Escholarship UC Item 9q75v9t9

Download as pdf or txt
Download as pdf or txt
You are on page 1of 80

UC Irvine

UC Irvine Electronic Theses and Dissertations

Title
Reduction of Tonal Propeller Noise by Means of Uneven Blade Spacing

Permalink
https://escholarship.org/uc/item/9q75v9t9

Author
Kim, Tae

Publication Date
2016

Peer reviewed|Thesis/dissertation

eScholarship.org Powered by the California Digital Library


University of California
UNIVERSITY OF CALIFORNIA,
IRVINE

Reduction of Tonal Propeller Noise by Means of Uneven Blade Spacing

THESIS

submitted in partial satisfaction of the requirements


for the degree of

MASTERS OF SCIENCE

in Mechanical and Aerospace Engineering

by

Tae Young Kim

Thesis Committee:
Professor Dimitri Papamoschou, Chair
Professor Robert H. Liebeck
Professor Feng Liu

2016
c 2016 Tae Young Kim
TABLE OF CONTENTS

Page

LIST OF FIGURES iv

LIST OF TABLES vi

ACKNOWLEDGMENTS vii

CURRICULUM VITAE viii

ABSTRACT OF THE DISSERTATION xi

1 Introduction 1
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2.1 Elements of Propeller Noise . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.2 Previous Works . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Thesis Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Theoretical Work 6
2.1 Noise Prediction of a Propeller with Even Blade Spacing . . . . . . . . . . . . . . 6
2.2 Superposition of Propellers with Even Blade Spacing . . . . . . . . . . . . . . . . 9
2.3 Results of Theoretical Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3 Propeller Design 13
3.1 McCauley Propeller . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.2 Subscale Propellers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

4 Experimental Details 19
4.1 Power Plant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.2 Data Measurement and Collection . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.2.1 Acoustic Pressure Measurement . . . . . . . . . . . . . . . . . . . . . . . 20
4.2.2 Thrust Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.2.3 RPM Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.3 Data Processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.3.1 Sound Pressure Level . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

ii
4.3.2 Overall Sound Pressure Level . . . . . . . . . . . . . . . . . . . . . . . . 24

5 Results and Analysis 26


5.1 Tip Mach Number and Disc Loading . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.2 Power Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.3 Overall Sound Pressure Level . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

6 Conclusion 43
6.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
6.2 Recommendations for future work . . . . . . . . . . . . . . . . . . . . . . . . . . 44

Bibliography 45

A Blade Element Momentum Theory 48

B Stress Analysis 54

C Unweighted vs. A-weighted Spectra 58

D Unweighted Spectra 62

E Unweighted Spectra, Compound Propellers 65

iii
LIST OF FIGURES

Page

1.1 Applications of uneven blade spacing found in the Nissan Sentra’s radiator cooling
fan (left) [22] and in the Eurocopter EC135 helicopter tail rotor (right), courtesy of
Christoph Hansa. (https://en.wikipedia.org/wiki/File:RTH Christoph Hansa 06.jpg.) 4

2.1 Distribution of monopole sources (blue dots) along the normalized chord X. H(X)
represents the normalized thickness distribution. . . . . . . . . . . . . . . . . . . . 8
2.2 Distribution of dipole sources (blue dots) along the normalized chord X. P(X)
represents the normalized net pressure distribution acting on the blade section. . . . 9
2.3 Acoustic phase shift validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.4 ∆OASPL(A) map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.5 ∆OASPL(A) for ε = 15◦ , 20◦ , 25◦ , 30◦ . . . . . . . . . . . . . . . . . . . . . . . . . 12

3.1 The latest update to the 350 series, the Beechcraft King Air 350i [4]. . . . . . . . . 14
3.2 Chord and thickness modifications. . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.3 Twist modification. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.4 Airfoil profile modification. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.5 3D drawings of propellers with separation angles: 90◦ (top), 30◦ (bottom left), 20◦
(bottom right). Drawings were generated on Solidworks. . . . . . . . . . . . . . . 17
3.6 Printed propellers with separation angles: 90◦ (left), 30◦ (center), 20◦ (right). . . . 18

4.1 Electronics block diagram. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19


4.2 Anechoic chamber schematic. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.3 Anechoic chamber setup with the ε30 propeller. . . . . . . . . . . . . . . . . . . . 21
4.4 Motor support and thrust measurement assembly. . . . . . . . . . . . . . . . . . . 22
4.5 RPM sensor block diagram, designed by Truong [33]. . . . . . . . . . . . . . . . . 23
4.6 A-weighting correction curve [2]. . . . . . . . . . . . . . . . . . . . . . . . . . . 25

5.1 Comparison of thrust (left) and disc loading (right) between theory and experiments. 27
5.2 Spectra for ε90 (top), ε30 (middle), and ε20 (bottom) with labeled tones for down-
ward noise at θ = 96.7◦ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.3 Spectra showing the increases in broadband noise in ε30 (top) and ε20 (bottom)
for downward noise at θ = 44.2◦ . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5.4 Section view of propeller showing the merging of blades near the hub. . . . . . . . 30
5.5 Compound propeller mounted onto motor for testing. . . . . . . . . . . . . . . . . 31
5.6 Compound propellers: ε30 (top) andε20 (bottom) for downward noise at θ = 44.2◦ . 31

iv
5.7 Extraction of tonal noise from power spectrum. The fixed width of 120 hz trans-
lates into 4 data points for a 4096 point-Fourier transform. . . . . . . . . . . . . . 32
5.8 Filtered spectra for downward noise at θ = 96.7◦ . Extracted tones are shaded in
red. The acoustic intensity for each tone is displayed above the tone in Pa2 . . . . . 33
5.9 Tonal acoustic intensity for downward (left) and sideline (right) noise. . . . . . . . 34
5.10 Percent difference in acoustic intensity from the ε90 propeller. . . . . . . . . . . . 34
5.11 Percent of total acoustic intensity contained in the 0.5BPF tone. . . . . . . . . . . 35
5.12 Total acoustic intensity without the 0.5BPF tone. . . . . . . . . . . . . . . . . . . 35
5.13 Changes in acoustic intensity in “regular tones.” . . . . . . . . . . . . . . . . . . . 37
5.14 Shifting of acoustic intensity into “new” tones. . . . . . . . . . . . . . . . . . . . 38
5.15 Unweighted and A-weighted spectra of ε90 (top), ε30 (middle), and ε20 (bottom)
at θ = 96.7◦ . A-weighting is performed after scaling the frequencies with the
model scale factor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.16 OASPL(A) of 0.5BPF through 5BPF tones. . . . . . . . . . . . . . . . . . . . . . 40
5.17 Change in OASPL(A) from the ε90 propeller compared to theoretical predictions. . 40
5.18 OASPL(A) from the ε90 propeller with the 5.5BPF tone. . . . . . . . . . . . . . . 41
5.19 Change in OASPL(A) from the ε90 propeller with the 5.5BPF tone. . . . . . . . . 41
5.20 Changes in OASPL(A) from the ε90 propeller with both tonal and broadband noise. 42

v
LIST OF TABLES

Page

5.1 RPM, tip Mach number, and disc loading. . . . . . . . . . . . . . . . . . . . . . . 27

vi
ACKNOWLEDGMENTS

First and foremost, I would like to thank my advisor, Professor Papamoschou. He has been men-
toring and guiding me ever since I was an undergraduate student who knew absolutely nothing. I
would also like to thank all of my friends and family members who encouraged to take the leap
and pursue a graduate education.

vii
CURRICULUM VITAE

Tae Young Kim

EDUCATION

Master of Science in Mechanical and Aerospace Engineering 2016


University of California, Irvine Irvine, California
Bachelor of Science in Mechanical and Aerospace Engineering 2014
University of California, Irvine Irvine, California

RESEARCH EXPERIENCE
Graduate Student Researcher 2014–2016
University of California, Irvine Irvine, California

TEACHING EXPERIENCE
Teaching Assistant 2015–2016
University of California, Irvine Irvine, California

viii
NOMENCLATURE

Abbreviations
BEMT Blade Element Momentum Theory
BET Blade Element Theory
BPF Blade Passage Frequency
DAQ Data Acquisition
ESC Electronic Speed Control
FAA Federal Aviation Administration
FAR Federal Aviation Regulation
FFT Fast Fourier Transform
OASPL Overall Sound Pressure Level
OASPL(A) A-Weighted Overall Sound Pressure Level
psi Pounds Per Square Inch

Symbols
A Blade Section Area
vi
a V∞
B Blade Count
C Chord
c Speed of Sound
Cd Section Drag coefficient
Cl Section lift coefficient
CLα Lift-curve slope
D Drag
E Young’s Modulus
f Frequency
F Prandtl’s tip loss function
I Area Moment of Inertia

ix
L Lift
ṁ Mass flow rate through rotor
p Static Pressure
p0 Acoustic Pressure Disturbance
PDm Drag Noise Fourier Coefficient at harmonic m
PLm Lift Noise Fourier Coefficient at harmonic m
PV m Thickness Fourier Coefficient Noise at harmonic m
q Transverse Beam Load
T Propeller Thrust
vi Induced Velocity
V∞ Freestream Velocity
w Slipstream velocity
w Beam Deflection
x, y, z Cartesian coordinate system position
α Angle of attack
αi Induced Angle of attack
β Blade pitch angle
δ Θ − αi
ε Separation Angle
η efficiency
Θ Geometric Angle of Attack
θ Polar Angle
ρ Density
σ Stress
φ Azimuth angle
Ω Angular velocity

x
ABSTRACT OF THE DISSERTATION
Reduction of Tonal Propeller Noise by Means of Uneven Blade Spacing

