SIZE DWG. NO.

UTIASAA-RLGM-010

SCALE: 1:2

WT:

REV

SHEET 1 OF 1

67

1/2-20 Machine Threads


0.551

0.315

0.315

0.500

2 X M2.5x0.45 TAP

0.138

R0.118 TYP.

1.100

DETAIL A
SCALE 1 : 1

2.000
5.295
2.000

0.630

3.118

1.063

0.118 TYP.

4.173
4.528

1.100

2.126

1.282
1.717

1.260

12.795

11.795

0.197

DRAWN

NAME

DATE

RC

05/03/2013

CHECKED
UNLESS OTHERWISE SPECIFIED:
DIMENSIONS ARE IN INCHES

TOLERANCES:
.XX
.02

THIS DRAWING WAS PRODUCED USING:


SOFTWARE: SOLIDWORKS

.005

TITLE:

RUDIMENTARY LANDING GEAR


MAIN STRUT

FRACTIONS
1

1/16

REMOVE ALL BURRS AND SHARP EDGES.

NEXT ASSY

INSTITUTE FOR AEROSPACE STUDIES


UNIVERSITY OF TORONTO
4925 DUFFERIN STREET
TORONTO, ONTARIO
M3H 5T6
DO NOT SCALE DRAWING

.XXX

VERSION: 2012

FILENAME: UTIASAA-RLGM-011

Appendix C. Landing Gear Model Drawings

USED ON

PROPRIETARY AND CONFIDENTIAL


THE INFORMATION CONTAINED IN
THIS DRAWING IS THE SOLE
PROPERTY OF THE UNIVESITY OF
TORONTO. ANY REPRODUCTION IN
PART OR AS A WHOLE WITHOUT
THE WRITTEN PERMISSION OF THE
UNIVERSITY OF TORONTO IS
PROHIBITED.

SIZE DWG. NO.

UTIASAA-RLGM-011

SCALE: 1:2

WT:

REV

SHEET 1 OF 2

68

0.315

0.059

2.000

3.118

0.118 TYP.

C
D

0.315

3.898

DETAIL C
SCALE 1 : 1

SECTION B-B

R0.315

0.177 THRU

2.500

1.717

60
1.063

1.378

1.282

0.315

Appendix C. Landing Gear Model Drawings

1.292

DETAIL D
SCALE 1 : 1
0.315

M5 TAP X 2

1.149

DRAWN

NAME

DATE

RC

05/03/2013

CHECKED
UNLESS OTHERWISE SPECIFIED:
DIMENSIONS ARE IN INCHES

TOLERANCES:
.XX
.02

THIS DRAWING WAS PRODUCED USING:


SOFTWARE: SOLIDWORKS

1/16

REMOVE ALL BURRS AND SHARP EDGES.

NEXT ASSY

RUDIMENTARY LANDING GEAR


MAIN STRUT

FRACTIONS
1

VERSION: 2012

FILENAME: UTIASAA-RLGM-011

TITLE:

DO NOT SCALE DRAWING

.XXX
.005

INSTITUTE FOR AEROSPACE STUDIES


UNIVERSITY OF TORONTO
4925 DUFFERIN STREET
TORONTO, ONTARIO
M3H 5T6

USED ON

PROPRIETARY AND CONFIDENTIAL


THE INFORMATION CONTAINED IN
THIS DRAWING IS THE SOLE
PROPERTY OF THE UNIVESITY OF
TORONTO. ANY REPRODUCTION IN
PART OR AS A WHOLE WITHOUT
THE WRITTEN PERMISSION OF THE
UNIVERSITY OF TORONTO IS
PROHIBITED.

SIZE DWG. NO.

UTIASAA-RLGM-011

SCALE: 1:2

WT:

REV

SHEET 2 OF 2

69

3.630
2.646
0.059 TYP.

1.661

0.677

SECTION A-A

0.315 TYP.

R0.276

0.315

0.177 TYP.

2 X M2.5 CSK

0.128

4.528

Appendix C. Landing Gear Model Drawings

DETAIL B
SCALE 4 : 1

SECTION C-C

DRAWN

0.089 TYP.

0.030 TYP.

0.220 TYP.

05/03/2013

UNLESS OTHERWISE SPECIFIED:


DIMENSIONS ARE IN INCHES
TOLERANCES:
.XX
.02

.005

TITLE:

RUDIMENTARY LANDING GEAR


MAIN STRUT MICROPHONE INSERT

FRACTIONS
1

1/16

REMOVE ALL BURRS AND SHARP EDGES.

