Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

AIAA JOURNAL

Vol. 56, No. 7, July 2018

Multirotor Drone Noise at Static Thrust


Charles E. Tinney∗
University of Texas at Austin, Austin, Texas 78713
and
Jayant Sirohi†
University of Texas at Austin, Austin, Texas 78712
DOI: 10.2514/1.J056827
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

A first principles understanding of the sound field produced by multirotor drones in hover is presented. Propeller
diameters ranging from 8 to 12 in. are examined and with configurations comprising an isolated rotor, quadcopter,
and hexacopter configuration. The drone pitch, defined as the ratio of drone diameter to rotor diameter, is the same
for all multirotor configurations and is valued at 2.25. A six-degree-of-freedom load cell is used to assess the
aerodynamic performance of each configuration, whereas an azimuthal array of 1∕2 in. microphones, placed between
two and three hub-center diameters from the drone center, is used to assess the acoustic near field. The analysis is
performed using standard statistical metrics such as sound pressure level and overall sound pressure level and is
presented to demonstrate the relationship between the number of rotors, the drone rotor size, and its aerodynamic
performance (thrust) relative to the near-field acoustics.

Nomenclature ϵp = precision error


Acs = airfoil cross-section area ϵRSS = root sum of squares errors
A = propeller disk area θ = polar coordinate
a = acoustic velocity λ = drone pitch
CT = thrust coefficient ξ = propeller sweep
Cτ = torque coefficient ρ = gas density
c = propeller chord ζ = time delay
D = drone diameter τ = torque
d = propeller diameter ϕ = phase
E = wavelet energy density φ = radial coordinate measured from drone center
Fi = force ψ = mother wavelet
FM = figure of merit Ω = motor rotation speed
f = frequency Ωb = blade-pass frequency, passes per second
fs = sampling frequency ω = time per rotation, s
f = f∕Ωb , nondimensional frequency
G = premultiplied spectra Subscripts
l = wavelet time scale
x, y, z = Cartesian coordinates
Mtip = propeller tip Mach number
∞ = ambient condition
Mi = moment
* = complex conjugate
N = data partition size
n = number of propellers
OASPL = overall sound pressure level
P, p = pressure I. Introduction
p~ = wavelet coefficient of p
PSD
R
=
=
power spectral density
propeller radius
S MALL-SCALE unmanned air vehicles (UAVs) using multirotor
propulsion systems have become increasingly popular in recent
years due to their affordability and versatility. Commonly called
r = radial coordinate measured from propeller root “drones”, these UAVs provide a stable airborne platform on which a
SPL = sound pressure level variety of equipment can be mounted, such as cameras and ultrasonic
T = temperature sensors. These multirotor vehicles are available in a wide range of
t = propeller thickness, time sizes, with the smallest drones for hobbyists measuring less than 2 in.
β = propeller twist distribution, deg in diameter, whereas the largest commercially available drones can be
γ = specific heat ratio in excess of 5 ft. New and useful applications for drone vehicles
δf = frequency increment unfold daily, which in turn demands a range of propulsive
δθ = polar coordinate increment requirements, with configurations ranging from quadcopters to larger
ϵb = bias error octocopter shapes.
The performance of the propeller blade is undoubtedly important
to the design and efficiency of the vehicle because it affects the
Received 16 October 2017; revision received 22 January 2018; accepted for allowable size of its payload, its maneuverability, and ultimately
publication 28 January 2018; published online 6 April 2018. Copyright loiter time. Some efforts to characterize propeller performance and to
© 2018 by the authors. Published by the American Institute of Aeronautics and improve propeller design optimization tools can be seen in the work
Astronautics, Inc., with permission. All requests for copying and permission to of Deters and Selig [1] and Ol et al. [2]; the task of studying all
reprint should be submitted to CCC at www.copyright.com; employ the available configurations is daunting. A secondary effect of the
ISSN 0001-1452 (print) or 1533-385X (online) to initiate your request. See also
AIAA Rights and Permissions www.aiaa.org/randp. propeller design, which has received considerably less attention, is
*Research Associate, Applied Research Laboratories; cetinney@utexas. the acoustic signature that it produces. Given the growing demand for
edu. Associate Fellow AIAA (Corresponding Author). drone related tasks and the proximity of these activities to populated
† areas, the sound issue is becoming a topic of broad importance. It is
Associate Professor, Center for Mechanics of Solids, Structures and
Materials. Member AIAA. for this reason that the current study was performed.
2816
TINNEY AND SIROHI 2817

Here, we will show the results from a test campaign designed to A. Multirotor Drone and Electrical System
develop a basic understanding for the acoustic signatures produced The multirotor drone was designed to mimic many of the off-the-
by multirotor (quadcopter and hexacopter) drone configurations shelf configurations that are available to hobbyists and enthusiasts
operating under static thrust conditions. Variables of primary interest alike. These commonly fall into three categories: quadcopter (four
are the size and rotational speed of the propeller blades as well as the propellers), hexacopter (six propellers), and octocopter (eight
number of propeller blades. propellers) configurations; the former appear to be more popular (and
less expensive) but have a limited payload capacity. Therefore, the
idea was to distribute an array of motors/propellers (four, six, and
eventually eight) azimuthally with equidistant spacing so that the
II. Experimental Hardware and Setup drone pitch λ (the ratio of the drone diameter D to propeller diameter
To minimize interference from background noise, the experiments d) could be fixed for a range of propeller diameters of practical
were performed in the anechoic chamber at the J. J. Pickle Research interest. Some design constraints must be considered of course. For
Campus of the University of Texas at Austin (UT Austin). example, if all rotor disks are at the same p plane and are
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

Descriptions of this facility are provided by Mula et al. [3] and Fiévet nonoverlapping, then d ≤ D sinπ∕4  D∕ 2 for a quadcopter,
et al. [4], with the current setup being configured in a quasi-hemi-
dp ≤ 
D sinπ∕6  D∕2 for a hexacopter, and d ≤ D sinπ∕8 
anechoic arrangement. That is, only the ceiling of the anechoic p
chamber received full anechoic treatment (melamine foam wedges D 2 − 2∕2 for an octocopter. With this in mind, a drone pitch of
with air cavity followed by 5.5 in. of recycled cotton-fiber insulation), λ  D∕d  2.25 was chosen, which was not only within the
whereas the foam wedges on all four walls were removed to increase constraints of the D∕d limits for the quadcopter and hexacopter
air flow to the rotor; the floor was exposed concrete with one set of setups but appeared to reflect many of the commercially available
wedges located beneath the microphone array. The resulting interior drone kits.
dimensions (from wedge tips to walls) measured 18.5 ft (width) The structural frame for the multirotor drone was fabricated in
×22.5 ft (length) ×14 ft (height). house by combining aluminum plates with lightweight miniaturized
The three primary pieces of hardware used in this study were a extruded aluminum rails. The aluminum plates were machined to
support structure with a six-degree-of-freedom load cell (for accommodate quadcopter, hexacopter, and octocopter configura-
tions, whereas the extruded aluminum rails formed stiff adjustable
measuring the aerodynamic performance of the drone), a microphone
support arms that were extended to increase D for larger diameter
array (to capture the near-field acoustic signatures), and a multirotor
propellers. An image of the fully assembled hexacopter is provided in
drone with relevant power and electrical systems. Unlike other
Fig. 1a alongside a close-up of one of the adjustable arms in Fig. 2b.
studies that use fully assembled off-the-shelf drone kits, a custom
Each propeller was driven directly (without gears) by its own
fabricated multirotor drone test stand was built for this endeavor,
dedicated motor such that the angular position of each propeller was
which resulted in a relatively more generic set of hardware that different for each test. The motors are a brushless outrunner type with
provided additional degrees of freedom in the test matrix. An 13 circumferentially distributed magnets capable of handling up to
illustration of the quadcopter configuration undergoing testing is 229 W of power (∼11 V dc). As seen in Fig. 2b, each motor is
provided in Fig. 1 identifying some of these major hardware connected to a dedicated speed controller that can transmit up to 36 A
components. A description of these three components follows. of direct current from the power supply to the motor. Electrical power is
supplied by a 10 kW (maximum) Lambda TKE ESS 50-200
programmable dc power supply that outputs up to 50 Vat 200 A. Power
from the Lambda unit is transmitted to the drone test stand using 2
American wire gauge (AWG) welders cable to reduce electrical
transmission losses; 6 AWG was used during earlier stages of the test
program and proved to be problematic. The speed of each motor was
monitored and controlled by a closed-loop controller via a 1∕rev
magnetic sensor on each motor. This worked by changing the pulse
width of the modulated signal that the speed controllers used to govern
power to the motors. The 1∕rev sensor was also used for aligning
acoustic data during the isolated propeller tests described later.
There are numerous propeller blades to choose from for propelling
a drone, which will have a profound influence on both its
aerodynamic and aeroacoustic performance [5]. The propellers
chosen here were models 8 × 4.5 MRP, 9 × 4.5 MRP,
10 × 4.5 MRP, 11 × 4.5 MRP, and 12 × 4.5 MRP, manufac-
tured by APC Propellers. These self-tightening propellers are
intended for multirotor vehicle applications because they are
manufactured in both standard and reverse screw configurations.
Fig. 1 Quadcopter during testing with 8-in.-diam propeller blades. An image of the 8 and 12-in.-diam propellers is provided in Fig. 3.