By

Tae Young Kim

Masters of Science in Mechanical and Aerospace Engineering

University of California, Irvine, 2016

Professor Dimitri Papamoschou, Chair

An experimental study was conducted on the noise characteristics of four-bladed general aviation
propellers with uneven blade spacing. The subscale propeller designs were inspired by the four-
bladed McCauley propellers used on the Beechcraft King Air 350 series aircraft. The 4-inch diam-
eter (1:22.5 scale) propellers were manufactured using high-resolution stereolithography and were
powered by a high performance, radio controlled brushless electric motor. Acoustic measurements
were taken with a 24-microphone array. The use of uneven blade spacing created additional tones
over which the acoustic intensity was distributed. Large amounts of acoustic intensity were shifted
into the lowest frequency tone (occuring at half of the blade passage frequency of the propeller
with evenly spaced blades), resulting in reductions of A-weighted overall sound pressure levels of
up to 5 dB for polar angles near 90◦ . These reduction are partly offset by increases in A-weighted
overall sound pressure levels of up to 4 dB at polar angles less than 50◦ . Although the theory used
to predict propeller noise does not show good agreement with experiments, it does show this trend
of increasing noise at low polar angles. Since noise at low polar angles are weighted less in noise
metrics such as flyover noise, the use of uneven blade spacing has potential for providing noise
reduction without adding excessive complexity to propeller design.

xi
Chapter 1

Introduction

1.1 Motivation

In the past few decades, the demand for air transportation in the United States has risen dramati-
cally. This demand was met by an increase in the number of both commercial and general aviation
aircraft, which in turn increased the noise pollution around airports and airport communities. Stud-
ies have shown that this increase in noise can devalue not only property [9, 25], but general health
and quality of life as well [10, 13, 32]. To combat the increase in noise pollution, the Federal
Aviation Agency introduced noise limit requirements for aircraft certification in the form of the
Federal Aviation Regulations (FAR) Part 36 [12], and efforts in identifying and reducing aircraft
noise sources have since been a topic of interest. The propeller is usually the most dominant source
of noise in propeller driven aircraft and is therefore the subject of this study.

1.2 Background

The propeller is the most efficient method of producing aerodynamic thrust at low subsonic speeds.
A simple momentum analysis neglecting compressibility effects (shown in Appendix A), shows
that the efficiency of any propulsion device is related to the ratio of the induced velocity at the
propeller plane to the freestream velocity by equation 1.1. This ratio is called the induced velocity

1
factor and is denoted by the symbol a [30]. Equation 1.1 shows that higher theoretical efficiency
is achieved by applying smaller acceleration to a larger mass of air. In the realm of low subsonic
flight, where the effects of compressibility are not prominent, the propeller is the most practical
method of imparting small acceleration to a large mass of air. Due to its high theoretical efficiency
and relative simplicity, the propeller is still widely used in general aviation and military aircraft
despite huge advances in jet engine technology.

1
ηpropulsor = (1.1)
1+a

1.2.1 Elements of Propeller Noise

Propeller noise is often categorized into three components: harmonic or tonal noise, narrow-band
noise, and broadband noise. The three components are described as follows by Magliozzi [18] and
Smith [31].
Harmonic or tonal noise is caused by periodic motion of the propeller blades and is concen-
trated in deterministic frequencies. From a stationary point of view, a rotating propeller passes by
with a frequency equal to the product of the shaft rotational frequency and the number of blades.
This frequency is called the blade passage frequency or BPF. Each time a propeller blade passes, it
generates pressure disturbances which propagate into the far-field as sound waves at the BPF and
its harmonics. The main components of tonal noise are thickness, lift, and drag induced noise.
Narrow-band noise is caused by periodic motion of the propeller blades in unsteady flow
conditions. Interaction between the flow and aircraft components such as the fuselage and mis-
alignment of the propeller axis with respect to the incoming flow are examples that can cause
unsteady blade loading. Because narrow-band noise is both deterministic and stochastic, it shows
characteristics of both and occupies a narrow range of frequencies centered around the frequencies
of tonal noise. In carefully controlled environments such as a research laboratory, narrow-band
noise should be minimal.
Broadband noise is caused by the creation of, or interactions with turbulence. Since turbu-

2
lence is stochastic in nature, broadband noise is found at all frequencies. Atmospheric turbulence
as well as turbulence created by aircraft components such as the fuselage and the propeller itself
can contribute to broadband noise. Narrow-band and broadband noise will be neglected in this
study because of the dominance of tonal noise propeller acoustics [23].

1.2.2 Previous Works

Conventional methods of reducing propeller and fan noise focused on isolating individual noise
sources and reducing the source strengths. Attempts at lowering tonal noise include using thin-
ner blade sections to lower thickness noise, and adjusting the number of blades, tip speed, blade
chord, and other shape parameters to lower lift and drag induced noise [19]. Attempts at lowering
broadband noise include suppressing the amount of turbulent vortex shedding by the blades [24].
However, these methods require significant design changes to the propellers, which tend to create
losses in efficiency that outweigh the benefits in noise reduction. One method of noise reduction
that is feasible is the use of blade sweep as shown by Wright and Simmons [34], and many modern
propellers incorporate blade sweep. However, the main purpose of blade sweep is to improve pro-
peller aerodynamics against the effects of compressibility, making noise reductions due to blade
sweep an incidental benefit.
An alternative method of reducing propeller noise is to adjust the tonal characteristics of the
propeller by unevenly spacing the blades around the hub. This preserves the geometry of the pro-
peller blades and therefore does not harm propeller performance except in extreme configurations.
The concept of using unevenly spaced blades for noise reduction was first introduced by
Mellin, et al [21], Ewald, et al [11] and Shahady, et al [29]. These studies showed that the use
of uneven blade spacing distributed the tonal noise over a larger number of tones spread across a
broader range of frequencies. This broadening of the tonal noise reduced perceived noise by giving
it a higher degree of “whiteness.”
More recent work by Dobrzynski [8] and Cattanei, et al [7] have found promising results
with uneven blade spacing, but commercial fixed-wing aircraft have not adopted the concept. Ap-

3
plications of uneven blade spacing can be found in helicopter tail rotors and in automotive cooling
fans shown in figure 1.1 to ensure passenger comfort as described by Boltezar, et al [6] and Lee, et
al [16]. The adoption of uneven blade spacing in helicopters and automobiles imply that they have
potential in reducing tonal noise and should be investigated for use in commercial aviation.

Figure 1.1: Applications of uneven blade spacing found in the Nissan Sentra’s radiator cooling fan
(left) [22] and in the Eurocopter EC135 helicopter tail rotor (right), courtesy of Christoph Hansa.
(https://en.wikipedia.org/wiki/File:RTH Christoph Hansa 06.jpg.)

1.3 Objectives

This study aims to understand and reduce the perceived noise from general aviation propellers with
uneven blade spacing. The following objectives are established:

• Demonstrate the viability of using rapid prototyping methods in subscale propeller experi-
ments.

• Prove that the noise of propellers with uneven blade spacing can be predicted using simplified
theoretical models.

• Investigate the effects of various uneven blade spacing angles on propeller noise intensity
and directivity.

4
1.4 Thesis Overview

Chapter 1: Introduction of topic of interest.

Chapter 2: Summary of the theory used to predict propeller noise.

Chapter 3: Outline of the design procedures of subscale propellers.

Chapter 4: Details the experimental facilities and instruments.

Chapter 5: Analysis of experimental results.

Chapter 6: Summary of the findings in this study.

Appendix A: Blade element momentum theory and blade stress analysis.

Appendix B: Comparison of unweighted and weighted spectra.

Appendix C: Comparison with baseline propeller unweighted spectra.

Appendix D: Comparison with baseline compound propeller unweighted spectra.

5
Chapter 2

Theoretical Work

2.1 Noise Prediction of a Propeller with Even Blade Spacing

The harmonic noise of the propeller was predicted using far field helicoidal surface theory derived
by Hanson [15]. The model theorizes that the pressure-time signal of the propeller is given by


p(t) = ∑ PmB e−imBΩt (2.1)
m=−∞

where m is the harmonic or integer multiple of the blade passage frequency, B is the number
of blades, and Ω is the angular velocity. The Fourier coefficient PmB can be expressed as the
summation of the all noise sources:

PmB = PV m + PDm + PLm + P11m + P12m + P22m (2.2)

PV m , PLm , and PDm are the Fourier coefficients of the sources due to thickness, lift, and drag re-
spectively and represent the strength of each source for the mth harmonic of the blade passage
frequency. The coefficients are related to the source strengths ΨV , ΨL , and ΨD by equations 2.3 to

6
2.5.
r̄=1
Z
PV m = K1 K2 kx2tb ΨV (kx )d r̄ (2.3)
r̄hub

r̄=1
i
Z
PLm = K1 − K2 kyCL ΨL (kx )d r̄ (2.4)
2
r̄hub

r̄=1
i
Z
PDm = K1 K2 kxCD ΨD (kx )d r̄ (2.5)
2
r̄hub

The integration is performed over the normalized radial coordinate r̄ and tb is the ratio of maximum
thickness to chord. The coefficients K1 and K2 are computed using equations 2.6 and 2.7, where L
is the observer distance from the origin, y is the observer distance from the propeller axis, D is the
propeller diameter, Mx is the flight Mach number, and θ is the polar angle.

ρc2 B sin θ exp[imB( ΩL π


c )− 2]
K1 = − (2.6)
8π Dy (1 − Mx cos θ )

 
mBr̄MT sin θ
K2 = Mr̄2 JmB (2.7)
1 − Mx cos θ

The wavenumbers kx and ky are computed using equations 2.8 and 2.9, where BD is the chord to
diameter ratio and Mr̄ is the blade section Mach number.

2mBBD MT
kx = (2.8)
Mr̄ (1 − Mx cos θ )

Mr̄2 cos θ − Mx
 
2mBBD
ky = (2.9)
r̄Mr̄ Mr̄ (1 − Mx cos θ )

The source strength for thickness noise is determined using a distribution of monopole sources.
The physical blade surfaces are replaced by a distribution of monopole sources that have the same
effect on the fluid. Using the thin blade assumption, the sources can be distributed along the
propeller blade section chord as shown in figure 2.1 rather than the upper and lower surfaces. The
effect of the source distribution is integrated over the section chord to represent the blade section

7
geometry.

Figure 2.1: Distribution of monopole sources (blue dots) along the normalized chord X. H(X)
represents the normalized thickness distribution.

