Candeloro Et Al (2022) SmallScale Rotor Aeroacoustics For Drone Propulsion - A Review of Noise Sources and Control Strategies
Candeloro Et Al (2022) SmallScale Rotor Aeroacoustics For Drone Propulsion - A Review of Noise Sources and Control Strategies
Candeloro Et Al (2022) SmallScale Rotor Aeroacoustics For Drone Propulsion - A Review of Noise Sources and Control Strategies
Review
Small-Scale Rotor Aeroacoustics for Drone Propulsion:
A Review of Noise Sources and Control Strategies
Paolo Candeloro 1, * , Daniele Ragni 2 and Tiziano Pagliaroli 1
1 Engineering Department, Universitá Degli Studi Niccoló Cusano, Via Don Carlo Gnocchi 3, 00166 Rome, Italy
2 Flow Physics and Technology Department, Faculty of Aerospace Engineering, Delft University of Technology,
Kluyverweg 1, 2629 HS Delft, The Netherlands
* Correspondence: [email protected]
Abstract: In the last decade, the drone market has grown rapidly for both civil and military purposes.
Due to their versatility, the demand for drones is constantly increasing, with several industrial
players joining the venture to transfer urban mobility to the air. This has exacerbated the problem
of noise pollution, mainly due to the relatively lower altitude of these vehicles and the proximity
of their routes to extremely densely populated areas. In particular, both the aerodynamic and
aeroacoustic optimization of the propulsive system and of its interaction with the airframe are key
aspects of unmanned aerial vehicle design that can signify the success or the failure of their mission.
The industrial challenge involves finding the best performance in terms of loading, efficiency and
weight, and, at the same time, the most silent configuration. For these reasons, research has focused on
an initial localization of the noise sources and, on further analysis, of the noise generation mechanism,
focusing particularly on directivity and scattering. The aim of the present study is to review the
noise source mechanisms and the state-of-the-art control strategies, available in the literature, for
its suppression, focusing especially on the fluid-dynamic aspects of low Reynolds numbers of the
propulsive system and on the interaction of the propulsive system flow with the airframe.
Citation: Candeloro, P.; Ragni, D.;
Pagliaroli, T. Small-Scale Rotor
Keywords: drones; aerodynamics; aeroacoustics; rotor noise; airframe noise; porous material
Aeroacoustics for Drone Propulsion:
A Review of Noise Sources and
Control Strategies. Fluids 2022, 7, 279.
https://doi.org/10.3390/
fluids7080279
1. Introduction
The term “drone” refers to an automatized vehicle with high manoeuvrability in
Academic Editors: Mehrdad
both hovering and cruise operations. In the most interesting configurations, unmanned
Massoudi and Goodarz Ahmadi
aerial vehicles (UAVs), often labelled as unmanned aerial systems (UAS) or micro-aerial
Received: 30 June 2022 vehicles (MAV) are already designed with vertical or horizontal take-off and landing
Accepted: 12 August 2022 (VTOL) capabilities and can manoeuvre with extremely high versatility and speed. Due
Published: 15 August 2022 to their unique properties, MAVs are often used in tactical surveillance missions or for
Publisher’s Note: MDPI stays neutral reconnaissance purposes. In order to gain information about the scouting area without
with regard to jurisdictional claims in being easily identified, achieving an acoustic stealth mode is an essential feature for mission
published maps and institutional affil- success. Despite the different aims, the noise footprint of these vehicles is extremely
iations. important even when employed in civilian roles due to their flight proximity to populated
urban areas. Some of their mission tasks still require geographical mapping, infrastructure
inspections, precision agriculture, delivery and e-commerce. Small drones will have an
enormous social and economic impact. In fact, this technology opens new possibilities in
Copyright: © 2022 by the authors. several application fields. For example, drones equipped with cameras can resolve the
Licensee MDPI, Basel, Switzerland. problem of the images taken by satellites (which are often expensive, weather-dependent
This article is an open access article and in low resolution) or car-based images (which are limited to human-level perspectives
distributed under the terms and
and the availability of accessible roads). In addition, farmers can check the quality of
conditions of the Creative Commons
crop growth by using cameras mounted on specific UAVs. These particular drones will
Attribution (CC BY) license (https://
also enable construction companies to verify work advancement in real time. For mining
creativecommons.org/licenses/by/
companies, interest focuses on the possibility of obtaining precise volumetric data, leading
4.0/).
to lower risks for their employers. Humanitarian organizations will be able to evaluate
and adapt aid efforts for refugee camps, while medical supplies can be delivered quickly
by rescue organizations where necessary. By using MAVs for transportation, developing
countries (i.e., countries without appropriate road networks) could deliver goods simply.
Inspection drones, vehicles able to fly in confined space, can be used by fire-fighting and
emergency units to assess danger faster and safely, or by logistic companies to detect
damage to both inner and outer shells of ships, or by road maintenance companies to
measure deterioration in bridges or tunnels. Security agencies will be able to improve
building safety by monitoring even the areas outside cameras range. Drones will enable
disaster mitigation agencies to inspect partially collapsed buildings in the event of obstacles
for terrestrial robots. Teams of autonomous drones coordinators will enable missions to last
longer than the flight time of a single drone by allowing it to leave the swarm for a short
time to replace its battery [1]. The interest in the use of drones in delivery is also related
to the fact that these vehicles enable reduced greenhouse gases and other environmental
impacts compared to conventional diesel delivery trucks [2–5].