a) b)
Fig. 2 Photographs of a) the hexacopter with 8-in.-diam propellers, and b) support arms and motor mounting configuration.
2818 TINNEY AND SIROHI

Fig. 3 The 8 and 12-in.-diam propellers manufactured by APC


Propellers.
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

Various geometrical properties are shown in Fig. 4 as provided by the


manufacturer. Here, β is the twist distribution, ξ is the blade sweep, Fig. 5 Isolated propeller with a 12-in.-diam propeller.
and Acs is the airfoil cross-section area.
so half of the propellers were spinning clockwise, whereas the other
B. Loads Measurement Apparatus
half were spinning counterclockwise.
All loads were measured by a six-component load cell (ATI The total error in the measured thrust was determined by the root
Industrial Automation Mini40E), sampled at 10 kHz using a National sum of squares errors (ϵRSS ) of the bias (ϵb ) and precision (ϵp ) errors
Instruments PXI system. The electronics and data acquisition system of the Fz measurements, as given by
are similar to the setup used by Cameron et al. [6] to study coaxial
counter-rotating rotors during hover. This strain-gauge-based load q
cell can measure three orthogonal forces (Fi  Fx , Fy , Fz ) and three ϵRSS  ϵ2b  ϵ2p (1)
orthogonal moments (Mi  Mx , My , Mz ). The load cell was
manufacturer calibrated to a full-scale Fz of 15 lbf and a full-scale Bias errors were obtained from the manufacturer specified
torque Mz of 10 lbf·in. The drone and load cell assembly was elevated calibration sheet for the load cell. Although the manufacturer
to the center of the anechoic chamber using a stiff support structure. specifies a maximum, or worst-case bias error of 0.75% of full-scale
This positioned the rotor disk plane at 100 in. from the concrete floor for the Fz load, the calibration sheet indicates the actual bias error for
below to suppress ground effects; see Mula et al. [7] for a description the range of loads encountered in these experiments to be less than
of this rotor test stand. Two separate setups were tested. The first 0.06% full scale. Therefore, a realistic estimate of the bias error is
comprised isolated propellers in which a propeller and motor taken to be 0.06% of the full-scale 15 lbf load, which is 0.009 lbf. The
combination was mounted on a long cylindrical support, which was
mean and standard deviation of the loads at each test condition were
then attached to the load cell. The cylindrical support was similar in
calculated from a set of 25,000 samples. Precision error is taken as the
diameter to the motor diameter to minimize wake interference effects.
standard deviation of these Fz samples. The largest standard
An illustration of this setup is shown in Fig. 5. The process was
deviation is 0.05 lbf, which is taken as a conservative estimate of the
repeated for different diameter propellers and with propeller thrust
and torque being taken as the Fz and Mz force and moment acting on precision error in thrust. The resulting total estimated error in the
the load cell, respectively. thrust Fz using Eq. (1) is 0.0508 lbf.
The second setup comprised the multirotor drone in which the long Thrust coefficient CT , torque coefficient Cτ , and the rotor’s figure
cylindrical support was removed so that the entire vehicle could be of merit (FM) were then calculated for each test case using the
mounted directly to the load cell. Like the isolated propeller study, following well-known definitions:
vehicle thrust was measured as Fz and vehicle torque was measured
Fz
as Mz . Note that the propellers in the multirotor drone experiments CT  (2)
were configured to cancel torque (or was nearly zero in all cases), and nρAΩR2

0.25 40

0.2
30
0.15
20
0.1
β

10
0.05

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
a) b)
0.3 0.3
0.25 0.25
0.2 0.2
0.15 0.15
0.1 0.1
0.05 0.05
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
c) d)
Fig. 4 Representations of a) propeller chord, b) twist distribution, c) thickness ratio and maximum thickness distributions, and d) sweep and cross-
section area distributions.
TINNEY AND SIROHI 2819

τ Table 1 Multirotor drone testing configurations and


Cτ  (3) position relative to the arc array
nρAΩR2 R
D, in. d, in. D∕d φ∕D 2φ∕D  d
p
C3∕2
T ∕ 2
18.00 8.0 2.25 2.42 3.35
FM  (4) 20.25 9.0 2.25 2.15 2.97
Cτ 22.50 10.0 2.25 1.93 2.68
24.75 11.0 2.25 1.76 2.43
where τ is the measured torque, R  d∕2 is the propeller radius, 27.00 12.0 2.25 1.61 2.23
A  πd2 ∕4 is the disk area, Ω is the motor rotation speed in rad∕s,
and n is the number of propellers. For the isolated propeller case, n is
valued at 1.0.
and propeller diameter. Only two propeller sizes were tested in the
hexacopter configuration. However, the forthcoming analysis will
C. Microphone Array
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