The strength of the monopole source ΨV as a function of the wavenumber kx is given by


equation 2.10.
Z1/2
ΨV (kx ) = H(X)eikx X dX (2.10)
−1/2

The source strengths for the loading noise are determined using distributions of dipole sources. The
effects of the source distributions are integrated over the section chord to represent the distribution
of net pressure from the upper and lower surfaces of the blade section as shown in figure 2.2. This
distribution of net pressure can be predicted using a combination of blade element momentum
theory and the assumption that the pressure distribution along the chord is parabolic. In addition,
the noise due to drag is considered negligible on the assumption that the drag is small compared
to the lift. These assumption were also made by Hanson and resulted in good correlation between
experiment and theory [15].

8
Figure 2.2: Distribution of dipole sources (blue dots) along the normalized chord X. P(X) repre-
sents the normalized net pressure distribution acting on the blade section.

Z1/2
ΨL (kx ) = P(X)eikx X dX (2.11)
−1/2

The strength of the dipole source for lift ΨL as a function of the wavenumber kx is given by equation
2.11.
The source strengths for the noise due to shear stresses in the propeller wake are determined
using distributions of quadrupole sources. However, the calculation of quadrupole source strengths
require the knowledge of the Lighthill’s stress tensor which is beyond the scope of this study.
Therefore the quadrupole terms P11m , P12m , and P22m will be ignored.

2.2 Superposition of Propellers with Even Blade Spacing

It is important to note that equation 2.1 only applies to propellers with even blade spacing. How-
ever, a simple method described by Dobrzynski allows for the superposition of propellers with
even blade spacing to create propellers with uneven blade spacing [8]. If the noise of an evenly
spaced, two-bladed propeller is known (can be predicted by Hanson’s theory), the pressure signal
from this propeller can be added to itself with a phase shift corresponding to the desired separation
angle in the propeller with uneven blade spacing. The combined pressure signal resembles a signal

9
from an unevenly spaced, four-blade propeller and is shown in equation 2.12.

∞ ∞
p(t) = ∑ Pm(2) e−im2Ωt + ∑ Pm(2) e−im2(Ωt+ε) (2.12)
m=−∞ m=−∞

To verify that equation 2.12 produces dependable results, the solution to equation 2.12 with B = 2
and ε = 90◦ is compared to the solution to Hanson’s original theory (equation 2.1) with B = 4. All
variables except the number of blades and separation angle were identical between the two cases.
The peaks corresponding to tonal noise in figure 2.3 are identical and therefore validate equation
2.12. The regions between the peaks are a result of errors in the Fourier Transform of a finite
pressure signal and do not represent broadband noise.

120
Four−blade, evenly spaced
110 2 two−blade pairs with 90° phase lag
100

90
SPL (dB)

80

70

60

50

40

30
0 2000 4000 6000 8000 10000 12000 14000 16000 18000
Frequency (Hz)

Figure 2.3: Acoustic phase shift validation

2.3 Results of Theoretical Analysis

The method of superimposing propeller noise signals was employed to predict the A-weighted
overall sound pressure levels OASPL(A) for separation angles 10◦ < ε < 90◦ and polar angles
20◦ < θ < 160◦ . To determine the effectiveness of the uneven blade spacing, the OASPL(A) of
the propellers with uneven blade spacing were compared to those of the propeller with even blade

10
spacing by equation 2.13.

∆OASPL(A) = OASPL(A) − OASPL(A)ε=90◦ (2.13)

This quantity, ∆OASPL, represents the change in noise quantity of each propeller from the baseline
propeller with evenly spaced blades at various polar angles and is shown in figures 2.4 and 2.5.

Figure 2.4: ∆OASPL(A) map. Negative values of ∆OASPL(A) denote reductions in noise while
positive values denote increases in noise.

11
ε = 15° ε = 20°
15 15

10 10
∆ OASPL(A)

∆ OASPL(A)
5 5

0 0

−5 −5
20 40 60 80 100 120 140 160 20 40 60 80 100 120 140 160
θ° θ°

ε = 25° ε = 30°
15 15

10 10
∆ OASPL(A)

5
∆ OASPL(A) 5

0 0

−5 −5
20 40 60 80 100 120 140 160 20 40 60 80 100 120 140 160
θ° θ°

Figure 2.5: ∆OASPL(A) for ε = 15◦ , 20◦ , 25◦ , 30◦ .

The location of maximum OASPL(A) reductions vary with separation angle, but the reduc-
tions tend to be concentrated near the plane of the propeller (θ = 90◦ ). At very low and very high
polar angles, increases in OASPL(A) are found for all separation angles.

12
Chapter 3

Propeller Design

3.1 McCauley Propeller

The subscale propellers were designed based on a 4-blade McCauley Propeller (model SA1241GL)
used on the twin-engine Beechcraft King Air 350 series aircraft. The latest models feature pro-
pellers with swept blades from the Raisbeck Engineering Company. However, the propeller by
McCauley Propeller Systems used on older models, shown in figure 3.1, was chosen over the
Raisbeck propeller for its geometric simplicity [20].

13
Figure 3.1: The latest update to the 350 series, the Beechcraft King Air 350i [4].

The Pratt and Whitney PT6A engines that power the propellers each provide 550 shaft horse-
power at an RPM of 2200. This propeller and engine combination produces a disc loading of 0.23
psi and a tip Mach number of 0.77. The subscale propellers were designed to meet these two
criteria for acoustic similarity.

3.2 Subscale Propellers

The subscale baseline propeller has a diameter of 4 inches and is geometrically similar to the Mc-
Cauley propeller. It was designed using the NACA 64A410 airfoil to create the required tip speed
and disc loading at 50,000 RPM. To prevent the propeller blades from bending from axial (thrust)
loads and failing under centrifugal loads, the chord and thickness distributions were modified as
shown in figure 3.2. Final stress predictions were made using the finite element analysis tool in
Solidworks, which predicted a factor of safety of approximately 1.2 under centrifugal loads at
50,000 RPM.

14
Chord Modification
0.2

Chord (in)
0

−0.2
0 0.5 1 1.5 2
Radius (in)

Thickness Modification
6
Thickness (%)

original
4 reinforced

0
0 0.5 1 1.5 2
Radius (in)

Figure 3.2: Chord and thickness modifications.

The propeller twist distribution of the McCauley propeller is unknown, so the twist of the
subscale propeller was determined using blade element-momentum theory (described in Appendix
A). The twist distribution that produces the required disc loading at the optimum aerodynamic
efficiency was found to be a logarithmic curve shown in black in figure 3.3. This distribution was
then modified to favor higher amounts of twist towards the hub to create a more stiff structure that
is less likely to bend under axial (thrust) loads. The twist towards the tip was reduced to maintain
the desired disc loading. The modified twist distribution is shown by the red logarithmic curve in
figure 3.3.

15
Twist Modification
35
Original
Reinforced
30

Section Twist (°)


25

20

15

10
0 0.5 1 1.5 2
Radius (in)

Figure 3.3: Twist modification.

Three propeller designs were chosen to be manufactured using stereolithography. The first
design was the baseline propeller with evenly spaced blades that served as a reference for all other
propellers. The remaining two propellers were propellers with uneven blade spacing. With the
guidance of the results in Chapter 2, the propellers with uneven blade spacing were chosen to
have separation angles of 20◦ and 30◦ . All other parameters were kept constant for all propellers.
Separation angles less than 20◦ were not taken into consideration for experimental testing due to
speculations that such small angles would cause too much aerodynamic interference between the
blades and result in significant losses in efficiency. The propellers were manufactured using the
high resolution stereolithography services from Proto Labs, Inc. The details of the stereolithogra-
phy material (called Accura 60) are specified in Reference [1].
Due to the finite resolution of the manufacturing process, tolerances are limited to a minimum
thickness of 0.016 inches. To account for this limitation, the propeller leading and trailing edges,
where the blade section airfoil profile converges to a single point, were thickened to 0.016 inches.
An example of this is shown in figure 3.4.

16
Airfoil Profile Modification
0.05
original profile
modified profile

y (in) 0

−0.05
−0.1 −0.05 0 0.05 0.1
x (in)

Figure 3.4: Airfoil profile modification.

The propellers were sanded to restore its original geometry at the leading and trailing edges
and to minimize drag due to surface roughness. Final 3D drawings and pictures of printed pro-
pellers are shown in figures 3.5 and 3.5 respectively.

Figure 3.5: 3D drawings of propellers with separation angles: 90◦ (top), 30◦ (bottom left), 20◦
(bottom right). Drawings were generated on Solidworks.

17
Figure 3.6: Printed propellers with separation angles: 90◦ (left), 30◦ (center), 20◦ (right).

After testing was completed, it was discovered that the blade element momentum theory used
in the design process was creating an over prediction of thrust. This was corrected by applying the
Betz condition for minimum energy loss [27]. However, it must be noted that the propellers that
were tested experimentally were designed to meet an incorrect value of disc loading due to the
over prediction of thrust.

18
Chapter 4

Experimental Details

4.1 Power Plant

The propellers were powered by a Hoffman Magnetics EDF 70HW ”Wild Beast” DC brushless
motor, Castle Creations Pheonix 125 electronic speed controller (ESC), and a 4-cell lithium poly-
mer battery. The ”Wild Beast” is capable of handling up to 1700 Watts and 60,000 RPM. Motor
throttle controls were transmitted and received using the Spektrum DX7 2.4 GHz 7 channnel radio
transmitter and Spektrum AR6200 DSM26 receiver. The schematic of this system is shown in
figure 4.1.

Figure 4.1: Electronics block diagram.

19
4.2 Data Measurement and Collection

4.2.1 Acoustic Pressure Measurement

Acoustic pressure was measured using the Bruel & Kjaer type 4138 microphone with a frequency
response of 140 kHz and Bruel & Kjaer type 2690-A-0S4 amplifier. 24 microphones and 6
amplifiers were housed in an anechoic chamber lined with Illbruch SONEX super, 6-inch tall
polyurethane foam acoustic wedges. The chamber inlet and outlet, where the usage of tall wedges
are impractical, were lined with ILLbruck SONEX mini acoustic baffles. Microphone locations
are defined by the polar angle θ and the azimuth angle φ as shown in the schematic in figure 4.2.
Twelve microphones were arranged on the downward arm (φ = 0◦ ) and twelve on a sideline arm
(φ = 60◦ ) at polar angles of approximately 20◦ < θ < 120◦ .
To minimize the noise from the interactions between the flow and the supporting structure,
the motor was supported by an aerodynamic strut hanging down from the ceiling. All of the
sensors and electronics were located on the ceiling or high on the strut and the wires connecting to
the motor were run internally in the strut to create the cleanest environment possible.