The combination of distributed or multi-rotor propulsive systems, generally preferred
for manoeuvrability, and proximity to civil areas makes drone noise a challenging issue
for the European scientific community at both industrial and academic levels. In a 2018
document, the European Aviation Safety Agency (EASA) specified the noise level require-
ment for drones at a fixed value of 60 dB(A), measured at a distance of 3 m from the
source [6]. Generally, the strategic objectives for drone market growth are greater en-
durance and acoustic impact reduction. These two aspects are also key issues to improve
the safety of this technology in the future. Drone noise pollution is also a problem from
the point of view of public acceptance of the widespread deployment of flying drones in
urban areas. Just to give an idea of the problem related to a large-scale use of drones in
residential areas, information about the effects on the population of a large-scale test drone
for delivering can be found in an article from the Wall Street Journal (“Delivery Drones
Cheer Shoppers, Annoy Neighbors, Scare Dogs”, WSJ 2018 [7]). In this article, drone noise
is indicated as the main obstacle to widespread public acceptance of this technology in
residential areas. Furthermore, Torija et al. [8] also identify the noise produced as the
main impediment to the spread of drones in urban areas, even though, particularly in the
delivery sector, they are perceived as more environmentally friendly than classic delivery
trucks [3,8]. Moreover, the problem cannot be underestimated as it is also related to the
health of the people involved. In fact, it is known that exposure to aircraft noise might be
a significant cause of community reaction and social disturbance. Using a definition of
health that includes both mental and social well-being, it is true and a well-known fact that
being exposed to aircraft noise causes ill health. Several studies indicated that aircraft noise
exposure can be associated with a prevalence of psychological and psychiatric symptoms.
Studies show a strong link between aircraft noise and sleep loss and awakenings [9]. These
effects can be a further motivation to find a way to reduce noise generated by UAVs.
Despite the clear drawbacks related to acoustic emissions, drones are earmarked to
transform the marketplace of deliveries and civil urban transports, speeding up delivery
times and reducing costs, which is what companies are betting on them [7]. The global drone
market, according to the reports of Drone Industry Insight, amounted to ∼USD 14.1 billion
in 2018 with the prospect of growth by almost 3 times in 2025 [10]; such growth is expected
despite pandemic issues. Just to give an overview, in the USA, there were 1.32 million
UAS registered for leisure purposes and 0.385 million for professional applications in
2019, and these numbers are expected to increase by 12% to 1.48 million for the former
applications and 115% up to 0.83 million for the latter in 2024 [11,12]. A drawback of such
fast growth is that this technology is expected to encourage innovations that may disrupt
existing industries.
The interest in this topic can also be seen in the European Union U-space project.
U-space is a set of new services designed to guarantee safe, efficient and secure access
Fluids 2022, 7, 279 3 of 23
to airspace under 150 m for a great number of drones. This would facilitate any kind of
routine mission in all classes of airspace and all types of environment.
One additional and often overlooked drone application is the monitoring and scouting
of wildlife. The impact of UAVs on the animal population has been the object of recent
research [13]. What the studies have found is that drones constitute a potential new source
of anthropogenic disturbance, which depends both on UAV configurations themselves
and on additional environmental factors. Mulero-Pazmany et al. [13] suggest that animal
reactions are not only influenced by the magnitudes of the noise levels but also by the
sound intermittency and timbre. In the case of UASs, these changes in intensity may be
associated with aircraft on-flight engine variations due to sudden trajectory changes or
due to wind gusts, which has led to the extension of the aeroacoustic problem to unsteady
regimes. Noise signature has been additionally addressed as one of the main influencing
parameters on both human and animal behaviour [14]. Long-term exposure studies based
upon the acoustic emissions of UASs have yet to be performed. Even marine mammals
could be negatively affected by noise emission from UAVs. Christiansen et al. [15] and
also Smith et al. [16] demonstrate that drone noise and visual cues are the main problems
for the utilization of MAVs in wildlife science. These situations require drones to fly
at close range (less than 10 m), increasing the risk of disturbance for marine mammals.
On-air recording showed that the noise level generated by UAVs (they considered two
commonly used drones in marine mammal research) were within the level known to cause
disturbance in some animals. However, according to recent studies, the physiological and
behavioural aspects associated with psycho-acoustic stress [9] are expected to potentially
cause relatively higher energy expenditures, decreases in reproduction and survival, and
space-use changes, which might compromise the average fitness or even viability of certain
populations. Researchers at NASA Langley performed a test aimed at understanding the
psycho-acoustic effects on 38 subjects related to exposure of different recorded vehicle
sounds, in particular of small UAVs [17]. The results reveal a difference of 5.6 dB between
UAVs and the considered road vehicles. This can be interpreted to show that UAVs were
5.6 dB lower in the A-weighted sound exposure level (SEL) than road vehicles but were
perceived to be equally as annoying. The experiments of Torija et al. [8] suggests UAV
flight paths along road infrastructure. In this way, the road traffic noise could mask the
drone sound emissions and alleviate the annoyance for the residents.
In this complex framework, there is a clear need for legislation and standardisation.
The ANSI Standardization Roadmap of 2020 [11,18] reports that, at present, no specific
standards for drone are available, in particular regarding UAV acoustic emissions. Limiting
the noise pollution at an early stage is fundamental, even more so because the population
is still forming its opinion about drones [19]. In 2011, the International Civil Aviation
Organization (ICAO) in Circular 328 identified the noise radiated from drones as a possible
problem in the future [20]: “As new products and aircraft come into use, it may become
apparent that additional noise and emission standards are necessary.”
There is interest in this topic from both the academic and industrial spheres. The
main manufacturer moving to design a silent configuration is DJI, which designed the
Mavic low-noise propeller, which seems to reduce noise to almost 60%, measurable at 4 dB.
In addition, Master Airscrew has designed a low noise propeller for the DJI Mavic Air
that generates a low-pitched sound compared to the original props. The new designed
propeller reduces the aircraft noise by up to 3.5 dB and increases the flight time by 12%,
which means 2.5 min of extra flight time for the standard Mavic Air battery. From an
academic point of view, different research groups are focusing on UAV noise. The main
examples are the University of Southampton, the Institut Supérieur de l’Aéronautique et de
l’Espace (ISAE-SUPAERO) and Niccolò Cusano University. The University of Southampton is
working mainly on leading edge modifications to reduce interaction noise [21,22]. Niccolò
Cusano University is focusing on trailing edge modification to reduce the broadband
noise component generated by propellers [23–25]. Other research groups working on this
topic are, for example, Penn State University; which is working on reducing noise from
Fluids 2022, 7, 279 4 of 23
quadricopters. Finally, NASA is pursuing two projects concerning UAVs: advanced air
mobility (AAM) and revolutionary vertical lift technology. The goal is to develop tools
for designing aircraft to meet noise levels. The obtained data will also be useful to define
and optimise AAM and low-noise flight paths according to community needs and assist
the Federal Aviation Administration (FAA) in setting policy. Furthermore, the number of
research groups working on this topic has increased.