show how the trends from the quadcopter and hexacopter


Given the dynamic range of the sound levels produced by these configurations can be extrapolated for different propeller diameters.
kinds of vehicles, as well as the bandwidths of interest, four G.R.A.S.
IEPE-powered 1∕2 in. free-field microphones were selected for
measuring the sound field. These microphones (model 46AE) with
matching preamplifiers (model 26CA) have a frequency response
III. Results
range of 5 Hz to 10 kHz (1 dB accuracy) or 3.15 Hz to 20 kHz A. Aerodynamic Performance
(2 dB accuracy) as well as a dynamic range of 17 dB(A) to 138 dB Measurements were acquired over a duration of several
with nominally 50 mV∕Pa sensitivity. To improve accuracy, both the consecutive days with the drone operating in static hover conditions
microphone capsules and preamplifiers were calibrated by the only. Though atmospheric properties were not recorded, it is assumed
manufacturer as one unit. IEPE power was provided by a National that they are near standard sea-level conditions. Therefore, P∞ 
Instruments PXI system (NI-PXI-4472 board), which also filtered 14.696 psia (101,325 Pa), T ∞  525.6 R (292 K), γ  1.3991, and
(low pass Butterworth set to 84% of the Nyquist frequency) and ρ  2.34 × 10−3 slug∕ft3 (1.2082 kg∕m3 ) so that the sound speed of
digitized their voltages using dedicated 24-bit accurate converters. air is valued at a∞  1; 122 ft∕s (342 m∕s). An initial set of
The acquisition rate was set to 40 kHz for an uninterrupted duration of measurements focused on gathering aerodynamic performance data
20.48 s at each test condition. Digitized signals were converted to for the isolated propeller using the configuration shown in Fig. 5.
engineering units and partitioned into 100 data blocks comprising 213 Doing so provided a base set of measurements without the effect of
data points per block. neighboring propellers. These measurements were conducted for
Both an arc array and a line array of microphone measurement motor speeds up to the maximum allowable power (within the
points were used in this study, with each array being in line with one thermal limits of the motor).
rotor (not split between two). An illustration of this is provided in Thrust and torque coefficients corresponding to all five propeller
Fig. 6 with coordinate system. The arc array is such that 0 deg is at the diameters are shown in Figs. 7 and 8 for a range of rotor speeds (and
rotor disk plane, with positive angles θj being measured in the propeller tip Mach numbers, defined as ΩR∕a∞ ). Overall, the trends
direction of the thrust vector; a description of the line array is are as expected, with differences (between propeller diameters) being
provided in Sec. III.B.2. For the arc array, eight unique measurement attributed to Reynolds number effects; similar findings were reported
points were selected, with microphones focused on the drone center. by Brandt and Selig [8] for the same propeller shapes. When plotted
The first four points, further referred to as j  1; : : : ; 4, covered the using the propeller tip Mach number, a clear barrier forms for all
locations at −45, −30, −15, and 0 deg, whereas the second, further propeller diameters just beyond Mtip of 0.3. It is postulated that the
referred to as j  5; : : : ; 8, covered locations at 7.5, 15, 22.5, and motors’ maximum allowable safe amperage draw is limited by the
30 deg. Preliminary measurements revealed increased sound pressure elevated drag levels that accompany the onset of compressibility
levels above the rotor disk plane, and so the positive angles were effects. Turning one’s attention to Fig. 8b, the figure of merit is shown
designed to capture a finer grid (δθ  7.5 deg) relative to the to collapse reasonably well using the propeller tip Mach number, with
negative angles (δθ  15 deg). Microphone diaphragms were maximum values hovering between 0.65 and 0.70.
oriented so that they faced the multirotor drone at a distance of A comparison of the thrust values measured using the isolated
φ  43.5 in: from the drone center. Table 1 provides the placement propeller, quadcopter, and hexacopter configurations is shown in
of the arc array of microphones in dimensionless form. Given the Fig. 9a for the 8 and 10-in.-diam propellers. A second-order least-
bandwidths of interest and the close proximity of this microphone squares fit of the data (denoted by solid and dashed lines in Fig. 9) for
the individual cases is also displayed and reveals subtle differences
array to the multirotor drone, corrections for atmospheric absorption
between the three configurations for a given propeller diameter. The
were not implemented because they were found to be insignificant.
thrust values for the quadcopter and hexacopter have been divided by
The test matrix for this study is provided in Table 2 and was chosen
the number of propellers in their respective setups to match the
to cover the maximum safe operating range, in revolutions per second
isolated propeller data. Because the drone pitch is the same for both
(rev∕s), allowed by the motors for a given multirotor configuration
quadcopter and hexacopter configurations (λ  2.25), the spacing
between neighboring propeller tips is smaller for the hexacopter
configuration than for the quadcopter, which has been found to
elevate interactions between neighboring propeller disks and
adversely affect thrust. For example, Intaratep et al. [9] measured the
thrust and acoustics of a commercial quadcopter drone kit comprising
a drone pitch of approximately 1.45 and observed 5.8 and 7.3%
reductions in thrust when going from the isolated propeller to a
bicopter configuration and then from a bicopter to quadcopter
configuration, respectively. The current data exhibit this same
phenomenon where the hexacopter configuration yields slightly
smaller thrust values relative to the quadcopter. Previously tested
quadcopters at UTAustin (not shown) had the propellers thrusting up
and with the support arms in close proximity to the propeller disk
plane. Doing so placed the adjustable support arms directly in the
Fig. 6 Arrangement of instrumentation relative to the quadcopter propeller downwash. This impedance not only reduced thrust
configuration with coordinate system. (through some wake/pylon interaction effect) but also generated
2820 TINNEY AND SIROHI

Table 2 Test matrix for quadcopter and hexacopter configurations


d, in. Configuration 30 rev∕s 45 rev∕s 60 rev∕s 75 rev∕s 90 rev∕s 105 rev∕s 120 rev∕s 135 rev∕s 150 rev∕s 165 rev∕s
p p p p p p p p p p
8.0 Quad p p p p p p p p p p
9.0 Quad p p p p p p p p p
10.0 Quad p p p p p p p p ——
11.0 Quad p p p p p p p —— ——
12.0 Quad p p p p p p p —p— —p— —p—
8.0 Hex p p p p p p p p
10.0 Hex —— ——

vortex interaction noise; these kinds of effects have been studied tested. Offsets between different propeller diameters due to Reynolds
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

recently by Zawodny and Boyd [10]. Therefore, the final multirotor number effects are also uniform. In subsequent analysis, the
configuration entailed inverted propellers (thrusting down) that were polynomial coefficients corresponding to the individual configura-
extended away from the drone body using motor extensions tions are preserved, though the averages (corrected to the isolated
(identified in Fig. 2), which ultimately provided a better comparison propeller configuration values) are provided in Table 3 for the
to the isolated propeller case. The rotor offset distance, measured interested reader.
from the top of the adjustable arms to the rotor disk plane, was
measured to be 4.25 in. B. Acoustic Performance
In Fig. 9b, all thrust data are shown using log scales on both the Given the overwhelming amount of data that were acquired during
ordinate and abscissa axes. The agreement between the three this test campaign, only a fraction of it will be presented here. Our
configurations (isolated, quadcopter, hexacopter) is quite good for all emphasis is on developing a general understanding of the acoustic
propeller diameters and motor speeds. These trends exhibit a linear footprint in terms of its spectral behavior, its decay with distance from
second-order growth in thrust over the range of propeller speeds the source, and its directivity at angles surrounding the rotor disk

16 16

14 14

12 12

10 10

8 8

6 6

20 40 60 80 100 120 140 160 180 0.05 0.1 0.15 0.2 0.25 0.3 0.35
a) b)
Fig. 7 Thrust coefficient for an isolated propeller versus a) motor rotation speed, and b) propeller tip Mach number.

0.7
2

0.6

1.5 0.5

0.4

1
0.3

0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.05 0.1 0.15 0.2 0.25 0.3 0.35
a) b)
Fig. 8 Representations of a) torque coefficient, and b) figure of merit for an isolated propeller.