Figure 4.2: Anechoic chamber schematic.

20
Figure 4.3: Anechoic chamber setup with the ε30 propeller.

4.2.2 Thrust Measurement

Propeller thrust was measured by a Transducer Techniques ESP-35 beam load cell with a maximum
load of 77 pounds and OMEGA Engineering DP25B-S strain gauge panel meter. The load cell
was located at the top above the strut and is oriented as shown in figure 4.4. The propeller thrust is
amplified by the strut, hinge, and arm assembly and is transmitted to the load cell. The system was
designed to allow the load cell to move closer or farther from the hinge axis, and doing so changes
the amplification factor of the force due to the change in moment arm. To minimize instrument
error from the load cell, this amplification factor was carefully chosen to amplify the expected
propeller thrust close to the load cell’s maximum load of 77 pounds.

21
Figure 4.4: Motor support and thrust measurement assembly.

4.2.3 RPM Measurement

The propeller RPM was calculated using the rate of fluctuation of the power signal being fed
into the motor. A coil of magnet wire wrapped around one of the motor leads was used to sense
the fluctuating voltage by magnetic induction. This signal was then conditioned using a voltage
divider, buffer, and low-pass filter and was processed using a fast Fourier Transform code to obtain
the frequency of the fluctuation as depicted in figure 4.5. Due to the nature and design of brushless
motors, the motor RPM is related to the fluctuation frequency by equation 4.1. The ”Wild Beast”
motor has four magnetic poles, so the RPM can be found by equation 4.2.

2 ∗ 60
RPM = Frequency ∗ ( ) (4.1)
number of poles

RPM = Frequency ∗ 30 (4.2)

22
Figure 4.5: RPM sensor block diagram, designed by Truong [33].

4.3 Data Processing

Raw data from the microphones must be processed in order to provide meaningful results. This
section will describe three important acoustic metrics used to evaluate the effectiveness of uneven
blade spacing in reducing propeller noise. These metrics are the sound pressure level (SPL), overall
sound pressure level (OASPL), and acoustic pressure.

4.3.1 Sound Pressure Level

The sound pressure level distribution, or power spectrum, represents the distribution of acoustic
energy over frequency. It is obtained using a 4096-point fast Fourier transform of the acoustic
pressure fluctuation as shown in equation 4.3, where the reference pressure pref is 20 µPa.

2
ZT
1 p(t) −i2π f t
SPLraw ( f ) = lim e dt (4.3)
T →∞ T pref
−T

The spectrum must be corrected for microphone frequency response, microphone freefield re-
sponse, and atmospheric absorption. The frequency and freefield response corrections were per-
formed based on microphone specifications provided by the manufacturer, and the atmospheric
absorption correction was performed using the relation in Bass et al. [5] based on the relative hu-

23
midity and temperature readings measured at the time of testing. The corrected spectrum, shown by
equation 4.4, represents the small scale lossless power spectrum. To obtain the full scale spectrum,
equation 4.4 is divided by the scale factor [26].

SPL( f ) = SPLraw ( f ) + CFR ( f ) + CFF ( f ) + CAA ( f ) (4.4)

CFR =Correction for microphone frequency response


CFF =Correction for microphone freefield respone
CAA =Correction for atmospheric absorption

The acoustic intensity within a given frequency band [ f1 , f2 ] is calculated by integrating the SPL
spectrum over this band. This result can be presented as the square of the raw pressure fluctuation
within this band
Zf2
2
prms = pref2
10SPL( f )/10 d f (4.5)
f1

This procedure is used to calculate the energy content of tonal noise in section 5.2.

4.3.2 Overall Sound Pressure Level

The overall sound pressure level provides a measure of the variance of the pressure signal. It is
calculated by integrating the power spectrum as shown in equation 4.6. The limit fupper refers to
the highest frequency resolved in the experiments.

fZupper
 

OASPL = 10 log10  10SPL/10 d f  (4.6)


0

However, this measure of noise does not take the human ear’s frequency response into ac-
count. Human ears are more sensitive to certain frequencies, and one way to account for this is to
“correct” the power spectrum using A-weighting. This method adds or subtracts decibels from the
power spectrum at each frequency based on the weight curve shown in figure 4.6.

24
Figure 4.6: A-weighting correction curve [2].

The A-weighted power spectrum SPL(A) can be used to calculate the A-weighted overall
sound pressure level OASPL(A).

fZupper
 

OASPL(A) = 10 log10  10SPL(A)/10 d f  (4.7)


0

25
Chapter 5

Results and Analysis

For simplicity, the subscale propellers with 20◦ and 30◦ of uneven blade spacing will be referred to
as the ε20 and ε30 propellers respectively. The baseline propeller with evenly spaced blades will
be referred to as the ε90 propeller. This chapter will focus on the analysis of acoustic intensity and
OASPL. The power spectra will be reserved for Appendices B, C, and D.

5.1 Tip Mach Number and Disc Loading

Experimental thrust and disc loading results are plotted in figure 5.1 along with predicted values
from blade element momentum theory. The theory predicts the same thrust for all separation angles
because it neglects interactions between blades. Although experiments show good agreement with
theory, the target disc loading value of 0.23 psi was not reached experimentally. During a test, the
motor shaft sheared off and destroyed the ε30 propeller in the process. The thrust and disc loading
data were only taken up to 36,600 RPM, while the acoustic data (collected separately from thrust
data) were taken up to a slightly higher RPM of around 39,000.
The tip Mach number and disc loading of the subscale propellers are listed in table 5.1. Tests
showed that the propellers with uneven blade spacing typically produce lower thrust at a given
RPM, with the ε20 propeller experiencing a rather large thrust loss of 8.5% compared to the ε90
propeller at 36,600 RPM. At first, the loss of thrust was believed to be the result of some turbulence

26
related effects, such as the ingestion of the leading blade’s trailing edge separation by the trailing
blade. But the lack of narrowband noise in the ε20 and ε30 spectra (found in appendix B) suggests
otherwise.

Propeller RPM Tip Mach Number Disc Loading at 36,600 RPM

ε90 38,800 0.60 0.103 psi

ε30 38,500 0.60 0.096 psi

ε20 39,800 0.61 0.094 psi

Target 50,000 0.77 0.23 psi at 50,000 RPM

Table 5.1: RPM, tip Mach number, and disc loading.

2.5 0.2

0.18
2
0.16
Disc loading (psi)

0.14
Thrust (lbs)

1.5
0.12

0.1
1
0.08

0.06 BEMT
0.5 ε 90
0.04 ε 30
ε 20
0 0.02
2 2.5 3 3.5 4 4.5 5 2 2.5 3 3.5 4 4.5 5
RPM 4
x 10 RPM x 10
4

Figure 5.1: Comparison of thrust (left) and disc loading (right) between theory and experiments.

Instead, the thrust loss is believed to be caused by the leading blade’s downwash. The blade
element momemtum theory that was used to determine the geometric twist distribution along the
propeller blades assumed that individual blades do not interact with each other in any way. So
the propeller blades sections have angles of attack that are designed to operate in undisturbed flow
from a rotating frame of reference. However, the adjacent blades of the ε30 and ε20 propellers

27
are so close to each other near the hub that the trailing blade gets caught in the downwash of the
leading blade. The trailing blade therefore operates with smaller angles of attack and produces
less than the expected amount of lift. This loss of lift in the trailing blade is believed to result in
the overall propeller thrust loss of up to 8.5% at 36,000 RPM and possibly more at higher speeds.
Future designs should have modified twist distributions on the trailing blades to account for the
downwash effects from the leading blades.

5.2 Power Spectra

The ε30 and ε20 propellers produce twice as many tones as the ε90 propeller, at locations in-
between the tones found in the ε90 propeller. To make it easier to refer to individual tones, the
blade passage frequency (BPF) and its harmonics of the ε90 propeller will be referred to by 1BPF,
2BPF, 3BPF, and on. Since the ε30 and ε20 propellers produce tones that fall in between these
tones, the tones of the ε30 and ε20 propellers will be referred to by 0.5BPF, 1BPF, 1.5BPF, and
on. These tones are labeled in an example shown in figure 5.2.
The tones that are not labeled in figure 5.2 are unexpected tones and do not correlate to any
type of tonal noise that the propellers should be emitting. These tones are believed to be a result of
imbalance in the propeller blades causing vibrations in the propellers and the rig.

28
100
ε 90

SPL (dB/Hz)
50

1BPF 2BPF 3BPF 4BPF 5BPF


0
0 5 10 15
Frequency (kHz)
100
ε 30
SPL (dB/Hz)

50

0.5BPF 1BPF 1.5BPF 2BPF 2.5BPF 3BPF 3.5BPF 4BPF 4.5BPF 5BPF 5.5BPF
0
0 5 10 15
Frequency (kHz)
100
ε 20
SPL (dB/Hz)

50

0.5BPF 1BPF 1.5BPF 2BPF 2.5BPF 3BPF 3.5BPF 4BPF 4.5BPF 5BPF 5.5BPF
0
0 5 10 15
Frequency (kHz)

Figure 5.2: Spectra for ε90 (top), ε30 (middle), and ε20 (bottom) with labeled tones for downward
noise at θ = 96.7◦ .

An initial inspection of the spectra showed significant increases in broadband noise in the ε30
and ε20 propellers. Example spectra showing increases in broadband noise are shown in figure 5.3
(the rest can be found in Appendix C). Since the increase in broadband noise appears to be higher
on average in the ε20 propeller than the ε30 propeller, it was hypothesized that this increase in
broadband noise is due to some unknown interaction effect regarding the close proximity of the
propeller blades. The adjacent blades in propellers with unevenly spaced blades come very close to
each other near the hub and ultimately merge together as shown in figure 5.4. This close proximity
at the propeller hub is hypothesized to result in increases in broadband noise.