For propeller-driven aircraft, the main noise sources are the engine and the propeller
itself, and this problem strongly affects the low Reynolds regime as well. Therefore, to
reduce a drone’s noise signature, the only way to proceed is to optimize both components
at the same time. For this reason, in recent years, there has been renewed interest in the
first aeronautical propulsion device: the propeller. Rotor noise is becoming a very central
issue because of its several fields of drone applications.
Several authors investigated the aerodynamic performance of small-scale propellers,
including [26–31], and the main evidence from both the experimental and numerical
point of view is that UAV propellers are less efficient than their full-scale counterparts.
Such an effect is related to the low Reynolds number in which they operate. Due to
constraints in size and power density, MAVs are typically equipped with electric motors,
which contribute to simplifying operations and significantly reducing the mechanical noise
signature. As reported in literature, a greater benefit is achieved through the usage of
brushless motors [32]. Huff and Henderson [33] experimentally investigated the engine
noise contribution by using various motors on which the same propeller is mounted.
Results shows a small broadband component buried in the tonal counterpart in the range
of 4.5–5.5 kHz in the frequency domain. In addition, they found that, for typical drone
loading conditions, the most important noise source is the propeller (which is the main
focus of the paper).
In the last few years, the reduction of noise from the propulsive system of small
rotors has been the subject of several works in the literature [32,34–41]. Propeller noise
reduction requires particular attention in the design process because the achievement of
an aeroacoustic optimum may affect the generation of aerodynamic forces. While several
previous works focused on relatively high-Reynolds-numbers propeller, few studies have
focused on low-Reynolds small-scale propellers. For the latter kind of propellers, especially
in hovering conditions, where a considerable area of the rotor is subjected to stall and to
self-interaction with its own slipstream, the effect of the flow features, such as recirculation
bubbles, stall cells and non-uniform boundary layer transitions, are exacerbated. The
presence of the aforementioned flow structures, associated with laminar separation or
three-dimensional spanwise flow non-uniformity, contributes to a reduction in effective
loading and an increase in the unsteadiness at the blade edge. For these reasons, small-
scale UAVs provide a great challenge to the task of noise characterization and prediction.
Indeed, the main noise sources remain consistent with those associated with helicopters,
but there are numerous unknowns that could be investigated, such as, for example, the
effect of reduced size and the balance between tonal noise and broadband noise. Small-scale
propellers are characterized by diameters up to 300 mm and chord lengths on the order of
50 mm. In addition, their rotational speed generally ranges from 1440 [35] to 10,000 RPM [42],
depending on the desired thrust. Furthermore, one of the main differences between small
UAVs and conventional rotorcraft is the flow speed regime in which they fly, measured by
the chord-based Reynolds number at 75% span:
0.75R ρ∞ Ω c
Rec = (1)
µ∞
where Ω is the rotational regime, R is the rotor tip radius, ρ∞ is the air density, c is
the rotor blade chord and µ∞ is the air dynamic viscosity. For a full-scale helicopter, a
representative Rec is on the order of 106 , while for a UAV, it may range from 104 to 105 . The
corresponding tip Mach number are on the order of 0.3. In terms of conventional flat plate
aerodynamics, the former Reynolds number explicates in a turbulent flow regime, while
Fluids 2022, 7, 279 5 of 23
the latter explicates in a laminar-transitional flow regime [43,44]. This discrepancy calls
into question the applicability of traditional noise prediction tools. When summarizing the
different contributions, the literature shows that the main noise sources in a UAV propeller
are: tonal self-noise, generated by the volume displacement and aerodynamic loading on
the blade; the incoming flow turbulence at the blade leading edge (i.e., LE noise from highly
turbulent flow in harsh environments); the interaction of the boundary layer with the blade
trailing edge (i.e., TE noise due to turbulent boundary layers but also due to unsteady flow
separation of re-circulation bubbles, etc.), flow separation of the flow on the different blade
sections (i.e., stall and flow separation noise), blade vortex interaction (i.e., BVI due to the
interaction of a rotor blade with the shed tip vortices from a previous blade) and blade
wake interaction (BWI, which occurs when the turbulent wake impinges on a subsequent
blade) [35,45]. Predicting and reducing the noise radiation from these contributions is even
more complicated due to the variety and sensitivity of the noise to the flow field. These
reasons clarify the complexity of the problem and the importance of improving knowledge
in this field. The relative importance of the noise sources listed here depends significantly
on the specific operating conditions [46]. For a schematic representation of the main noise
sources involved in these kind of applications, see Figure 1. In the literature, a few studies
have been devoted to the analysis of the noise due to the interaction between the propulsive
system and the airframe in the case of small propellers. Zawodny et al. [47], in their
experimental analysis, found that the presence of airframe surfaces is a not-negligible noise
source. In fact, it generates noise levels analogous or even greater than the rotor blade
surfaces in particular rotor tip conditions. This study analyzed the effects of both airframe
to rotor distance and airframe size. Results show prominent tonal peaks in the Fourier
domain related to airframes in the case of close proximity between the airframe and the
rotor plane. This effect seems to decay rapidly if the rotor-airframe distance increases. Even
the airframe shape seems to influence noise generation. Generic constant cross-section
systems were found not to affect noise generation in the plane of the rotor. Instead, a
conical airframe shows an increase in the tonal noise component. The present manuscript
is organized as follows. In Section 2.1 there is a brief explanation of the most important
noise sources for rotors. Section 2.2 reviews the state of art of passive control strategies
currently in use. Finally, Section 3 draws conclusions and provides a brief overview of
future configurations.
where p0TN is the tonal component of pressure fluctuations, and p0BB is the broadband
counterpart. Tonal components are directly related to the periodic motion of the blade in
the surrounding fluid. Therefore, the frequency and magnitude of the radiated noise is
Fluids 2022, 7, 279 6 of 23
related to the rotational velocity. The physical mechanism associated with the production
of the tonal contributions is related to blade thickness and to aerodynamic loading. On
the other hand, broadband noise is radiated by the interaction of turbulent flow structures
with the blade edge. Therefore, it is either generated at the blade leading/trailing edge or
at the blade tip. The theoretical prediction of the periodic noise generated by propellers
is based on the solution of the Ffowcs, Williams and Hawkings non-homogeneous wave
equation, known as the Ffowcs–Williams/Hawkings equation [36,50].