2.5
1
2

1.5

1 0.1
0.5

0
0.05 0.1 0.15 0.2 0.25 0.3 0.35 40 60 80 100 120 140 160 180
a) b)
Fig. 9 Representations of a) thrust of all configurations using 8 and 10-in.-diam propellers [comparison between raw data (symbols) and a second-order
least-squares fit (lines)], and b) thrust for all propeller diameters and configurations.
TINNEY AND SIROHI 2821

Table 3 Averaged coefficients (×103 ) from second-order fit δf  81.97 Hz for 30 and 165 rev∕s rotation speeds, respectively.
of thrust with motor speed (revolutions per second) Doing so reduces spurious noise and low-frequency waveform
d, in. Slope Quadratic coefficient modulations from contaminating the spectral estimate. As for the
multirotor drone, a constant partition size of N 213 data points is
8.0 −1.1593 9.5111
9.0 −1.7459 10.5742 used, thus yielding a narrow spectral resolution of δf  4.88 Hz;
10.0 −2.0468 8.1556 propeller blades in the multirotor configuration are independently
11.0 −2.6858 8.0788 operated with varying clock positions, and so the processing method is
12.0 −3.5147 1.4668 less elaborate, relative to the isolated propeller case. Converting Eq. (5)
to the decibel scale yields the sound pressure level (SPL):
a
Coefficients normalized by the number of propellers in each configuration.
 
PSDθj ; f
SPLθj ; f  10log10 (6)
pref 2
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

plane. For the isolated propeller configuration, the 1∕rev magnetic


sensor allowed acoustic pressure waveforms to be partitioned and and uses thep standard reference pressure for air of
then aligned on a per revolution basis. This is demonstrated in Fig. 10 pref  20 μPa∕ Hz. Corrections for human ear effects are achieved
where the average acoustic waveforms are displayed for two using the A-weighting standard described by the International
propeller diameters (9 and 12 in.) and different rotor speeds. Each Organization for Standardization (ISO 226:2003). Published values for
average waveform encompasses two rotor revolutions as recorded by the A-weighting function are first interpolated to match the resolution
the 0 deg observer. As expected, increasing rotation speed is of the narrowband spectra. The correction is then performed as follows:
accompanied by increasing pressure amplitude. At low rotation
speeds, the periods of the smaller-amplitude waveforms are difficult SPLA θj ; f  SPLθj ; f  Ad f
to identify because the total waveform shape is dominated by higher-  
PSDθj ; f ⋅ Ap f
frequency waveforms that are phase aligned with the propeller (the  10log10
averaging process removes only incoherent noise). The source of pref 2
 
these high-frequency waveforms will become apparent in the PSDA θj ; f
upcoming discussion.  10log10 (7)
pref 2
To characterize the spectral makeup of these acoustic waveforms,
power spectral densities (PSDs) are computed. For a sensor set where PSDA θj ; f  PSDθj ; f ⋅ Ap f. Premultiplied spectra are
comprising θj sensors, the two-sided autospectral density function is then computed for both the original, Gθj ;f  PSDθj ;f ⋅ f∕σ 2 θj ,
defined as
and A-weighted, GA θj ; f  PSDA θj ; f ⋅ f∕σ 2A θj , spectra
Z
1 ∞ following the method described by Baars et al. [11]. Here, σ 2 θj 
PSDθj ; f  < pθj ; tpθj ; t  ζ > e−i2πfζ dζ (5) and σ 2A θj  are the variances of the original and A-weighted signals,
2π −∞
respectively.
where ζ signifies a time delay, hi denotes ensemble averaging, and Figures 11 and 12 illustrate these different processing methods
limits of integration are confined to the size of the partition. For the using the 0 deg microphone observer and the 9 and 12-in.-diam
isolated propeller, the partition size varies and is based on the number of isolated propellers operating at 105 rev∕s. Figure 11a shows how the
samples required to cover two rotations of the propeller (N  2fs ∕Ω). motor and speed controllers alone (without propeller blades installed)
Therefore, spectral resolutions range from δf  14.88 Hz to produce a spectral peak around 1365 Hz corresponding to clogging of

0.1
0.1
0.05

0 0

-0.05
-0.1
-0.1
0 0.5 1.0 1.5 2.0 0 0.5 1.0 1.5 2.0
a) b)
Fig. 10 Average acoustic waveforms corresponding to two rotor revolutions of a) 9, and b) 12-in.-diam isolated propellers as measured by the θ4  0 deg
observer.

60 60

50 50

40 40
30 30
20 20
10 10
0 0
-10 -10
-20 -20
2 3 4
10 10 10 1 10 100
a) b)
Fig. 11 Representations of a) SPL of the isolated propeller at 105 rev∕s for the θ4  0 deg observer, and b) effect of partition size on SPLA using the
12-in.-diam isolated propeller at 105 rev∕s and the 0 deg observer.
2822 TINNEY AND SIROHI

15 2

1.5
10

5
0.5

0 0
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

0.1 1 10 100 1 10 100

a) b)
Fig. 12 Acoustic filtering effects applied to a) 9, and b) 12-in.-diam isolated propellers at 105 rev∕s and the θ4  0 deg observer.

the motor magnets (13 magnets per revolution spinning at 105 To now gauge the effect of rotor speed on the spectral footprint of
revolutions per second). The second equally significant peak at this multirotor configuration, we turn our attention to Fig. 13, where
3.9 kHz is believed to be pulse noise produced by the speed the 9 and 12-in.-diam propellers are installed on the quadcopter
controller, though this has not been verified. Nonetheless, the high- configuration. This employs the same observer location as before and
frequency waveforms displayed in Fig. 10 are shown here to be with A-weighting applied. Because these spectra are rich with
related to nonrotor harmonic noise. harmonic activity, the clutter that forms when overlaying the results
As for the propeller noise in Fig. 11, the fundamental and its from multiple test conditions conceals the details that are of interest to
harmonics appear crisp for both propeller sizes; the frequency scale in this study. Therefore, a detailed understanding of the effect of
Fig. 11b is defined by f  f∕Ωb, which uses the blade-pass propeller size and speed on the spectral footprint is not self-evident
frequency Ωb to more easily identify peaks corresponding to the without additional work; this kind of discussion is reserved for
fundamental frequency, its harmonics (appearing as integer multiples Sec. III.B.3, where the amplitudes of the first few rotor harmonics are
of f ), and even subharmonics. The noise floor from the facility and segregated and compared for a range of operating conditions. For
microphone data acquisition system is also included, though it has now, the discussion is confined to relatively general observations.
been made nondimensional along the frequency axis in Fig. 11b The most conspicuous of these observations is the increased sound
using 60 Hz (common line noise). Comparing the unweighted levels that accompany increased rotor speeds. Additional propellers
(Fig. 11a) and A-weighted signals (Fig. 11b) reveals the significance also manifest increased sound amplitudes relative to the isolated
of the higher harmonics with respect to the human ear. Furthermore, rotor. Once again, noise from the motor and speed controller persists
the effect of partition size in Fig. 11b is seen to have little effect on the and is a significant source of sound at low rotation speeds when the
shape of the spectra, with the first few harmonics being displayed as noise from the propeller (in the form of thickness noise) is quite small.
the dominant components of the signal. Overall, the spectral behavior It should be noted that any additional sources of noise caused by
is similar to the measurements reported by Sinibaldi and Marino [12], scattering of sound waves by electrical cables, extendable arms, or
Intaratep et al. [9], and Zawodny et al. [5], who studied single rotors the rotor test stand were not isolated in any of the configurations
operating at static thrust. tested. Thus, although we believe these additional sources of noise to
A dimensionless display of these spectra using the premultiplied be small, they are currently unknown, where comparison to the
spectra GA is provided in Fig. 12 for the same operating conditions. isolated propeller studies are concerned.
Relative to the SPL in Fig. 11, the higher harmonics are much more
significant sources of audible noise. A-weighted premultiplied 1. Time-Frequency Analysis
spectra are nearly void of energy in the fundamental with peak Because each propeller is driven by a motor that operates
energies residing in the 7th and 11th harmonics for the 9 and 12 in. independently from all other motors and speed controllers, any
propellers, respectively. Cross referencing the SPL in Fig. 11a with differences in the individual rotation speeds will generate beating
the peaks in Fig. 12a reveals how the two peaks at f  7 and 19 are phenomena due to quadratic interactions between their fundamental
caused by motor and speed controller noise, respectively, as opposed blade-pass frequencies. Thus, quadratic interactions in the near field
to propeller noise. From a human detection point of view, it is the between frequencies f1 and f2 , and with associated phases ϕ1 and ϕ2 ,
higher harmonic signatures, including the noise from the motor and result in a spectral peak at frequency f and phase ϕ in the far field
speed controller, that are the prominent sources of noise. according to the selection rules f  f1  f2 and ϕ  ϕ1  ϕ2 .