29
100
ε90
ε30

SPL (dB/Hz)
80

60

40

20
0 20 40 60 80 100 120
Frequency (kHz)
100
ε90
ε20
SPL (dB/Hz)

80

60

40

20
0 20 40 60 80 100 120
Frequency (kHz)

Figure 5.3: Spectra showing the increases in broadband noise in ε30 (top) and ε20 (bottom) for
downward noise at θ = 44.2◦ .

Figure 5.4: Section view of propeller showing the merging of blades near the hub.

In efforts to prove this hypothesis, additional experiments were conducted using hobby-grade
RC propellers, as shown in figure 5.5. Two sets of two-bladed propellers (4.2x4 propellers by APC
Propellers) were mounted onto a single motor shaft in tandom. Because the two sets of propeller
blades are on separate hubs, the adjacent blades are naturally given some axial separation between
them. These propellers will be referred to as the compound propeller, since each four-bladed
propeller is formed by two separate two-bladed propellers. For example, the compound propeller

30
with 30◦ of separation will be labeled as “compound ε30” in figures.

Figure 5.5: Compound propeller mounted onto motor for testing.

The separation between the adjacent blades in the compound propellers were adjusted to
match those of the 3D printed propellers (20◦ and 30◦ ). Due to the relatively high drag on the RC
propellers however, the motor was only able to spin the compound propellers up to 29,300 RPM.
Example spectra are shown in figure 5.6 (the rest of the spectra can be found in appendix D). At
29,300 RPM, the compound propellers show no signs of increases in broadband noise.

100
compound ε90
compound ε30
SPL (dB/Hz)

80

60

40

20
0 20 40 60 80 100 120
Frequency (kHz)
100
compound ε90
compound ε20
SPL (dB/Hz)

80

60

40

20
0 20 40 60 80 100 120
Frequency (kHz)

Figure 5.6: Compound propellers: ε30 (top) andε20 (bottom) for downward noise at θ = 44.2◦ .

Future designs should incorporate some degree of axial spacing to prevent as much increases
in broadband noise as possible. To keep the following analysis of data relevant, only the tonal noise
will be considered. Tones were extracted by taking data within a fixed width of 120 hz centered

31
around peak frequencies as shown in figure 5.7. The width of 120 hz was chosen because the
half-width of 60 hz consistently gave values of SPL that were more than 10 dB less than that of the
peak. Extracted tones in figure 5.8 are shown shaded in red as an example. Tones that are thought
to be the result of propeller imbalance and rig vibrations (and therefore not predicted by theory)
were not extracted.

Figure 5.7: Extraction of tonal noise from power spectrum. The fixed width of 120 hz translates
into 4 data points for a 4096 point-Fourier transform.

32
100 43.238 ε90

SPL (dB/Hz)
2.877
80 0.744
0.285 0.276
60

40
1BPF 2BPF 3BPF 4BPF 5BPF
20
0 5 10 15
Frequency (kHz)

100 70.321 ε30


7.652
SPL (dB/Hz)

1.876 0.819
80 0.378
0.075 0.107 0.063 0.115 0.158
60

40
0.5BPF 1BPF 1.5BPF 2BPF 2.5BPF 3BPF 3.5BPF 4BPF 4.5BPF 5BPF
20
0 5 10 15
Frequency (kHz)

100 106.011 37.571 ε20


SPL (dB/Hz)

5.176
80 0.548
0.124 0.177 0.168 0.082 0.075
0.071
60

40
0.5BPF 1BPF 1.5BPF 2BPF 2.5BPF 3BPF 3.5BPF 4BPF 4.5BPF 5BPF
20
0 5 10 15
Frequency (kHz)

Figure 5.8: Filtered spectra for downward noise at θ = 96.7◦ . Extracted tones are shaded in red.
The acoustic intensity for each tone is displayed above the tone in Pa2 .

The acoustic intensity in each tone was summed and plotted in figure 5.9 for all directions.
The decrease in separation angle shifts the direction of maximum total acoustic intensity to lower
polar angles.

33
Downward Sideline
160 160
ε90 ε90
140 ε30 140 ε30
ε20 ε20
120 120
tonal (Pa2)

p2rms tonal (Pa2)


100 100

80 80

60 60
rms
p2

40 40

20 20

0 0

20 40 60 80 100 120 20 40 60 80 100 120


θ (degrees) θ (degrees)

Figure 5.9: Tonal acoustic intensity for downward (left) and sideline (right) noise.

Although the downward and sideline noise in figure 5.9 were measured simultaneously, stark
differences are seen in the two directions. This difference is believed to be caused by the strut
hanging down from the ceiling, blocking and interfering the noise traveling in the sideline direc-
tion. From this point onwards, only the downward noise will be used in the analysis since it is
believed to be less contaminated.
The ε30 and ε20 propellers appear to emit much larger amount of acoustic intensity in most
directions compared to the ε90 propeller. Very large differences (in %) are found in figure 5.10 for
polar angles below 80◦ . This large increase in acoustic intensity was unexpected since the ε30 and
ε20 propellers are identical to the ε90 propeller in blade geometry and number.

ε30
1000 ε20
Percent Increase (%)

800

600

400
rms

200
p2

20 40 60 80 100 120
θ (degrees)

Figure 5.10: Percent difference in acoustic intensity from the ε90 propeller.

34
Inspection of spectra at angles where the acoustic intensity increases were largest showed
that large fractions of the total intensity were contained in the 0.5BPF tone (the spectra shown in
figure 5.8 is a good example of this). 0.5BPF tones can often contain over 50% of the intensity of
all the tones as shown in figure 5.11.

80 ε30
Percent of p2rms in 0.5BPF tone (%) ε20
70

60

50

40

30

20

10

20 40 60 80 100 120
θ (degrees)

Figure 5.11: Percent of total acoustic intensity contained in the 0.5BPF tone.

If the intensity contained in the 0.5BPF tone is subtracted from the total intensity in figure
5.9, the result would be that shown in figure 5.12. With the 0.5BPF ignored, the tonal acoustic
intensity in the ε30 and ε20 propellers decreases in almost every direction. The use of uneven blade
spacing can cause the emission of far greater levels of tonal acoustic intensity, but the additional
intensity is highly concentrated in the lowest frequency.

70 ε90
ε30
p2rms without 0.5BPF tone (Pa2)

60 ε20

50

40

30

20

10

20 40 60 80 100 120
θ (degrees)

Figure 5.12: Total acoustic intensity without the 0.5BPF tone.

35
Because the 0.5BPF tone is at such a low frequency where the human ear is least sensitive,
large amounts of sound at these frequencies can be “acceptable.” The degree to which the human
ear’s sensitivity will play in this phenomena will be analyzed in the next section. But before moving
on, some interesting trends are shown in figures 5.13 and 5.14. Figure 5.13 shows the changes in
acoustic intensity of the “regular” tones (1BPF, 2BPF, 3BPF, ...) that are found in the ε90 propeller.
The ε30 and ε20 propellers are found to have significant decreases in acoustic intensity contained
in these “regular” tones compared to the ε90 propeller.

36
1BPF 2BPF

ε90 ε90
7 ε30
60 ε30
ε20 ε20
6
50

5
40

p2rms (Pa2)
(Pa2)

30
rms

3
p2

20
2

10
1

0 0

20 40 60 80 100 120 20 40 60 80 100 120


θ (°) θ (°)
3BPF 4BPF
0.8
ε90 0.3 ε90
ε30 ε30
0.7 ε20 ε20
0.25
0.6

0.2
0.5
p2rms (Pa2)

p2rms (Pa2)

0.4 0.15

0.3
0.1

0.2
0.05
0.1

0
20 40 60 80 100 120 20 40 60 80 100 120
θ (°) θ (°)
5BPF
0.3
ε90
ε30
ε20
0.25

0.2
prms (Pa )
2

0.15
2

0.1

0.05

0
20 40 60 80 100 120
θ (°)

Figure 5.13: Changes in acoustic intensity in “regular tones.”

This decrease in intensity can be accounted for by the creation of the 0.5BPF, 1.5BPF,
2.5BPF, ... tones. The amount of intensity shifted into these tones are shown in figure 5.14.

37
1.5BPF
0.5BPF
ε90
ε90 9 ε30
ε30
ε20
100 ε20
8

7
80
6

p2rms (Pa2)
p2rms (Pa2)

60 5

4
40
3

2
20
1

0 0

20 40 60 80 100 120 20 40 60 80 100 120


θ (°) θ (°)
2.5BPF 3.5BPF
1.2
ε90 ε90
ε30 0.3 ε30
ε20 ε20
1

0.25

0.8
0.2
p2rms (Pa2)

p2rms (Pa2)

0.6
0.15

0.4
0.1

0.2
0.05

0 0

20 40 60 80 100 120 20 40 60 80 100 120


θ (°) θ (°)
4.5BPF

ε90
ε30
0.25 ε20

0.2
prms (Pa )
2

0.15
2

0.1

0.05

20 40 60 80 100 120
θ (°)

Figure 5.14: Shifting of acoustic intensity into “new” tones.

38
5.3 Overall Sound Pressure Level

The shift of acoustic intensity into the low frequency 0.5BPF tone is beneficial because the human
ear is much less sensitive to low frequency sounds. A typical spectrum, with the first tone contain-
ing the most acoustic intensity is shown before and after A-weighting in figure 5.15. Because the
0.5BPF tone is of such a low frequency, the A-weighting tends to bring it far below the other tones.