NEXT ASSY

INSTITUTE FOR AEROSPACE STUDIES


UNIVERSITY OF TORONTO
4925 DUFFERIN STREET
TORONTO, ONTARIO
M3H 5T6
DO NOT SCALE DRAWING

.XXX

VERSION: 2012

FILENAME: UTIASAA-RLGM-012

DATE

RC

CHECKED

THIS DRAWING WAS PRODUCED USING:


SOFTWARE: SOLIDWORKS

NAME

USED ON

PROPRIETARY AND CONFIDENTIAL


THE INFORMATION CONTAINED IN
THIS DRAWING IS THE SOLE
PROPERTY OF THE UNIVESITY OF
TORONTO. ANY REPRODUCTION IN
PART OR AS A WHOLE WITHOUT
THE WRITTEN PERMISSION OF THE
UNIVERSITY OF TORONTO IS
PROHIBITED.

SIZE DWG. NO.

UTIASAA-RLGM-012

SCALE: 2:1

WT:

REV

SHEET 1 OF 1

70

0.197 TYP.

0.374

0.295 TYP.

2 X M4 THRU

C
0.591 TYP.
0.787 TYP.
9.646

DRAWN

NAME

DATE

RC

05/03/2013

CHECKED
UNLESS OTHERWISE SPECIFIED:
DIMENSIONS ARE IN INCHES

TOLERANCES:
.XX
.02

THIS DRAWING WAS PRODUCED USING:


SOFTWARE: SOLIDWORKS

1/16

REMOVE ALL BURRS AND SHARP EDGES.

NEXT ASSY

TITLE:

RUDIMENTARY LANDING GEAR


DRAG STRUT

FRACTIONS
1

VERSION: 2012

FILENAME: UTIASAA-RLGM-020

.005

INSTITUTE FOR AEROSPACE STUDIES


UNIVERSITY OF TORONTO
4925 DUFFERIN STREET
TORONTO, ONTARIO
M3H 5T6
DO NOT SCALE DRAWING

.XXX

USED ON

Appendix C. Landing Gear Model Drawings

PROPRIETARY AND CONFIDENTIAL


THE INFORMATION CONTAINED IN
THIS DRAWING IS THE SOLE
PROPERTY OF THE UNIVESITY OF
TORONTO. ANY REPRODUCTION IN
PART OR AS A WHOLE WITHOUT
THE WRITTEN PERMISSION OF THE
UNIVERSITY OF TORONTO IS
PROHIBITED.

SIZE DWG. NO.

UTIASAA-RLGM-020

SCALE: 1:1

WT:

REV

SHEET 1 OF 1

71

1
3.543

D
2

C
4
1.181

R0.354 TYP.

B
ITEM NO.
1
2
3
4
5

PART NUMBER / DWG. NUMBER


UTIASAA-RLGM-031
UTIASAA-RLGM-032
UTIASAA-RLGM-033
B18.3.1M - 3 x 0.5 x 20 Hex SHCS -- 20NHX
EK-26899-P03
DRAWN

DESCRIPTION

QTY.

RUDIMENTARY LANDING GEAR WHEEL TREAD


RUDIMENTARY LANDING GEAR OUTER WHEEL CAP
RUDIMENTARY LANDING GEAR INNER WHEEL CAP
M3 X 0.5 X 20 SHCS
SURFACE MICROPHONES (KNOWLES)

1
1
1
3
5

NAME

DATE

RC

05/03/2013

CHECKED
UNLESS OTHERWISE SPECIFIED:
DIMENSIONS ARE IN INCHES

TOLERANCES:
.XX
.02

THIS DRAWING WAS PRODUCED USING:


SOFTWARE: SOLIDWORKS

1/16

REMOVE ALL BURRS AND SHARP EDGES.

NEXT ASSY

TITLE:

RUDIMENTARY LANDING GEAR


WHEEL ASSEMBLY

FRACTIONS
1

VERSION: 2012

FILENAME: UTIASAA-RLGM-030

.005

INSTITUTE FOR AEROSPACE STUDIES


UNIVERSITY OF TORONTO
4925 DUFFERIN STREET
TORONTO, ONTARIO
M3H 5T6
DO NOT SCALE DRAWING

.XXX

USED ON

Appendix C. Landing Gear Model Drawings

PROPRIETARY AND CONFIDENTIAL


THE INFORMATION CONTAINED IN
THIS DRAWING IS THE SOLE
PROPERTY OF THE UNIVESITY OF
TORONTO. ANY REPRODUCTION IN
PART OR AS A WHOLE WITHOUT
THE WRITTEN PERMISSION OF THE
UNIVERSITY OF TORONTO IS
PROHIBITED.

SIZE DWG. NO.

UTIASAA-RLGM-030

SCALE: 1:1

WT:

REV

SHEET 1 OF 1

72

3 X M3 THRU

2
3.543

0.307

2.362

0.089

120

0.089

0.220

45 TYP.

120
0.030 TYP.