1 ∂2 ( p 0 ) ∂2 ( p 0 ) ∂2 Tij
∂ ∂f ∂f
· − = + ρ a · v i · δ ( f ) · − ∇ ∆p ij · δ ( f ) · (3)
a2 ∂t2 ∂xi2 ∂xi · ∂xi ∂t ∂xi ∂xi
where:
• a is the speed of sound;
• p0 is the perturbation on the static pressure;
• t is the observer time;
• xi are the components of the position vector;
• Tij are the components of the Lighthill stress tensor;
• ρ a is the air density;
• vi is the components of the source velocity vector;
• δ is Kronecker’s delta function;
• f is a function that defines the surface of the body producing the pressure wave;
• pij are the components of the generalized stress tensor.
In this equation, there are 3 forcing terms on the right-hand side, which are related
to vortex, thickness and loading, respectively. For thin blades and low Mach numbers
(M < 1), the vortex term is negligible, and the narrow band contribution is given by the
sum of a sound source related to blade thickness p0T and one related to aerodynamic loading
p0L as the distributed force over the blade. Since it is possible to further decompose p0TN ,
we obtain:
p0TN (x, t) = p0T (x, t) + p0L (x, t) (4)
The thickness term takes into account the fluid displacement due to the body, while the
loading counterpart takes count of the unsteady force distribution over the body surface.
A numerical evaluation of these two quantities can be achieved by discretizing the blade in
N finite elements along the span. The resulting overall radiation field can be approximated
as the sum of N pointwise sources.
N
p0L (x, t) = ∑ pl,k (x, t) (5)
k =1
N
p0T (x, t) = ∑ pt,k (x, t) (6)
k =1
2
Φk
ρ ∂
p0T,k (x, t) = (8)
4π ∂2 τ 2 rk (1 − Mr )
where:
Fluids 2022, 7, 279 7 of 23
• r̂k is the position vector of an observer relative to the k-point noise source |rˆ | = 1 ,;
• Fk is the aerodynamic force on the k-point blade element of volume Φk ;
v
• Mrk is a scalar magnitude that represents the component of the Mach vector Mk = a
on rk .
Figure 2. Representation of the reference coordinate system considered for the definition of the
aeroacoustic model.
where p0TE is the TE component, p0LE is the LE counterpart and p0S is the separation term.
Several authors have addressed the prediction of trailing-edge broadband noise in literature.
See, among many studies, Sinibaldi et al. [35], where a relationship for calculating the
power spectral density at the trailing edge is reported:
B ωc 2
S TE
pp (r, θ, ω ) = ∆R D (θ, φ) | I |2 Φ pp ly (11)
8π 2ar
where c is the chord, ∆R is the spanwise length of the blade, I is the radiation integral
function, B is the number of the blades, ω = 2π f is the angular frequency, f is the rotational
frequency, D (θ, ϕ) is the directivity function and Φ pp is the wall power spectral density
of the pressure fluctuations. The wall pressure spectral density S TE pp and the spanwise
correlation length ly can be evaluated experimentally or numerically. There are different
models for S TEpp estimation, e.g., the one proposed by Schklinker and Amiet [52], or the
more recent model proposed by Rozenberg et al. [53], which takes into account the effect of
Fluids 2022, 7, 279 8 of 23
the adverse pressure gradient. On the other hand, for ly evaluation. the most used model
is the Corcos’ model [54]. The effect of the flow separation on broadband noise can be
significant as well. According to [40], an estimation of the power spectral density related to
that source is provided by the following expression:
( 3 )
ω 2 z 2 c2 2πU
sep 2 2 2 D −7
S pp (x, ω ) = ρ cU AS 8.6 · 10 (12)
4πar r 4 ωc(1 − Mr )
where c D is the drag coefficient, AS is the body cross-sectional area where separation is
localized, and U is the velocity of the flow.
the different disciplines are addressed simultaneously. In this case of study, aerodynamic,
structural and acoustic problems were analysed at the same time.
One of the most interesting MDO models in the literature was presented by Gur and
Rosen [32,36,62,63] and developed to reach the best compromise between the opposite
requirements of efficiency and quietness. In particular, the target of this design process
was first to mitigate the tonal component of the noise, dependent on the actual loading
of the blade (as seen in Section 2.1). In [36], the cost function J was based on the sound
pressure level (SPL). However, the presence of power and stress constraints were taken
into account. The first step was to optimize only the acoustic footprint of the blade, and
MDO results provided for a blade with a very large chord and relatively small radius. This
is, of course, unfeasible in a small rotor due to the power required by such a non-optimal
aerodynamic design in combination with the noise increase from the electric engine to
deliver such power at the hub. Furthermore, a limit on the extracted power from the battery
produces a significant increase in propeller noise. Instead, stress constraints lead to an
increase in cross-sectional thickness and rotational speed. These results clarified the need
for a multidisciplinary optimization. In fact, the presence of both structural and acoustic
constrains is fundamental to achieve feasible results.
This model was enhanced in [32], where the propeller design model was extended to
the entire propulsion system. In other words, a model for electric motor and battery was
added to the previous model. For this purpose, theoretical models of these components
were required. The models presented were based on a comprehensive investigation of
existing motors and batteries. The performance of the vehicle greatly depends on the
interaction between propeller, electric motor and battery. Clearly, then, it is important to
study these three components contemporaneously.
By using the Gur and Rosen model [36], Sinibaldi and Marino [35] employed a quiet
propeller and carried out an experimental analysis to characterize its behaviour as compared
to a conventional one (conventional in the sense of a propeller not specifically designed
to achieve noise reduction). In their study, the focus was on the optimization of the chord
distribution along the span-wise direction. The results of the comparison of the two
propeller showed that, by using the MDO approach, significant noise reduction can be
achieved, at least for the narrow-band contribution. An unexpected result is that by increasing
rotational velocity in order to achieve high thrust values, strong vibrations occur that can
be ascribed to the increased thickness of the optimized blade. This phenomenon produces
noise that make the optimized propeller comparable to the conventional one.