100
unweighted ε90
SPL (dB/Hz)

A−weighted ε90

50

1BPF 2BPF 3BPF 4BPF 5BPF


0
0 5 10 15
Frequency (kHz)
100
unweighted ε30
SPL (dB/Hz)

A−weighted ε30

50

0.5BPF 1BPF 1.5BPF 2BPF 2.5BPF 3BPF 3.5BPF 4BPF 4.5BPF 5BPF 5.5BPF
0
0 5 10 15
Frequency (kHz)
100
unweighted ε20
SPL (dB/Hz)

A−weighted ε20

50

0.5BPF 1BPF 1.5BPF 2BPF 2.5BPF 3BPF 3.5BPF 4BPF 4.5BPF 5BPF 5.5BPF
0
0 5 10 15
Frequency (kHz)

Figure 5.15: Unweighted and A-weighted spectra of ε90 (top), ε30 (middle), and ε20 (bottom) at
θ = 96.7◦ . A-weighting is performed after scaling the frequencies with the model scale factor.

The OASPL(A) for the tonal noise in all directions is shown in figure 5.16. The ε20 and ε30
propellers show reductions in OASPL(A) of up to 2 and 5 decibels respectively near the plane of
the rotor (θ = 90◦ ). It is important to mention distortion of data due to thrust loss. Since the ε90
propeller produced the most thrust, it is expected to create more tonal noise. How much this affects
the OASPL(A) measurements is unknown.

39
84 ε90
ε30
83
ε20
82
81

OASPL(A) (dB)
80
79
78
77
76
75
74

20 40 60 80 100 120
θ (degrees)

Figure 5.16: OASPL(A) of 0.5BPF through 5BPF tones.

The change in OASPL(A) is shown in figure 5.17. Figure 5.17 clearly shows that the reduc-
tions in tonal noise is concentrated near the plane of the propeller. For directions behind of the
propeller (θ approaching 0◦ ), tonal noise of the ε30 and ε20 propellers are found to be higher than
that of the ε90 propeller. Comparison to theory shows that experiments are in agreement with the
general trend of decreasing OASPL(A) near θ = 90◦ and increasing OASPL(A) below θ = 50◦ .
However, levels of OASPL(A) reductions for individual points often disagree with theory with
relatively large errors.

2
∆ OASPL(A) (dB)

−2

−4
ε30 experiment
ε30 theory
−6
ε20 experiment
ε20 theory
−8
20 40 60 80 100 120 140 160
θ (degrees)

Figure 5.17: Change in OASPL(A) from the ε90 propeller compared to theoretical predictions.

The results in figure 5.17 were calculated using tones 0.5BPF through 5BPF, which are the
first five tones in the ε90 propeller and the first ten tones for the ε30 and ε20 propellers. How-

40
ever, the ε30 and ε20 spectra often show an 11th tone, the 5.5BPF tone. The 5.5BPF tones are
usually small, but because they appear at high frequencies, they can add a significant amount of
A-weighted noise. The noise reduction calculations of figures 5.16 and 5.17 were redone with the
addition of the 5.5BPF tone and are shown in figures 5.18 and 5.19. These tones appear to add no
more than 1 decibel in any direction. However, it is still significant considering that it is coming
from a single tone.

84 ε90
ε30
83
ε20
82 ε30 with 5.5BPF
81 ε20 with 5.5BPF
OASPL(A) (dB)

80
79
78
77
76
75
74

20 40 60 80 100 120
θ (degrees)

Figure 5.18: OASPL(A) from the ε90 propeller with the 5.5BPF tone.

ε30
4
ε20
3 ε30 with 5.5BPF
2 ε20 with 5.5BPF
∆ OASPL(A) (dB)

1
0
−1
−2
−3
−4
−5

20 40 60 80 100 120
θ (degrees)

Figure 5.19: Change in OASPL(A) from the ε90 propeller with the 5.5BPF tone.

The OASPL(A) values in figures 5.16, 5.17, 5.18, and 5.19 were calculated after filtering
out the broadband noise. For completeness, figure 5.20 shows the changes in OASPL(A) using
the unfiltered spectra that includes both the tonal and broadband noise. Because of the unexpected

41
increase in broadband noise, the OASPL(A) for the ε30 and ε20 propellers is noticeably higher
than before. The only reason that the ε30 and ε20 propellers are not any noisier than the ε90
propeller close to the plane of the propeller is that the rise in broadband noise is canceled out by
the decrease in tonal noise. Again, the use of axial spacing between adjacent propeller blades
should eliminate the increases in broadband noise.

6
ε30
ε20
5

4
∆ OASPL(A) (dB)

20 40 60 80 100 120
θ (degrees)

Figure 5.20: Changes in OASPL(A) from the ε90 propeller with both tonal and broadband noise.

42
Chapter 6

Conclusion

6.1 Summary

An experimental study was conducted on the noise characteristics of four-bladed general aviation
propellers with uneven blade spacing. The subscale propellers were initially designed using the ge-
ometry of the four-bladed McCauley propeller used on the Beechcraft King Air 350 series aircraft
and later modified to increase the strength of the propellers. The 4-inch diameter (1:22.5 scale)
subscale propellers were manufactured using high-resolution stereolithography and were powered
by a high performance, radio controlled brushless electric motor. In general, these experiments
have shown that it is possible to study propeller noise using 3D printed propellers. However, the
plastic material is prone to failing under centrifugal and bending loads so they must be carefully
designed to maximize structural strength.
Acoustic measurements were taken with a 24-microphone array housed inside an anechoic
chamber for polar angles between 20◦ to 120◦ . The results showed significant increases in total
acoustic intensity for propellers with uneven blade spacing, with very high concentrations of en-
ergy in the 0.5BPF tone. As predicted by the theory, the use of uneven blade spacing tends to shift
tonal noise away from the plane of the propeller. Decreases in OASPL(A) of up to 5 dB were found
near polar angles of 90◦ while increases of up to 4 dB were found for polar angles below 50◦ . This

43
is beneficial for noise metrics such as flyover noise, as noise at low polar angles are “weighted”
less than noise near the plane of the propeller. Overall, the concept of uneven blade spacing has
potential to reduce propeller tonal noise without adding excessive complexity to propeller design.

6.2 Recommendations for future work

Future work should prioritize meeting the similarity conditions. Due to the destruction of the ε30
propeller, the target tip Mach number of 0.77 and disc loading of 0.23 psi were not reached. It
is unknown whether or not the propeller blades would have been able to withstand the centrifugal
and bending loads at 50,000 RPM, but the propellers would certainly not have reached the required
disc loading due to a combination of the downwash effects described in section 5.1 as well as an
over prediction of thrust by blade element momentum theory prior to applying the Betz condition
for minimum energy loss.
Future propeller designs should also incorporate axial spacing between adjacent propeller
blades. Acoustic measurements showed that the propellers with unevenly spaced blades suffered
from increases in broadband noise significant enough to completely negate the reductions in tonal
noise. The compound propeller experiments suggest that increases in broadband noise can be
avoided by adding axial spacing between the adjacent blades. Axial spacing will also have the
added benefit of making the blades more resistant to centrifugal loads, since the merging of blades
decreases the blade section area over which centrifugal loads must be distributed.
Finally, future work should incorporate forward flight effects. The acoustic and disc loading
results shown in this study were done in static conditions because of the limitations of the facilities
used. Understanding the effects of various flight regimes such as cruise, takeoff, and landing on
propeller noise will be required to fully understand the effects of uneven blade spacing on propeller
noise.

44
Bibliography

[1] 3D Systems Corporation. Accura 60 Plastic, April 2006.


[2] R. M. Aarts. A comparison of some loudness measures for loudspeaker listening tests. The
Journal of the Audio Engineering Society, 40(3):142–146, 1992.
[3] C. N. Adkins and R. H. Liebeck. Design of optimum propellers. Journal of Propulsion and
Power, 10(5):676–683, 1994.
[4] T. Aviation. King air 350i. "http://beechcraft.txtav.com/en/king-air-350i, 2016.
[5] H. E. Bass, L. C. Sutherland, A. J. Zuckerwar, D. T. Blackstock, and D. M. Hester. Atmo-
spheric absorption of sound: Further developments. The Journal of the Acoustical Society of
America, 97, 1995.
[6] M. Boltezar, M. Mesaric, and A. Kuhelj. The influence of uneven blade spacing on the spl
and noise spectra radiated from radial fans. Journal of Sound and Vibration, 216:697–711,
1998.
[7] A. Cattanei, R. Ghio, and A. Bongiovı̀. Reduction of the tonal noise annoyance of axial flow
fans by means of optimal blade spacing. Applied Acoustics, 68:1323–1345, 2007.
[8] W. Dobrzynski. Propeller noise reduction by means of unsymmetrical blade-spacing. Journal
of Sound and Vibration, 163:123–136, December 1993.
[9] M. Espeyand and H. Lopez. The impact of airport noise and proximity on residential property
values. Growth and Change, 31:408–419, 2000.
[10] G. W. Evans, M. Bullinger, and S. Hygge. Chronic noise exposure and physiological re-
sponse: A prospective study of children living under environmental stress. Psychological
Science, 9:75–77, 1998.
[11] D. Ewald, A. Pavlovic, and J. G. Bollinger. Noise reduction by applying modulation princi-
ples. The Journal of the Acoustical Society of America, 23:1381–1385, 1970.
[12] Federal Aviation Administration. Part 36, Noise Standards: Aircraft Type and Airworthiness
Certification, January 2002.
[13] E. A. M. Franssen, C. M. A. G. van Wiechen, N. J. D. Nagelkerke, and E. Lebret. Aircraft
noise around a large international airport and its impact on general health and medication use.
Occupational and Environmental Medicine, 61:405–413, 2004.

45
[14] R. Froude. On the part played in propulsion by differences of fluid pressure. Transactions
Institude Naval Architects, 30:390, 1889.

[15] D. B. Hanson. Helicoidal surface theory for harmonic noise of propellers in the far field.
AIAA, 18:1213–1220, October 1980.

[16] J. Lee and K. Nam. Development of low-noise cooling fan using uneven fan blade spacing.
Technical Report SAE Technical Paper 2008-01-0569, Hyundai Motor Co., 2008.

[17] G. Leishmam. Principles of Helicopter Aerodynamics. Cambridge University Press, New


York, 2 edition, 2002.

[18] B. Magliozzi, D. Hanson, and R. Amiet. Aeroacoustics of flight vehicles: Theory and prac-
tice. Technical report, National Aeronautics and Space Administration, Langley Research
Center, 1991.