60 60

40 40

20 20

0 0

1 10 100 1 10 100
a) b)
Fig. 13 SPLθ4 ; f for a range of propeller speeds of the quadcopter with a) 9, and b) 12-in.-diam propellers.
TINNEY AND SIROHI 2823

Quadratic interactions can be quantified using bispectral methods but signal, and so the Morlet wavelet is the suitable choice. This Morlet
are cumbersome to apply; see Baars and Tinney [13] and the wavelet is defined as
references therein. Forward flight and drone maneuvering exacerbate
ψt∕l  ejωψ t∕l e−jt∕lj
2 ∕2
the phenomena, which is of fundamental interest to the establishment (9)
of a broader class of acoustic models for characterizing drone noise.
For now, we reserve the application of higher-order spectral analysis with a central frequency of ωψ  6. The analysis shown here mirrors
of the sound produced by multirotor drones for future efforts and previous efforts described by Baars and Tinney [16], Stephenson
consider an alternate route for determining whether quadratic et al. [17], and Rojo et al. [18] and is performed in the Fourier domain
coupling may be present in this data set. using 81 unique scales distributed logarithmically over the frequency
An alternative way of identifying nonlinear coupling in these range 100 Hz < f < fs ∕2. Only regions inside the cone of influence
acoustic signatures is to use time-frequency analysis to visualize the are shown and are constructed by overlapping signal partitions
temporal behavior of the signal’s spectral makeup. Any beating comprising N 217 samples. Like the Fourier transform, the energy
phenomenon would be revealed by variations in the amplitude of the density can be computed and is known as the wavelet power spectrum
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

fundamental blade-pass frequencies over time. The process is (WPS):


thoroughly described elsewhere [14,15] and involves the convolution
of a mother wavelet ψt∕l with a time-dependent signal to produce jpθ
~ j ; l; tj2
wavelet coefficients as follows: Eθj ; l; t  (10)
l
Z  0 
t −t which is then converted to Fourier frequency to obtain the double-
~ j ; l; t  pθj ; t 0 ψ 
pθ dt (8)
l sided WPS, denoted as Ex; f; t. To simplify the analysis, we will
focus on the microphone signals located at θ1  −45 deg and θ4 
The signal in this case is the unsteady acoustic pressure pθj ; t 0 deg relative to the rotor disk plane, which is in the general direction
with l being the time scale of the predefined wavelet. The Morlet of where a far-field observer would reside.
wavelet will be used here, given that it offers crisper resolutions in The findings from this analysis are shown in Figs. 14 and 15 for the
frequency at the expense of coarser resolutions in time, relative to the quadcopter and hexacopter configurations, respectively, and for a
Mexican hat wavelet; we are interested in the frequency content of the duration of 8 s. For the sake of simplicity, the results are confined to a

10
60

40

20
1

1 2 3 4 5 6 7 8
a)
10
60

40
20
1

1 2 3 4 5 6 7 8
b)
Fig. 14 Morlet wavelet power spectra expressed as 10log10 Eθj ; f;t∕pref2  for a) θ1  −45 deg, and b) θ4  0 deg observer with the quadcopter
configuration using 10-in.-diam propellers at 135 rev∕s.

10
60

40

20
1

1 2 3 4 5 6 7 8
a)

10
60

40

20
1

1 2 3 4 5 6 7 8
b)
Fig. 15 Morlet wavelet power spectra expressed as 10log10 Eθj ; f;t∕pref2  for a) θ1  −45 deg, and b) θ4  0 deg observer with the hexacopter
configuration using 10-in.-diam propellers at 135 rev∕s.
2824 TINNEY AND SIROHI

motor rotation speed of 135 rev∕s (Ωb  270 passes per second) spherical decay laws are commonly used to propagate acoustic
using 10-in.-diam propeller blades. The y axis, again, has been pressure signals from the near field to the far field. To the authors’
normalized by the blade-pass frequency and is confined to illustrate knowledge, the 1∕r decay law has not yet been verified for multirotor
only the first 10 harmonics. At first glance, three striking features are drones. Figure 16 supports the conjecture that the acoustic pressure
observed. The first is the amplitude modulations at the fundamental signal spreads spherically beyond 2φ∕D  d ≥ 2.5 and 2.0 for the
blade-pass frequency (f  1), which suggests quadratic interactions 9 and 12-in.-diam rotor configurations, respectively. This verifies that
between acoustic waves produced by neighboring propeller tips. The the arc array is placed within the acoustic near-field regions of the
second is the differences between the higher harmonics at the different multirotor drone for all configurations and blade sizes tested.
observer stations. For a given multirotor configuration, the signals However, these measurements only verify this decay along the rotor
from the two microphones were acquired simultaneously. The disk plane, which may be different along other propagation paths
observer located below the rotor disk plane experiences a broader given the directivity pattern of the dipole sources.
spectrum of higher harmonic activity, relative to the observer aligned
with the disk plane, and is attributed to the directivity of dipole noise 3. Filtered Overall Sound Pressure Level and Directivity
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

sources; we will see evidence of this later on. A third observation is in Because the dominant features in these acoustic waveforms are
the hexacopter configuration where the amplitude modulation of the
integer multiples of the fundamental blade-pass frequency, an effort
second harmonic (2Ωb ) is much stronger than in the quadcopter
is undertaken to track the effect of propeller diameter, rotor speed,
configuration. This may be due to the closer proximity of the propeller
and configuration (quadcopter versus hexacopter) on the first few
blade tips in the hexacopter configuration that is amplifying
harmonics associated with the blade-pass frequency. This is
interactions between neighboring blade tips.
performed following Parseval’s theorem as a guide so that
2. Acoustic Pressure Decay Z f 2δf
i
It is an essential part of any acoustic study to know what the decay σ 2i θj   2 PSDθj ; f df (11)
fi −2δf
rate and decay path are for the prominent sound source. This decay
was measured for two quadcopter configurations comprising the 9
where the subscript i is an integer identifier corresponding to the
and 12-in.-diam propellers. Like the arc array, the line array was
fundamental blade-pass frequency. A factor of 2 is inserted to account
aligned with the rotor disk plane at θ  0 deg and with measurement
for the energy neglected in the negative frequencies. The integration
points spanning the range φ  18; : : : ; 72 in: for the 9 in. propellers
width is 5δ, which corresponds to 24.41 Hz in the multirotor drone
and from φ  24; : : : ; 72 in: for the 12 in. propellers using
increments of δφ  6 in: for both. The findings from this study are setup. This was needed to capture the energy in the tails of the spectral
shown in Fig. 16 using a high-pass filter to remove waveforms below peaks. Thus, σ 2i quantifies the variance of the signal corresponding to
70 Hz. Given that thickness noise is the prominent source of noise and the blade-pass frequency and its harmonics. OASPLi is then
is known to decay spherically from the source, the measured pressure calculated in standard fashion as follows:
amplitude from these multirotor drones should decay like 1∕r. This is  
important for several reasons. Foremost, it is known that disturbances σ 2i θj 
OASPLi θj   10log10 (12)
close to the propeller blade are dominated by evanescent pressure pref 2
waves (pseudosound waves) that decay within the first few
wavelengths from the source and do not propagate to the far field. so that OASPL1 θj  is the energy associated with the fundamental
Therefore, the radial position where these hydrodynamic pressure blade-pass frequency of the jth sensor, OASPL2 θj  is associated
waves are overtaken by acoustic pressure waves is needed to verify with the second harmonic, and so on.
that the microphone arc array is indeed located in a region where The findings from this are displayed in Figs. 17–19 as a function of
acoustic pressure waves dominate the microphone signal. Second, the measured thrust and for a given drone configuration. Clear trends