Furthermore, Serrè et al. [21,22,38] presented a model for reducing the noise of a
MAV rotor while preserving the endurance. The aim was to develop an optimisation
algorithm that is applicable in the industrial field, especially in terms of computational
cost. The method was based on low-order computational tools for the estimantion of
the aerodynamic loading and the tonal and braodnband noise. These tools were coupled
with the optimization algorithm in order to define a law for chord and twist distributions
of the blade. A relevant reduction of 8 dB(A) was measured. Results showed that the
optimized propeller had a higher aerodynamic efficiency. Such an effect permits to reduce
the rotational speed resulting in a lower main frequency of the tonal noise, a lower intensity
of the small turbulent eddies (which are related to high frequencies of ingestion noise) and
a lower tip Mach number (which implies a lower intensity of the radiated noise).
Moreover, Pagliaroli et al. [34] used an MDO approach in order to assess the effect of
the pitch angle on MAV noise signatures. An experimental analysis was carried out in order
to evaluate aeroacoustic behaviour. The experimental tests were carried out on a propeller
of 2 [mm] with a twist angle of zero. The blades were mounted on a collective pitch in
order to vary the pitch angle from 0 to 21 [deg]. All the measurements were taken at the
anechoic chamber of the Office National d’Études et de Recherches Aérospatiales (ONERA).
The optimization strategy seemed to be useful in reducing the number of variables in
the multiphysics problem. Furthermore, wall pressure measurements confirmed that
Fluids 2022, 7, 279 10 of 23
the pressure signature was dominated by the broadband component generated by the
separation bubble, showing that it is important to extend studies to broadband noise.
Wisnieski et al. [64] developed a propeller design program in order to construct quiet
propellers with improved efficiency. This approach enabled to design a propeller, evaluate
its performance and, finally, to create a 3D model for additive manufacturing. Such an
optimization approach lead to an high pitch propeller with low aspect ratio, reducing
the chord length this propeller has a larger surface which guarantee the desidered thrust
at lower rotational speed. The obtained propeller was characterized by 5 blades with
an oval tip, which proved to be 12 dB quieter than the stock commercial DJI phantom
II propeller and 6% more aerodinamically efficient. This investigation put emphasis on
tip geometry as a powerful strategy to reduce noise. In the context of propeller shape
optimisation, it is also interesting to consider optimal propeller blade spacing as in the
study of Cattanei et al. [65]. This study presents a general optimization strategy based on
the evaluation of the rotor interference function, which may be applied to rotors of arbitrary
blade number. By imposing unequal spacing between the blades, it is possible to reduce
the tonal noise component without having any effect on thrust generation. Based on their
results, a propeller has been realized and tested in semi-anechoic chamber to validate the
interference model. The experimental results show a good agreement, confirming that the
proposed algorithm is reliable for the design of a propeller.
(a) (b)
(c)
Figure 3. 3D rendering of the propeller blade: (a) baseline; (b) serrated trailing edge; (c) boundary
layer tripping system.
The idea for this control strategy was inspired by nature, in particular by the silent
flight of owls [66–68]. Owls are known to be one of the most silent predators in nature.
The quietness of their flight is due to their characteristic wings, with three main physical
features: (i) a suction wing surface with a soft downy coating, (ii) a comb of stiff feathers at
the wing leading edge, and (iii) TE feathers and wings with a fringe of flexible filaments.
The sawtooth pattern employed by manufacturers is the simplest geometric way to mimic
the permeability of owls’ wings.
Chong et al. [69] and Avallone et al. [70] focused on wind turbine applications. The first
study involved an experimental analysis on a flat plate, while the second was a numer-
Fluids 2022, 7, 279 11 of 23
Pang et al. [55] conducted an experimental analysis of the effect of pitch angle and
trailing edge serration on a small rotor. Results showed that sawtooth serrations employed
at the TE can noticeably suppress broadband noise in the high frequency region in the
far-field. The main drawback observed was that the tonal component seems to increase
in the low-frequency region. At low velocities, serrations seem to lead to greater noise
reduction. Such an effect suggests that STE is a potential solution for reducing UAV noise
when propellers are sure to operate at low speed. Near-field experiments shed light on
sound field characteristics, exhibiting a radial decay of SPL in the propeller rotation plane.
Ning et al. [72] also carried out an experimental analysis on STE, the aim being to
reduce noise while maintaining the thrust constant. This work defined three parameters
that ensure a beneficial employment of serration for noise reduction. The parameters
considered were:
• The non-dimensional tooth height, defined as the ratio between the tooth half-height
and the boundary layer thickness h∗ = h/2δ;
Fluids 2022, 7, 279 12 of 23
• The aspect ratio of the tooth, defined as the ratio between the width and the half-height
ARt = 2b/h;
• The boundary layer thickness-based Strouhal number Stδ = f δ/U.
The geometrical parameters employed are defined in Figure 5. Ning pointed out that
to achieve noise reduction, h∗ > 0.25. Otherwise the amplitude of the serration is too
small, and as a result, the turbulent eddies go beyond the sawtooth without significant
interaction. Furthermore, the inclination angle α (see Figure 4) must be lower than 45◦ ,
and this fact is guaranteed by imposing ARt < 4. In the definition of the Strouhal number,
f is the sound frequency, δ is the boundary layer thickness and U = 0.7 ∗ Urel by having
called Urel the relative velocity. This non-dimensional coefficient has to be greater than 1
(as stated in Howe’s theory), which means f > U/δ, in order to obtain a significant noise
reduction. Experiments have been carried out at Re > 1.5 × 105 . The results show that when
f > U/δ, noise reduction appears at a frequency lower than U/δ, while the overall noise
level increases. Therefore, this parameter gives the frequency range in which it is possible
to find noise reduction. This work analyses four rotors by varying the ARt coefficient. The
analysis involves aerodynamic and aeroacoustic measurements to characterize wake flow
statistics. The results show that the STE can reduce broadband noise in the high frequency
region without any loss in aerodynamic performance, while, in the low frequency region,
the noise generated is almost the same. Measurements also show that, in order to keep the
thrust constant, a higher rotational velocity of the propeller is required.