[19] J. E. Marte and D. W. Kurtz. A review of aerodynamic noise from propelleres, rotors, and lift
fans. Technical Report NASA-CR107568, National Aeronautics and Space Administration,
Jet Propulsion Laboratory, 1970.

[20] McCauley Propeller Systems. STC Reference Manual, May 2010.

[21] R. C. Mellin and G. Sovran. Controlling the tonal characteristics of the aerodynamic noise
generated by fan rotors. Journal of Basic Engineering, 92:143–154, 1970.

[22] MonsterAutoParts.com. Nissan sentra radiator cooling fans, 2016.

[23] A. Parry, M. Kingan, and B. Tester. Relative importance of open rotor tone and broadband
noise sources. In 17th AIAA/CEAS Aeroacoustics Conference, Portland, Oregon, 2011.

[24] R. W. Paterson, P. G. Vogt, and M. R. Fink. Vortex noise on isolated airfoils. Journal of
Aircraft, 10:296–302, 1973.

[25] G. Pennington, N. Topham, and R. Ward. Aircraft noise and residential property values
adjacent to manchester international airport. Journal of Transport Economics and Policy,
24:49–59, 1990.

[26] V. C. Phong. Intermediate pressure bleed valve noise characterization and suppression. Mas-
ter’s thesis, University of California, Irvine, 2011.

[27] L. Prandtl and A. Betz. Screw propeller with minimum energy loss. Gottinger Nachrichen,
pages 193–213, 1919.

[28] W. Rankine. On the mechanical principles of the action of propellers. Transactions Institute
Naval Architects, 6:13–39, 1865.

[29] P. A. Shahady, C. A. Lyon, J. J. Schauer, M. H. Chopin, and M. S. Ewing. The effects of


modulated blade spacing on static rotor acoustics and performance. AIAA, 73, 1973.

46
[30] R. S. Shevell. Fundamentals of Flight. Prentice-Hall, Inc., Eaglewood Cliffs, New Jersey, 2
edition, 1989.

[31] M. Smith. Aircraft Noise. Cambridge University Press, Cambridge, 1 edition, 1989.

[32] S. A. Stansfeld and M. P. Matheson. Noise pollution: Non-auditory effects on health. British
Medical Bulletin, 68:243–257, 2003.

[33] A. D. Truong. Acoustic simulation of counter-rotating open rotor noise using very small scale
model. Master’s thesis, University of California, Irvine, 2012.

[34] T. Wright and W. E. Simmons. Blade sweep for low-speed axial fans. Journal of Turboma-
chinery, 112:151–158, January 1990.

47
Appendix A

Blade Element Momentum Theory

Designing the subscale propeller requires knowledge of how much thrust the subscale propeller
will produce. The thrust must be known not only to match the disc loading of the McCauley
propeller, but to ensure that the blades will not fail under the thrust load and to accurately model
the propeller’s acoustics. The well known blade element momentum theory (BEMT) is used to
predict the forces on the propeller blades.

48
Figure A.1: Visualization of the streamtube around a propeller [17].

The momentum analysis was first done by Rankine and Froude in the analysis of marine
propellers [28, 14]. The flow is visualized in figure A.1 and is assumed to be incompressible and
isentropic everywhere except through the propeller disc. The conservation of total pressure in the
isentropic regions are written as the following

1 1
P0 + ρV∞2 = P1 + ρ[V∞ (1 + a)]2 (A.1)
2 2

1 1
P0 + ρVs2 = P2 + ρ[V∞ (1 + a)]2 (A.2)
2 2

Where a = Vv∞i and Vs = V∞ + w. Subtracting equation (A.1) from (A.2)

1
P2 − P1 = ρ(Vs2 −V∞2 ) (A.3)
2

Assuming that the pressure is constant on either side of the propeller disc, the thrust produced by

49
the propeller is
T = (∆P)A = (P2 − P1 )A (A.4)

Substituting equation (A.3) gives


1
T = ρ(Vs2 −V∞2 )A (A.5)
2

The differential thrust is therefore

1
dT = ρ(Vs2 −V∞2 )dA (A.6)
2

For a circular disc, this equation becomes

1
dT = ρ(Vs2 −V∞2 )2πrdr (A.7)
2

The thrust can also be calculated from the momentum deficit. The momentum of the fluid at the
freestream condition and the slipstream condition are

momentum in : ρV∞2 A∞ = (ρV∞ A∞ )V∞ = ṁV∞ (A.8)

momentum out : ρVs2 As = (ρVs As )Vs = ṁVs (A.9)

The thrust is the difference between equations (A.8) and (A.9)

T = ṁ(Vs −V∞ ) (A.10)

The differential thrust from this equation is

dT = d ṁ(Vs −V∞ ) = ρ(2πrdr)V∞ (1 + a)(Vs −V∞ ) (A.11)

50
Setting equations (A.7) and (A.11) equal to each other and canceling out like terms gives

1
V∞ (1 + a) = (V∞ +Vs ) (A.12)
2

Substituting equation (A.14) into equation (A.7) for Vs and simplifying gives

dT = 4ρaV∞2 (1 + a)πrdr (A.13)

dT = 4FρaV∞2 (1 + a)πrdr (A.14)

Where F is a function that corrects for the losses at the blade tip proposed by Prandtl [27]. F has
the form
2 B 1 − r̄
F= arccos(exp(− )) (A.15)
π 2 sin(δ + αi )
B 1 − r̄
f= (A.16)
2 sin(δ + αi )

Applying the Betz condition for minimum energy loss [3], the expression for f becomes

B 1 − r̄
f= (A.17)
2 sin(δtip )

Where δtip is the value of δ at the blade tip (the induced angle of attack αi is zero at the tip).
The only unknown in equation (A.14) is the inflow factor a. The inflow factor can be calculated
from the analysis of a blade element.

51
Figure A.2: Diagram of blade section.

From the geometry of the blade element with respect to the incoming flow, the differential
thrust from a single blade is

dT = dL cos(δ + αi ) − dD sin(δ + αi ) (A.18)

Multiplying by the number of blades B, and ignoring the drag and induced angle of attack αi terms
since they are small compared to the lift and angle of attack δ respectively gives

1
dT = BdL cos δ = B ρ(V∞2 + Ω2 r2 )clα (Θ − δ − αi )cdr cos δ (A.19)
2

Equating equation (A.19) to equation (A.14) and simplifying gives

B(V∞2 + Ω2 r2 )clα (Θ − δ − αi )c cos δ = 8FV∞2 a(1 + a)πr (A.20)

From the geometry, it can be shown that

vi
tan αi = p (A.21)
V∞ + Ω2 r2
2

52
Using small angle approximation,
q
vi = αi V∞2 + Ω2 r2 (A.22)

Equating the axial induced velocities from the blade element and momentum analysis gives

aU∞ = vi cos δ (A.23)

Therefore p
vi cos δ αi V∞2 + Ω2 r2 cos δ
a= = (A.24)
V∞ V∞

Substituting equation (A.24) into equation (A.20) and substituting for dimensionless variables:
√ 2 22
U∞ +Ω r V∞ Bc
V̄R = ΩR , λ = ΩR , σ = πR , r̄ = Rr

 
λ σ clα V̄R σ clα V̄R
αi2 + + α i − (Θ − δ ) = 0 (A.25)
r̄ 8F r̄2 8F r̄2

Solving this quadratic equation for the positive root gives

s
   2
1 λ σ clα V̄R 1 λ σ clα V̄R σ clα V̄R
αi = − + + + + (Θ − δ ) (A.26)
2 r̄ 8F r̄2 4 r̄ 8F r̄ 2 8F r̄2

Equation (A.26) shows that the induced angle of attack αi is a function of αi (since F is a function
of αi ). So equation (A.26) is iterated until the value for αi converges. Once αi is known, equation
(A.19) can be used to calculate the differential thrust.

53
Appendix B

Stress Analysis

There are two main sources of stress on a propeller blade. The first is the centripetal force, which
is applied in the radial direction as the propeller spins. The second is thrust distribution across the
blade, which is applied axially.
The vast majority of stresses applied to propeller blades are due to centripetal forces from
the rotation of the propeller blades. A simplified method of estimating stresses due to centripetal
forces is crucial in designing propellers for structural integrity. The centripetal force at any point
on the blade is described by
mv2
Fc = = mΩ2 r (B.1)
r

However, the radial coordinate (r) and the mass (m) supported at each radial coordinate varies
along the propeller blade. So the blade is divided into n radial segments (blade sections) and the
centripetal force is calculated at each blade section.

n
Fc (r) = Ω2 rn ∑ mn (B.2)
n=1

The centripetal force at each blade section is divided by the blade cross sectional area (A) to find
the stress due to centripetal force (σc ) at each blade section. The area of the blade section will be

54
discussed in the thrust contribution section.

σc (r) = Fc (r)/A(r) (B.3)

While the centripetal force acts radially and wants to pull the blade apart, the thrust force acts
axially and wants to bend the blade. This bending is modeled using Euler-Bernoulli beam theory.
To apply this beam theory, the propeller blade is simplified as a constant chord, straight blade with
no twist. The worst case scenario is applied by using the smallest geometric angle of attack found
in the real propeller blade.
Prior to applying the beam theory, a method of calculating the blade section area, centroid
locations (cy , cz ), and the second moment of inertias (Iy , Iz , Iyz ). Figure B.1 shows a blade cross
section. The complex geometry of the airfoil shape can be converted into a polygon by connecting
points along the upper and lower surfaces of the airfoil.

Figure B.1: Polygon Method. Only 8 vertices are shown for simplicity. More vertices are required
to more accurately represent the airfoil using a polygon.