100
100

90
90

80 80

70
70

60
60
1 1.5 2 2.5 3 3.5 4 4.5 1 1.5 2 2.5 3 3.5 4

a) b)
Fig. 16 Pressure decay along the line array compared with the spherical decay law (1∕r) for a quadcopter with a) 9, and b) 12-in.-diam propeller.

100 100
90 90
80 80
70 70
60 60
50 50
40 40
30 30
1 10 1 10
a) b)
Fig. 17 Fundamental blade-pass frequency noise at a) θ1  −45 deg, and b) θ4  0 deg observer positions.
TINNEY AND SIROHI 2825

100 100
90 90
80 80
70 70
60 60
50 50
40 40
30 30
1 10 1 10
a) b)
Fig. 18 Second harmonic noise at a) θ1  −45 deg, and b) θj  0 deg observer positions.
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

100 100
90 90
80 80
70 70
60 60
50 50
40 40
30 30
1 10 1 10
a) b)
Fig. 19 Third harmonic noise at a) θ1  −45 deg, and b) θ4  0 deg observer positions.

30 30 30

20 20 20

10 10 10

0 0 0

-10 -10 -10

-20 -20 -20

-30 -30 -30

-40 -40 -40

-50 -50 -50


20 40 60 80 20 40 60 80 40 50 60 70 80
a) b) c)
Fig. 20 OASPL directivity patterns for a) fundamental blade-pass frequency, b) second harmonic, and c) third harmonic for the quadcopter
configurations at 105 rev∕s.

are revealed that demonstrate the rate by which the sound pressure thickness noise is the predominant factor here, which is governed
amplitude of the fundamental frequency and its higher harmonics principally by blade thickness and tip Mach number. On the contrary,
increase with increasing motor speed (thrust). In general, the change the second and third harmonics are less symmetric and with
in OASPL due to changes in thrust is similar for both the hexacopter significant drops in amplitude occurring at, or just above, the rotor
and quadcopter configurations and for different size propellers. disk plane; this was also observed in Figs. 17–19. These higher
Closer inspection reveals how the sound pressure, for a given thrust harmonics are thus attributed to loading noise, which is consistent
value, decreases with increasing propeller diameter. The only with large-scale rotorcraft studies. Furthermore, the shape of the
distinguishing factor is the blade-pass frequency. Energy in the directivity pattern appears to be unaffected by Reynolds number
fundamental blade-pass frequency is also very similar between effects (due to changes in the propeller diameter). The only
the shallow (−45 deg) and rotor disk (0 deg) observer stations. On discrepancies in the shape of the directivity pattern appear in the
the contrary, the second and third harmonics are similar in amplitude higher harmonics and with the smaller-diameter propeller blades.
to the fundamental frequency at shallow angles but nearly 10 dB
weaker along the rotor disk plane, which complements the findings
from the wavelet analysis in Figs. 14 and 15. IV. Conclusions
Where the acoustic behaviors from all available measurement This paper presents a study to understand the sound produced by
stations are concerned, the same filtered data using Eq. (12) are multirotor drones operating at static thrust. Load measurements show
shown in Fig. 20 for the quadcopter configuration only and for a rotor that, for a drone pitch of 2.25, defined as a ratio of the rotor diameter
speed of 105 rev∕s. Several points of interest are seen here. to hub diameter, thrust levels resort to integer multiples of the number
Foremost, the directivity patterns for the fundamental blade-pass of propeller blades. The sound field is shown to encompass both
frequencies are found to be symmetric about a line located just below nonrotor harmonic noise (motor noise and noise from the speed
the rotor disk plane for all propeller diameters. This suggests that controller) and main rotor harmonic noise (thickness noise and
2826 TINNEY AND SIROHI

loading noise). The former of these is as equally important as the [5] Zawodny, N. S., Boyd, D. D., and Burley, C. L., “Acoustic
latter and is even more significant where human ear effects, on Characterization and Prediction of Representative, Small-Scale Rotary-
account of A-weighted filtering, are concerned. Time-frequency Wing Unmanned Aircraft System Components,” Proceedings of the
analysis of the sound field demonstrates modulations of the pressure American Helicopter Society 72nd Annual Forum, West Palm Beach,
FL, May 2016.
levels associated with the first few blade-pass frequencies that occur [6] Cameron, C., Karpatne, A., and Sirohi, J., “Performance of a Mach-
intermittently along the tip path plane caused by subtle variations in Scale Coaxial Counter-Rotating Rotor in Hover,” Journal of Aircraft,
the motor speeds among the multiple motors. Pressure levels Vol. 53, No. 3, 2016, pp. 746–755.
associated with the higher harmonics are shown to be more doi:10.2514/1.C033442
significant at observer angles below the tip path plane but are still [7] Mula, S. M., Stephenson, J. H., Tinney, C. E., and Sirohi, J., “Dynamical
significant at all observer angles studied. and Evolutionary Characteristics of the Tip Vortex from a Four-Bladed
For a given thrust, the sound pressure of the first few harmonics is Rotor in Hover,” Proceedings of the American Helicopter Society 68th
shown to decrease with both increasing propeller diameter and the Annual Forum, Ft. Worth, TX, May 2012.
number of propellers. These changes are attributed to the lower [8] Brandt, J. B., and Selig, M. S., “Propeller Performance Data at Low
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