Figure 5. Sketch of the serrated trailing edge. In the enlargement of figure, the main geometrical
parameters of the tooth are reported: tooth basis b, height h and the serration angle α.
Intravartolo et al. [49] alslo carried out an experimental analysis on STE by focusing
on the serration depth effect. Results showed that an increase in serration depth produces a
reduction in the intensity of the trailing edge wake. Nevertheless, benefits from the depth of
the serrations diminished with respect to the overall noise signature of the propeller. When
serration depth reaches a value comparable to half of the mean aerodynamic chord (MAC),
no further gain in aeroacoustic effect can be observed. On the contrary, an increase in the
overall noise may occur due mainly to aerodynamic effects. Serration depth effect was also
analysed by Pagliaroli et al. [23–25], in particular, as regards broadband noise component
and the directivity of the noise source in the near-field. A notable reduction in the noise
generated was obtained in the low-frequency region, and damping in the tails of the
probability density function (PDF) was observed. The statistical analysis shed light on the
physical phenomenon that lies behind the noise reduction. PDF tails are commonly related
to intermittent structures in the pressure field; as a result, serrations seem to eliminate
strong energetic events. The drawback of serration is a loss in aerodynamic efficiency,
so the optimal geometry had to be found. An analysis of the directivity showed that the
sawtooth pattern effect is bounded in the polar angle range θ = [60◦ :120◦ ] (the polar angle
considered is defined in Figure 6).
Fluids 2022, 7, 279 13 of 23
Anechoic Chamber
Microphone
Θ=90°
Θ=180° Θ=0°
Θ=-90°
Figure 6. Definition of the polar reference system for the directivity analysis.
because they work as a vortex generator and can mitigate velocity fluctuations and also
modify the laminar-turbulent transition on the propeller suction surface.
In addition, Chaitanya et al. [77,78] performed an experimental analysis regarding
the use of leading-edge serration on a flat plate. More in detail, in [78], an optimum
geometry for the serration was found; the maximum noise reduction was achieved when
the turbulence integral length scale corresponded to half of the serration wavelength.
Moreover, in [77], an improvement of serration geometry was reported; specifically, three
innovative serration geometry were defined and investigated. These configurations involve:
(a) a double-wavelength obtained by adding two components, one of which has twice the
frequency of the other; (b) a chopped peak, in which the source at the peak is increased
by clipping the peak so that it generates destructive interference with the root source;
(c) a slitted root leading edge obtained by adding a narrow slit at the root position. The
results show that the main advantages are achieved by means of the double-wavelength
serrations even if the other two geometries can lead to a not negligible noise reduction. It is
important to emphasise that this control strategy has no effect on aerodynamic performance.
Moreover, considering a flat plate, Narayan et al. [79] investigated the noise reduction
in the case of sinusoidal serrations. The results showed that by using this geometry, the
broadband noise can be significantly reduced, in particular in the mid-frequency range
(500 Hz–8 kHz). It was also found that the serration amplitude is the factor that influences
the sound emissions more.
A numerical model to investigate the effect of leading edge serration was reported
in [80] based on Amiet’s approach, which also makes it valid for high-mach-number
applications, where leading edge noise is a common problem. The numerical simulation
performed showed that the noise reduction is related to a destructive interference of the
scattered pressure induced by the serrations. They also defined geometrical parameters
that ensure the noise effect.
Hersh et al. [81] investigated the use of LE serrations on a stationary and rotating
NACA 0012 airfoil. They demonstrated that a sensible reduction in tonal noise (approxi-
mately between 4 and 8 dB) can be achieved, and it is related to wake vortex shedding at
high incidence.
chord, there are smaller contributions from low frequencies and increased high-frequency
contributions. Consequently, the tonal noise connected with these two phenomena seems
to be mitigated. This passive control technique seems very interesting because it should
not affect the aerodynamic properties of the propeller, but rather reduce the drag force.
In the literature, several authors have studied the effect of porosity on trailing edge
noise [92–94]. Rubio Carpio et al. [92] focused on a flat plate with different types of inserts.
The porous inserts, covering 20% of the chord, are manufactured with metal foams of cell
diameters dc = 450 (µm) and dc = 800 (µm) and permeability values of 6 × 10−10 and
2.7 × 10−9 m2 . The far-field measurements show low-frequency noise attenuation of up
Fluids 2022, 7, 279 16 of 23
to 7 and 11 [dB], respectively, for the first and second permeability value. On the other
hand, in the high-frequency region, an increase in noise up to 8–10 dB was observed. This
phenomenon is due to surface roughness. Increasing permeability also led to a reduction of
the frequency range affected by noise attenuation. A PIV measurement campaign shows
an increase in BL thickness δ and in displacement thickness δ∗ for the metal foam insert
with higher permeability. Analysis in the Fourier domain shows that the attenuation in
velocity fluctuations affects mostly the low-frequency region, suggesting that turbulence
intensity reduction may be one of the changes that contributes to noise reduction. On the
other hand, the results do not show an increase in high-frequency fluctuation content as
regards the solid case. Showkat Ali et al. [93] demonstrated that porous TE can delay
vortex shedding and significantly increase vortex formation length, leading to a very low
turbulent near-wake region. The usage of porous material also leads to significant lateral
coherence reduction of the turbulent structure.
Ref. [94] documents acoustic test on an airfoil with porous treatment at the TE. Maxi-
mum noise reduction reached was up to 2–6 dB. This effect may be attributed to a material-
dependent pressure field generated in the near-field related to the flow resistivity of the
TE material.
A very interesting control technique is the use of Poro-Serrated TE [95–97], which
combines the serrated TE (Section 2.2.2) and the effect of porosity. These poro-serrated
TE devices contain porous materials of various air flow resistances at the gaps between
adjacent members of the serrated sawtooth. The object of this study was to understand if
two control strategies for noise mitigation can co-exist, one related to the oblique edges
introduced by the serrations, the second arising from porosity, which allows the pressure
side and suction side to communicate, thereby reducing the acoustic dipole strength at the
trailing edge. In these studies, the focus was on a flat plate with a serrated trailing edge
with the addition of porous foam between adjacent members of the sawtooth. The porous
foam was cut in order to match perfectly with the volume and shape of the sawtooth gaps,
thus preserving the original airfoil profile. This technique can simultaneously suppress
vortex shedding and reduce broadband noise. Results showed that multiple broadband
noise reduction mechanisms occur (serration + porous material), but it is likely that the
porous material is enhancing the serration effect, rather than the porous material exerting
an effect of its own.