The area, centroid, and moment of inertias of a polygon can be found using the vertices of

55
the polygon by the following equations (index i = N + 1 refers to the i = 1):

1 N
A= ∑ (yizi+1 − yi+1zi)
2 i=1
1 N
cy = ∑ (yi + yi+1)(yizi+1 − yi+1zi)
6A i=1
1 N
cz = ∑ (zi + zi+1)(yizi+1 − yi+1zi)
6A i=1
(B.4)
1 N 2
Iy = ∑ (zi + zizi+1 + z2i+1)(yizi+1 − yi+1zi)
12 i=1
1 N 2
Iz = ∑ (yi + yiyi+1 + y2i+1)(yizi+1 − yi+1zi)
12 i=1
1 N
Iyz = ∑ (yizi+1 + 2yizi + 2yi+1zi+1 + yi+1zi)(yizi+1 − yi+1zi)
12 i=1

These equations can be applied to each blade section to find the radial distributions of area, centroid
locations, and the second moment of inertias. These distributions must be known to solve the
Euler-Bernoulli beam equation.
The Euler-Bernoulli beam bending theory states that the beam deflection (w) is related to
the Young’s Modulus (E), second moment of inertia (I), and axial load distribution (T ) by the
following fourth-order differential equation:

d2 d2w
(EI 2 ) = T (x) (B.5)
dx2 dx

If the beam is initially straight, maintains the same cross section throughout the beam, and has
homogeneous material properties, equation (B.5) can be simplified into equation (B.6), where E
and I are constants.
d4w
EI = T (x) (B.6)
dx4

The bending moment of the beam can be expressed in terms of the beam deflection and is given by

d2w
M(x) = −EI (B.7)
dx2

56
Combining equations (B.6) and (B.7) gives the following expression for the bending moment

1
Z Z
M(x) = −EI T (x)dx (B.8)
EI

T (x) is the axial load distribution on the beam, which in this case, is the thrust distribution along the
blade found from Blade Element Momemtum theory. Solving equation (B.8) for a real propeller
with varying chord lengths and twist along the blade is difficult as E and I are not constant. To
simplify the analysis, the bending moment M is found for a simple propeller blade with constant
chord length and no twist. To simulate the worst case scenario, the simplified propeller blade has
the chord length and geometric angle of attack found at tip cross section of the real propeller. With
this simplification, equation (B.8) simplifies into equation (B.9) and can be easily integrated.

Z Z
M(x) = − T (x)dx (B.9)

In general, the stress at any point in the beam cross section at a specific x-coordinate value is

Mz Iy + My Iyz My Iz + Mz Iyz
σx (y, z) = − 2
(y − cy ) + 2
(z − cz ) (B.10)
Iy Iz − Iyz Iy Iz − Iyz

In equation B.10, the second moment of inertias vary along x as the propeller blade twists and
varies in chord length. It is important to note that the stress due to the thrust distribution is a
function of x, y, and z while the stress due to the centripetal force is only a function of x. Adding
the two stresses at each point in the propeller blade gives the total stress.
The bending moment causes the highest amounts of stress when the values y − cy and z − cz
are maximized. Therefore, the maximum stress will be found along the edge of the blade cross
section, or along the surface of the propeller blade.

57
Appendix C

Unweighted vs. A-weighted Spectra

Downward, θ = 24.38° Downward, θ = 29.52°


100 100
unweighted ε90 unweighted ε90
80 A−weighted ε90 80 A−weighted ε90
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)
Downward, θ = 34.68° Downward, θ = 39.7°
100 100
unweighted ε90 unweighted ε90
80 A−weighted ε90 80 A−weighted ε90
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)
Downward, θ = 44.2° Downward, θ = 52.66°
100 100
unweighted ε90 unweighted ε90
80 A−weighted ε90 80 A−weighted ε90
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B

0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)

Figure C.1: Unweighted vs. A-weighted spectra of the ε90 propeller.

58
Downward, θ = 66.18° Downward, θ = 80.79°
100 100
unweighted ε90 unweighted ε90
80 A−weighted ε90 80 A−weighted ε90
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)
Downward, θ = 96.69° Downward, θ = 113.01°
100 100
unweighted ε90 unweighted ε90
80 A−weighted ε90 80 A−weighted ε90
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)

Figure C.2: Unweighted vs. A-weighted spectra of the ε90 propeller (continued).

59
Downward, θ = 24.38° Downward, θ = 29.52°
100 100
unweighted ε30 unweighted ε30
80 A−weighted ε30 80 A−weighted ε30
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0.

1.

2.

3.

4.

5.

0.

1.

2.

3.

4.

5.
0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)
Downward, θ = 34.68° Downward, θ = 39.7°
100 100
unweighted ε30 unweighted ε30
80 A−weighted ε30 80 A−weighted ε30
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0.

1.

2.

3.

4.

5.

0.

1.

2.

3.

4.

5.
0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)
Downward, θ = 44.2° Downward, θ = 52.66°
100 100
unweighted ε30 unweighted ε30
80 A−weighted ε30 80 A−weighted ε30
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0.

1.

2.

3.

4.

5.

0.

1.

2.

3.

4.

5.
0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)
Downward, θ = 66.18° Downward, θ = 80.79°
100 100
unweighted ε30 unweighted ε30
80 A−weighted ε30 80 A−weighted ε30
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0.

1.

2.

3.

4.

5.

0.

1.

2.

3.

4.

5.
0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)
Downward, θ = 96.69° Downward, θ = 113.01°
100 100
unweighted ε30 unweighted ε30
80 A−weighted ε30 80 A−weighted ε30
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0.

1.

2.

3.

4.

5.

0.

1.

2.

3.

4.

5.

0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)

Figure C.3: Unweighted vs. A-weighted spectra of the ε30 propeller.

60
Downward, θ = 24.38° Downward, θ = 29.52°
100 100
unweighted ε20 unweighted ε20
80 A−weighted ε20 80 A−weighted ε20
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0.

1.

2.

3.

4.

5.

0.

1.

2.

3.

4.

5.
0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)
Downward, θ = 34.68° Downward, θ = 39.7°
100 100
unweighted ε20 unweighted ε20
80 A−weighted ε20 80 A−weighted ε20
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0.

1.

2.

3.

4.

5.

0.

1.

2.

3.

4.

5.
0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)
Downward, θ = 44.2° Downward, θ = 52.66°
100 100
unweighted ε20 unweighted ε20
80 A−weighted ε20 80 A−weighted ε20
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0.

1.

2.

3.

4.

5.

0.

1.

2.

3.

4.

5.
0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)
Downward, θ = 66.18° Downward, θ = 80.79°
100 100
unweighted ε20 unweighted ε20
80 A−weighted ε20 80 A−weighted ε20
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0.

1.

2.

3.

4.

5.

0.

1.

2.

3.

4.

5.
0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)
Downward, θ = 96.69° Downward, θ = 113.01°
100 100
unweighted ε20 unweighted ε20
80 A−weighted ε20 80 A−weighted ε20
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
PF

PF

PF

PF

PF

PF

PF

PF

PF

PF
5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B

5B
1B

2B

3B

4B

5B

1B

2B

3B

4B

5B
0.

1.

2.

3.

4.

5.

0.

1.

2.

3.

4.

5.

0 0
0 5 10 15 0 5 10 15
Frequency (kHz) Frequency (kHz)

Figure C.4: Unweighted vs. A-weighted spectra of the ε20 propeller.

61
Appendix D

Unweighted Spectra

Downward, θ = 24.38° Downward, θ = 29.52°


100 100
ε90 ε90
80 ε30 80 ε30
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 34.68° Downward, θ = 39.7°
100 100
ε90 ε90
80 ε30 80 ε30
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 44.2° Downward, θ = 52.66°
100 100
ε90 ε90
80 ε30 80 ε30
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)

Figure D.1: ε90 vs ε30 unweighted Spectra.

62
Downward, θ = 66.18° Downward, θ = 80.79°
100 100
ε90 ε90
80 ε30 80 ε30
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 96.69° Downward, θ = 113.01°
100 100
ε90 ε90
80 ε30 80 ε30
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)

Figure D.2: ε90 vs ε30 unweighted Spectra (continued).

63
Downward, θ = 24.38° Downward, θ = 29.52°
100 100
ε90 ε90
80 ε20 80 ε20
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 34.68° Downward, θ = 39.7°
100 100
ε90 ε90
80 ε20 80 ε20
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 44.2° Downward, θ = 52.66°
100 100
ε90 ε90
80 ε20 80 ε20
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 66.18° Downward, θ = 80.79°
100 100
ε90 ε90
80 ε20 80 ε20
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 96.69° Downward, θ = 113.01°
100 100
ε90 ε90
80 ε20 80 ε20
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)

Figure D.3: ε90 vs ε20 unweighted Spectra.

64
Appendix E

Unweighted Spectra, Compound Propellers

Downward, θ = 24.38° Downward, θ = 29.52°


100 100
compound ε90 compound ε90
80 compound ε30 80 compound ε30
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 34.68° Downward, θ = 39.7°
100 100
compound ε90 compound ε90
80 compound ε30 80 compound ε30
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 44.2° Downward, θ = 52.66°
100 100
compound ε90 compound ε90
80 compound ε30 80 compound ε30
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)

Figure E.1: Compound ε90 vs compound ε30 unweighted Spectra.

65
Downward, θ = 66.18° Downward, θ = 80.79°
100 100
compound ε90 compound ε90
80 compound ε30 80 compound ε30
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 96.69° Downward, θ = 113.01°
100 100
compound ε90 compound ε90
80 compound ε30 80 compound ε30
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)

Figure E.2: Compound ε90 vs compound ε30 unweighted Spectra (continued).

66
Downward, θ = 24.38° Downward, θ = 29.52°
100 100
compound ε90 compound ε90
80 compound ε20 80 compound ε20
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 34.68° Downward, θ = 39.7°
100 100
compound ε90 compound ε90
80 compound ε20 80 compound ε20
SPL (dB/Hz)

SPL (dB/Hz)
60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 44.2° Downward, θ = 52.66°
100 100
compound ε90 compound ε90
80 compound ε20 80 compound ε20
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 66.18° Downward, θ = 80.79°
100 100
compound ε90 compound ε90
80 compound ε20 80 compound ε20
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)
Downward, θ = 96.69° Downward, θ = 113.01°
100 100
compound ε90 compound ε90
80 compound ε20 80 compound ε20
SPL (dB/Hz)

SPL (dB/Hz)

60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Frequency (kHz) Frequency (kHz)

Figure E.3: Compound ε90 vs compound ε20 unweighted Spectra.

67

You might also like