Reynolds Numbers,” 49th AIAA Aerospace Sciences Meeting, AIAA


rotation speeds, and hence lower thickness and loading noise, Paper 2011-1255, Jan. 2011.
required to maintain the same thrust levels. The spatial decay of the [9] Intaratep, N., Alexander, W. N., and Devenport, W. J., “Experimental
overall sound pressure level along the tip path plane also revealed the Study of Quadcopter Acoustics and Performance at Static Thrust
demarcation between evanescent and acoustic components of the Conditions,” 22nd AIAA Aeroacoustics Conference, AIAA Paper 2016-
pressure field; the latter is characterized by spherical spreading and 2873, 2016.
was found to occur at positions closer to the drone hub for smaller [10] Zawodny, N. S., and Boyd, D. D., “Investigation of Rotor-Airframe
propellers. Directivity patterns associated with the blade-pass Interaction Noise Associated with Small-Scale Rotary-Wing Unmanned
frequency for different rotor diameters show it to be the result of Aircraft Systems,” Proceedings of the American Helicopter Society
thickness noise produced by the rotor. The same study of the second 73rd Annual Forum, Ft. Worth, TX, May 2017.
[11] Baars, W. J., Tinney, C. E., and Hamilton, M. F., “Piecewise-Spreading
and third harmonics shows them to be loading-noise effects, similar Regime Model for Calculating Effective Gol’dberg Numbers for
to what is observed with full-scale helicopters. Supersonic Jet Noise,” AIAA Journal, Vol. 54, No. 9, Sept. 2016,
Although a basic understanding of the sound produced by pp. 2833–2842.
multirotor drones during static thrust conditions has been developed, doi:10.2514/1.J054790
there is still much to be learned about the effects of forward flight and [12] Sinibaldi, G., and Marino, L., “Experimental Analysis on the Noise of
maneuvers of the vehicle. Such a topic should be the focus of future Propellers for Small UAV,” Journal of Applied Acoustics, Vol. 74, No. 1,
efforts because they will enhance impulsivelike noise effects not seen 2013, pp. 79–88.
during static operations of a drone. doi:10.1016/j.apacoust.2012.06.011
[13] Baars, W. J., and Tinney, C. E., “Proper Orthogonal Decomposition-
Based Spectral Higher-Order Stochastic Estimation,” Physics of Fluids,
Acknowledgment Vol. 26, No. 5, 2014, Paper 055112.
doi:10.1063/1.4879255
The authors wish to thank Bryson Hill for assisting in the design [14] Farge, M., “Wavelet Transforms and Their Application to Turbulence,”
and fabrication of the multirotor drone test stand as well as Chris Annual Review of Fluid Mechanics, Vol. 24, Jan. 1992, pp. 395–458.
Cameron and John Valdez for helping with the thrust and acoustic doi:10.1146/annurev.fl.24.010192.002143
measurements. [15] Addison, P. S., The Illustrated Wavelet Transform Handbook, Taylor
and Francis, New York, 2002, pp. 1–353.
[16] Baars, W. J., and Tinney, C. E., “Transient Wall Pressure in an
References Overexpanded and Large Area-Ratio Nozzle,” Experiments in Fluids,
[1] Deters, R. W., and Selig, M. S., “Static Testing of Micro Propellers,” Vol. 54, No. 1468, Feb. 2013, pp. 1–17.
26th AIAA Applied Aerodynamics Conference, AIAA Paper 2008-6246, doi:10.1007/s00348-013-1468-8
Aug. 2008. [17] Stephenson, J. H., Tinney, C. E., Green, E., and Watts, M. E., “Time
[2] Ol, M., Zeune, C., and Logan, M., “Analytical–Experimental Comparison Frequency Analysis of Sound from a Maneuvering Rotorcraft,”
for Small Electric Unmanned Air Vehicles,” 26th AIAA Applied Journal of Sound and Vibration, Vol. 333, No. 21, 2014, pp. 5324–
Aerodynamics Conference, AIAA Paper 2008-7345, Aug. 2008. 5339.
[3] Mula, S. M., Stephenson, J. H., Tinney, C. E., and Sirohi, J., “Dynamical doi:10.1016/j.jsv.2014.05.018
Characteristics of the Tip Vortex from a Four-Bladed Rotor in Hover,” [18] Rojo, R., Tinney, C. E., and Ruf, J. H., “Effect of Stagger on the
Experiments in Fluids, Vol. 54, No. 10, Oct. 2013, Paper 1600. Vibroacoustic Loads from Clustered Rockets,” AIAA Journal, Vol. 54,
doi:10.1007/s00348-013-1600-9 No. 11, 2016, pp. 3588–3597.
[4] Fiévet, R., Tinney, C. E., Baars, W. J., and Hamilton, M. F., doi:10.2514/1.J055017
“Coalescence in the Sound Field of a Laboratory-Scale Supersonic Jet,”
AIAA Journal, Vol. 54, No. 1, 2016, pp. 254–265. D. Papamoschou
doi:10.2514/1.J054252 Associate Editor
This article has been cited by:

1. Yan Wu, Michael J. Kingan, Ryan S. McKay, Sung Tyaek Go, Young-min Shim. 2022. Extrapolation and inversion of
near-field propeller noise measurements. Applied Acoustics 185, 108395. [Crossref]
2. Reynard de Vries, Nando van Arnhem, Tomas Sinnige, Roelof Vos, Leo L.M. Veldhuis. 2021. Aerodynamic interaction
between propellers of a distributed-propulsion system in forward flight. Aerospace Science and Technology 118, 107009.
[Crossref]
3. Emanuele L. de Angelis, Fabrizio Giulietti, Gianluca Rossetti, Gabriele Bellani. 2021. Performance analysis and optimal
sizing of electric multirotors. Aerospace Science and Technology 118, 107057. [Crossref]
4. Dongwook Kim, Jeongwoo Ko, Vignesh Saravanan, Soogab Lee. 2021. Stochastic analysis of a single-rotor to quantify
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

the effect of RPS variation on noise of hovering multirotors. Applied Acoustics 182, 108224. [Crossref]
5. R. Niemiec, F. Gandhi, N. Kopyt. 2021. Relative rotor phasing for multicopter vibratory load minimisation. The
Aeronautical Journal 1-20. [Crossref]
6. Huanxian Bu, Han Wu, Celia Bertin, Yi Fang, Siyang Zhong. 2021. Aerodynamic and acoustic measurements of dual
small-scale propellers. Journal of Sound and Vibration 511, 116330. [Crossref]
7. C. T. Justine Hui, Michael J. Kingan, Yusuke Hioka, Gian Schmid, George Dodd, Kim N. Dirks, Shaun Edlin, Sean
Mascarenhas, Young-Min Shim. 2021. Quantification of the Psychoacoustic Effect of Noise from Small Unmanned Aerial
Vehicles. International Journal of Environmental Research and Public Health 18:17, 8893. [Crossref]
8. Jesse Callanan, Rayhaan Iqbal, Revant Adlakha, Amir Behjat, Souma Chowdhury, Mostafa Nouh. 2021. Large-aperture
experimental characterization of the acoustic field generated by a hovering unmanned aerial vehicle. The Journal of the
Acoustical Society of America 150:3, 2046-2057. [Crossref]
9. Caterina Poggi, Giovanni Bernardini, Massimo Gennaretti. Aeroacoustic analysis of wing-mounted propeller arrays .
[Abstract] [PDF] [PDF Plus]
10. Alper Celik, Nur Syafiqah Jamaluddin, Kabilan Baskaran, Djamel Rezgui, Mahdi Azarpeyvand. Aeroacoustic Performance
of Rotors in Tandem Configuration . [Abstract] [PDF] [PDF Plus]
11. Caterina Poggi, Monica Rossetti, Giovanni Bernardini, Umberto Iemma, Cristiano Andolfi, Christian Milano, Massimo
Gennaretti. 2021. Surrogate models for predicting noise emission and aerodynamic performance of propellers. Aerospace
Science and Technology 107016. [Crossref]
12. Paweł Zimroz, Paweł Trybała, Adam Wróblewski, Mateusz Góralczyk, Jarosław Szrek, Agnieszka Wójcik, Radosław Zimroz.
2021. Application of UAV in Search and Rescue Actions in Underground Mine—A Specific Sound Detection in Noisy
Acoustic Signal. Energies 14:13, 3725. [Crossref]
13. Ranieri Emanuele Nargi, Fabrizio De Gregorio, Paolo Candeloro, Giuseppe Ceglia, Tiziano Pagliaroli. 2021. Evolution of
flow structures in twin-rotors wakes in drones by time-resolved PIV. Journal of Physics: Conference Series 1977:1, 012008.
[Crossref]
14. Ryan S. McKay, Michael J. Kingan, Sung Tyaek Go, Riul Jung. 2021. Experimental and analytical investigation of contra-
rotating multi-rotor UAV propeller noise. Applied Acoustics 177, 107850. [Crossref]
15. Romain Gojon, Thierry Jardin, Hélène Parisot-Dupuis. 2021. Experimental investigation of low Reynolds number rotor
noise. The Journal of the Acoustical Society of America 149:6, 3813-3829. [Crossref]
16. Beat Schäffer, Reto Pieren, Kurt Heutschi, Jean Marc Wunderli, Stefan Becker. 2021. Drone Noise Emission
Characteristics and Noise Effects on Humans—A Systematic Review. International Journal of Environmental Research and
Public Health 18:11, 5940. [Crossref]
17. Haitao Hu, Yannian Yang, Yu Liu, Xiaomin Liu, Yong Wang. 2021. Aerodynamic and aeroacoustic investigations of multi-
copter rotors with leading edge serrations during forward flight. Aerospace Science and Technology 112, 106669. [Crossref]
18. Harini Kolamunna, Thilini Dahanayaka, Junye Li, Suranga Seneviratne, Kanchana Thilakaratne, Albert Y. Zomaya, Aruna
Seneviratne. 2021. DronePrint. Proceedings of the ACM on Interactive, Mobile, Wearable and Ubiquitous Technologies 5:1,
1-31. [Crossref]
19. Antonio J. Torija, Paruchuri Chaitanya, Zhengguang Li. 2021. Psychoacoustic analysis of contra-rotating propeller noise
for unmanned aerial vehicles. The Journal of the Acoustical Society of America 149:2, 835-846. [Crossref]
20. Woutijn J. Baars, Liam Bullard, Abdulghani Mohamed. Quantifying modulation in the acoustic field of a small-scale rotor
using bispectral analysis . [Abstract] [PDF] [PDF Plus]
21. A. Filippone, G.N. Barakos. 2021. Rotorcraft systems for urban air mobility: A reality check. The Aeronautical Journal
125:1283, 3-21. [Crossref]
22. Con Doolan, Yendrew Yauwenas, Danielle Moreau. Drone Propeller Noise Under Static and Steady Inflow Conditions
45-60. [Crossref]
23. Yannian Yang, Yong Wang, Yu Liu, Haitao Hu, Zhiyong Li. 2020. Noise reduction and aerodynamics of isolated multi-
copter rotors with serrated trailing edges during forward flight. Journal of Sound and Vibration 489, 115688. [Crossref]
24. Daniel Weitsman, James H. Stephenson, Nikolas S. Zawodny. 2020. Effects of flow recirculation on acoustic and dynamic
measurements of rotary-wing systems operating in closed anechoic chambers. The Journal of the Acoustical Society of America
148:3, 1325-1336. [Crossref]
Downloaded by DLR DEUTSCHES ZENTRUM FUER LUFT UND RAUMFAHRT on November 22, 2021 | http://arc.aiaa.org | DOI: 10.2514/1.J056827