2.2.6. Metamaterials
One way of achieving sound attenuation is the application of a sound barrier that
reflects or absorbs the incident acoustic energy. Such a solution cannot be directly applied
on MAVs because it eliminates the passage of air. To be employed on a rotor blade, it
is important to guarantee permeability to air by designing a ducted propeller. In recent
years, there has been a fast growth in metamaterial science, leading to new solutions for
manipulating acoustic energy. Metamaterials are composed of subwavelength structures
since their effective acoustic properties are governed by their structural shape rather than
their constitutive properties.
Ghaffarivardavagh et al. [98] presented a design methodology for an ultra-open meta-
material (UOM) composed of subwavelength unit-cell structures featuring a predominately
open area. The designed UOM works as a high-performance selective sound silencer for
applications where both sound attenuation and highly efficient ventilation are required.
The proposed method is based on Fano-like interference [99] for selective attenuation
of acoustic waves. The first part of their studies aimed to analytically demonstrate that
Fano-like interference is present in a transverse bilayer metamaterial. Then, the feasibility
of the metamaterial structure was proved by providing both analytic and experimental val-
idation. The designed UOM consist of two distinguishable regions: a central open part and
a peripheral helical part. The two regions are characterized by different acoustic properties.
Figure 8 reports a 3D representation of it. The contrast in the acoustic properties of the two
regions has been proved to be essential to achieve the required silencing functionality. The
Fluids 2022, 7, 279 17 of 23
(a)
(b)
Figure 9. Representation of the considered baseline blade (a) and of the new bio-inspired blade (b).
Finally, Noda et al. [102] developed a new quiet propeller inspired by owl wing
morphologies. A DJI Phanton 3 propeller was employed as the reference propeller, and
different structures were attached to the trailing edge of the propeller. The main interesting
results came from an aluminium plate attached at the trailing edge. The first step in order
to reduce sound emissions was to employ a greater propeller. The strategy herein proposed
showed a reduction in noise of almost 2 dB with no effect on the power consumption. This
effect could be related to the reduction in the rotational speed, while the rotational speed to
maintain a certain operating condition will be lower with a smaller radius propeller.
3. Conclusions
This paper focuses on the noise generated by small rotors, the aim being to identify
which passive noise control strategies can be employed on a drone propeller. The main
noise sources for this application concern the interaction between the BL and the TE of
the blade. This paper presents several strategies to control this noise source. Even though
noise control is the main focus, aerodynamic performance is also taken in count in order to
guarantee the success of the mission.
The first strategy to reduce noise emission was to employ an optimized geometry by
taking into account acoustic constraints in the multi-disciplinary optimization process.
These solutions led to a blade geometry that reduces noise for a specific operating config-
uration, so it is not sure that in other configurations, the behaviour would be exactly the
same, both in terms of thrust and noise generation. The effect of chord distribution and of
pitch angle were analyzed, indicating significant noise reduction, but the drawback was a
loss of aerodynamic features.
By taking ideas from nature, in particular the owl wing, innovative blade geometries
may be employed, with the most effective, seemingly, being the application of serration.
The most investigated configuration is the sawtooth pattern at the TE, which has been
shown to reduce the broadband noise component. From a theoretical point of view, this
effect can be related to a reduction in coherent structures in the pressure field. This
assumption is confirmed by statistical analysis, which shows lower PDF tails when serration
is employed at the TE. Another configuration involves the use of fractal serration, with the
Fluids 2022, 7, 279 19 of 23
effect, even for this configuration, being related to the interaction between the coherent
structures and the serration.
Serrations can also be employed at the leading edge. Through this technology, it seems
possible to achieve a significant reduction in noise and also an improvement in aerodynamic
performance. Leading-edge serrations act as vortex generators, limiting velocity fluctuations
and anticipating the laminar-turbulent transition of the boundary layer; such effects result
in noise reduction.
Another strategy to reduce TE noise is the use of porous materials. The effect of poros-
ity on rotor noise has been studied principally for wind turbines but could be interesting
for UAV rotors as well. In fact, it has been proved that porosity produces a reduction
in turbulence length scale. The next step is to employ this technology on a small-scale
propeller and test it. Furthermore, metamaterials can be designed as highly efficient sound
barriers for a target frequency. The development in metamaterials science may, in a few
years, lead to the realization of a ducted propeller with specific sound characteristics.
On the other hand, in order to reduce the tonal noise component, a boundary layer
tripping system can be applied on the suction side of the propeller blade in the form of a
simple adhesive aluminium strip. By using this system, laminar to turbulent transition
is forced at 5% of the chord. This effect results in broadband noise radiating from the TE
in the high-frequency region and seems to have no effect on thrust generation; rather, it
should reduce drag force since the efficiency of the propeller should increase.
Finally, this paper presented an innovative type of geometry inspired by nature. This
particular geometry mimics the planform shape of cicada wings and maple seeds [100].
The experimental results show that this bio-inspired wing can provide the same thrust as
a baseline propeller. Additionally, a reduction in loading noise was observed and can be
attributed to the reduction in standard thrust deviation.
Author Contributions: Investigation, Formal Analysis and Writing, P.C.; Investigation and Writing,
D.R.; Conceptualization and Writing, T.P. All authors have read and agreed to the published version
of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.