Appendix C. Landing Gear Model Drawings

0.089 TYP.

0.177 TYP.

DETAIL A
SCALE 2 : 1

DRAWN

TOLERANCES:

.005

TITLE:

RUDIMENTARY LANDING GEAR


WHEEL TREAD

FRACTIONS
1

1/16

REMOVE ALL BURRS AND SHARP EDGES.

NEXT ASSY

INSTITUTE FOR AEROSPACE STUDIES


UNIVERSITY OF TORONTO
4925 DUFFERIN STREET
TORONTO, ONTARIO
M3H 5T6
DO NOT SCALE DRAWING

.XXX

VERSION: 2012

FILENAME: UTIASAA-RLGM-031

05/03/2013

UNLESS OTHERWISE SPECIFIED:


DIMENSIONS ARE IN INCHES

.02

THIS DRAWING WAS PRODUCED USING:

DATE

RC

CHECKED

.XX

SOFTWARE: SOLIDWORKS

NAME

USED ON

PROPRIETARY AND CONFIDENTIAL


THE INFORMATION CONTAINED IN
THIS DRAWING IS THE SOLE
PROPERTY OF THE UNIVESITY OF
TORONTO. ANY REPRODUCTION IN
PART OR AS A WHOLE WITHOUT
THE WRITTEN PERMISSION OF THE
UNIVERSITY OF TORONTO IS
PROHIBITED.

SIZE DWG. NO.

UTIASAA-RLGM-031

SCALE: 1:1

WT:

REV

SHEET 1 OF 1

73

D
R0.354

3 x M3x0.5x0.325 TAP

120
0.276

3.543

C
0.260

0.630

0.394
120

DRAWN

NAME

DATE

RC

05/03/2013

CHECKED
UNLESS OTHERWISE SPECIFIED:
DIMENSIONS ARE IN INCHES

TOLERANCES:
.XX
.02

THIS DRAWING WAS PRODUCED USING:


SOFTWARE: SOLIDWORKS

1/16

REMOVE ALL BURRS AND SHARP EDGES.

NEXT ASSY

TITLE:

RUDIMENTARY LANDING GEAR


OUTER WHEEL CAP

FRACTIONS
1

VERSION: 2012

FILENAME: UTIASAA-RLGM-032

.005

INSTITUTE FOR AEROSPACE STUDIES


UNIVERSITY OF TORONTO
4925 DUFFERIN STREET
TORONTO, ONTARIO
M3H 5T6
DO NOT SCALE DRAWING

.XXX

USED ON

Appendix C. Landing Gear Model Drawings

PROPRIETARY AND CONFIDENTIAL


THE INFORMATION CONTAINED IN
THIS DRAWING IS THE SOLE
PROPERTY OF THE UNIVESITY OF
TORONTO. ANY REPRODUCTION IN
PART OR AS A WHOLE WITHOUT
THE WRITTEN PERMISSION OF THE
UNIVERSITY OF TORONTO IS
PROHIBITED.

SIZE DWG. NO.

UTIASAA-RLGM-032

SCALE: 1:1

WT:

REV

SHEET 1 OF 1

74

3.543

R1.417

R0.354

0.260

120
0.197 THRU

0.197 TYP.

0.630

2
0.256 TYP.

0.134 TYP.

91

0.5

0.480

120

DRAWN

NAME

DATE

RC

05/03/2013

CHECKED
UNLESS OTHERWISE SPECIFIED:
DIMENSIONS ARE IN INCHES

TOLERANCES:
.XX
.02

THIS DRAWING WAS PRODUCED USING:


SOFTWARE: SOLIDWORKS

1/16

REMOVE ALL BURRS AND SHARP EDGES.

NEXT ASSY

TITLE:

RUDIMENTARY LANDING GEAR


INNER WHEEL CAP

FRACTIONS
1

VERSION: 2012

FILENAME: UTIASAA-RLGM-033

.005

INSTITUTE FOR AEROSPACE STUDIES


UNIVERSITY OF TORONTO
4925 DUFFERIN STREET
TORONTO, ONTARIO
M3H 5T6
DO NOT SCALE DRAWING

.XXX

USED ON

Appendix C. Landing Gear Model Drawings

PROPRIETARY AND CONFIDENTIAL


THE INFORMATION CONTAINED IN
THIS DRAWING IS THE SOLE
PROPERTY OF THE UNIVESITY OF
TORONTO. ANY REPRODUCTION IN
PART OR AS A WHOLE WITHOUT
THE WRITTEN PERMISSION OF THE
UNIVERSITY OF TORONTO IS
PROHIBITED.

SIZE DWG. NO.

UTIASAA-RLGM-033

SCALE: 1:1

WT:

REV

SHEET 1 OF 1

75

You might also like