25. Michel Roger, Stéphane Moreau. 2020. Tonal-Noise Assessment of Quadrotor-Type UAV Using Source-Mode
Expansions. Acoustics 2:3, 674-690. [Crossref]
26. Kenneth W. Van Treuren, Ricardo Sanchez, Charles Wisniewski, Paul Leitch. Investigation of Aeroacoustics and Motor
Efficiency of a Two-Bladed Stock and Five-Bladed Propeller Designs for Static Quadcopter Applications . [Abstract]
[PDF] [PDF Plus]
27. Yannian Yang, Yu Liu, Haitao Hu, Xiaomin Liu, Yong Wang, Elias J.G. Arcondoulis, Zhiyong Li. 2020. Experimental
study on noise reduction of a wavy multi-copter rotor. Applied Acoustics 165, 107311. [Crossref]
28. Quinten Henricks, Zhenyu Wang, Mei Zhuang. 2020. Small-Scale Rotor Design Variables and Their Effects on
Aerodynamic and Aeroacoustic Performance of a Hovering Rotor. Journal of Fluids Engineering 142:8. . [Crossref]
29. William A. Jordan, Shreyas Narsipur, Robert Deters. Aerodynamic and Aeroacoustic Performance of Small UAV Propellers
in Static Conditions . [Abstract] [PDF] [PDF Plus]
30. Sangwon Do, Myeongjae Lee, Jong-Seon Kim. 2020. The Effect of a Flow Field on Chemical Detection Performance of
Quadrotor Drone. Sensors 20:11, 3262. [Crossref]
31. Madhavan Shanmugavel, V. Sujatha, H. Kristanto, B. Amin, L. H. Teo, Veera Ragavan. Quadrotor mounted Inverted
Spherical Pendulum Test Stand - Design and Characterization 922-927. [Crossref]
32. 2020. Rotor interactional effects on aerodynamic and noise characteristics of a small multirotor unmanned aerial vehicle.
Physics of Fluids 32:4, 047107. [Crossref]
33. Giovanni Bernardini, Francesco Centracchio, Massimo Gennaretti, Umberto Iemma, Claudio Pasquali, Caterina Poggi,
Monica Rossetti, Jacopo Serafini. 2020. Numerical Characterisation of the Aeroacoustic Signature of Propeller Arrays for
Distributed Electric Propulsion. Applied Sciences 10:8, 2643. [Crossref]
34. Charles E. Tinney, John Valdez. 2020. Thrust and Acoustic Performance of Small-Scale, Coaxial, Corotating Rotors in
Hover. AIAA Journal 58:4, 1657-1667. [Abstract] [Full Text] [PDF] [PDF Plus]
35. Dongyeon Han, Doo Young Gwak, Soogab Lee. 2020. Noise prediction of multi-rotor UAV by RPM fluctuation correction
method. Journal of Mechanical Science and Technology 34:4, 1429-1443. [Crossref]
36. Y. Yang, Y. Liu, Y. Li, E. Arcondoulis, Y. Wang. 2020. Aerodynamic and Aeroacoustic Performance of an Isolated
Multicopter Rotor During Forward Flight. AIAA Journal 58:3, 1171-1181. [Abstract] [Full Text] [PDF] [PDF Plus]
37. Jung-Hee Seo, Tyson L Hedrick, Rajat Mittal. 2020. Mechanism and scaling of wing tone generation in mosquitoes.
Bioinspiration & Biomimetics 15:1, 016008. [Crossref]
38. Chun Hyuk Park, Dae Han Kim, Young J. Moon. 2019. Computational Study on the Steady Loading Noise of Drone
Propellers: Noise Source Modeling with the Lattice Boltzmann Method. International Journal of Aeronautical and Space
Sciences 20:4, 858-869. [Crossref]
39. William N. Alexander, Jeremiah Whelchel, Nanyaporn Intaratep, Antonio Trani. Predicting Community Noise of sUAS .
[Citation] [PDF] [PDF Plus]
40. Austin Thai, Sheryl M. Grace. Prediction of small quadrotor blade induced noise . [Citation] [PDF] [PDF Plus]
41. Ryan McKay, Michael J. Kingan. Multirotor Unmanned Aerial System Propeller Noise Caused by Unsteady Blade Motion .
[Citation] [PDF] [PDF Plus]
42. Yendrew Yauwenas, Jeoffrey R. Fischer, Danielle Moreau, Con J. Doolan. The Effect of Inflow Disturbance on Drone
Propeller Noise . [Citation] [PDF] [PDF Plus]
43. James H. Stephenson, Daniel Weitsman, Nikolas S. Zawodny. 2019. Effects of flow recirculation on unmanned aircraft
system (UAS) acoustic measurements in closed anechoic chambers. The Journal of the Acoustical Society of America 145:3,
1153-1155. [Crossref]

You might also like