Abbreviations
The following abbreviations are used in this manuscript:
References
1. Floreano, D.; Wood, R.J. Science, technology and the future of small autonomous drones. Nature 2015, 521, 460–466. [CrossRef]
[PubMed]
2. Figliozzi, M. Lifecycle modeling and assessment of unmanned aerial vehicles (Drones) CO2e emissions. Transp. Res. Part D
Transp. Environ. 2017, 57, 251–261. [CrossRef]
Fluids 2022, 7, 279 20 of 23
3. Yoo, W.; Yu, E.; Jung, J. Drone delivery: Factors affecting the public’s attitude and intention to adopt. Telemat. Inform. 2018,
35, 1687–1700. [CrossRef]
4. Koiwanit, J. Analysis of environmental impacts of drone delivery on an online shopping system. Adv. Clim. Change Res. 2018,
9, 201–207. [CrossRef]
5. Goodchild, A.; Toy, J. Delivery by drone: An evaluation of unmanned aerial vehicle technology in reducing CO2 emissions in the
delivery service industry. Transp. Res. Part D Transp. Environ. 2018, 61, 58–67. [CrossRef]
6. EASA. Introduction of a Regulatory Framework for the Operation of Unmanned Aircraft Systems in the ‘Open’ and ‘Specific’ Categories;
Opinion No 01/2018; EASA: Nairobi, Kenya, 2018.
7. Cherney, M. Delivery Drones Cheer Shoppers, Annoy Neighbors, Scare Dogs. The Wall Street Journal, 26 December 2018.
8. Torija, A.J.; Li, Z.; Self, R.H. Effects of a hovering unmanned aerial vehicle on urban soundscapes perception. Transp. Res. Part
Transp. Environ. 2020, 78, 102195. [CrossRef]
9. Morrell, S.; Taylor, R.; Lyle, D. A review of health effects of aircraft noise. Aust. N. Z. J. Public Health 1997, 21, 221–236. [CrossRef]
10. Kapustina, L.; Izakova, N.; Makovkina, E.; Khmelkov, M. The global drone market: Main development trends. In Proceedings of
the SHS Web of Conferences, EDP Sciences, Muscat, Oman, 15–16 November 2021; Volume 129, p. 11004.
11. Schäffer, B.; Pieren, R.; Heutschi, K.; Wunderli, J.M.; Becker, S. Drone Noise Emission Characteristics and Noise Effects on
Humans—A Systematic Review. Int. J. Environ. Res. Public Health 2021, 18, 5940. [CrossRef]
12. FAA. FAA Aerospace Forecast, Fiscal Years 2019–2039; Technical Report; FAA: Washington, DC, USA, 2022.
13. Mulero-Pázmány, M.; Jenni-Eiermann, S.; Strebel, N.; Sattler, T.; Negro, J.; Tablado, Z. Unmanned aircraft systems as a new source
of disturbance for wildlife: A systematic review. PLoS ONE 2017, 12, e0178448. [CrossRef] [PubMed]
14. Ditmer, M.; Vincent, J.; Werden, L.; Tanner, J.; Laske, T.; Iaizzo, P.; Garshelis, D.; Fieberg, J. Bears Show a Physiological but Limited
Behavioral Response to Unmanned Aerial Vehicles. Curr. Biol. 2015, 25, 2278–2283. [CrossRef] [PubMed]
15. Christiansen, F.; Rojano-Doñate, L.; Madsen, P.; Bejder, L. Noise Levels of Multi-Rotor Unmanned Aerial Vehicles with Implications
for Potential Underwater Impacts on Marine Mammals. Front. Mar. Sci. 2016, 3, 277. [CrossRef]
16. Smith, C.; Sykora-bodie, S.; Bloodworth, B.; Pack, S.; Spradlin, T.; Leboeuf, N. Assessment of known impacts of unmanned aerial
systems (UAS) on marine mammals: Data gaps and recommendations for researchers in the United States. J. Unmanned Veh. Syst.
2016, 14, 31–44. [CrossRef]
17. Christian, A.; Cabell, R. Initial investigation into the psychoacoustic properties of small unmanned aerial system noise. In
Proceedings of the 23rd AIAA/CEAS Aeroacoustics Conference, Denver, CO, USA, 5–9 June 2017; p. 4051.
18. ANSI. Standardization Roadmap for Unmanned Aircraft Systems, Version 2.0., Prepared by the ANSI Unmanned Aircraft Systems
Standardization Collaborative (UASSC): June 2020; Technical Report; American National Standards Institute: New York, NY,
USA, 2020.
19. Eißfeldt, H.; Vogelpohl, V.; Stolz, M.; Papenfuß, A.; Biella, M.; Belz, J.; Kügler, D. The acceptance of civil drones in Germany.
CEAS Aeronaut. J. 2020, 11, 665–676. [CrossRef]
20. (ICAO). Cir 328 AN/190, Unmanned Aircraft Systems (UAS) Circular 328; Technical Report; International Civil Aviation Organization
(ICAO): Montreal, QC, Canada, 2011.
21. Serré, R.; Gourdain, N.; Jardin, T.; Jacob, M.C.; Moschetta, J. Towards silent micro-air vehicles: Optimization of a low Reynolds
number rotor in hover. Int. J. Aeroacoustics 2019, 18, 690–710. [CrossRef]
22. Serré, R.; Fournier, H.; Moschetta, J. A design methodology for quiet and long endurance MAV rotors. Int. J. Micro Air Veh. 2019,
11, 1756829319845937. [CrossRef]
23. Pagliaroli, T.; Candeloro, P.; Camussi, R.; Giannini, O.; Panciroli, R.; Bella, G. Aeroacoustic Study of small scale Rotors for mini
Drone Propulsion: Serrated Trailing Edge Effect. In Proceedings of the 2018 AIAA/CEAS Aeroacoustics Conference 2018, Atlanta,
GA, USA, 25–29 June 2018. [CrossRef]
24. Candeloro, P.; Nargi, R.; Patanè, F.; Pagliaroli, T. Experimental Analysis of Small-Scale Rotors with Serrated Trailing Edge for
Quiet Drone Propulsion Experimental Analysis of Small-Scale Rotors with Serrated Trailing Edge for Quiet Drone Propulsion.
J. Phys. 2020, 1589, 012007. [CrossRef]
25. Candeloro, P.; Nargi, R.E.; Grande, E.; Ragni, D.; Pagliaroli, T. Experimental Fluid Dynamic Characterization of Serrated Rotors
for Drone Propulsion. J. Phys. Conf. Ser. 2021, 1977, 012007. [CrossRef]
26. Deters, R.; Ananda Krishnan, G.; Selig, M. Reynolds number effects on the performance of small-scale propellers. In Proceedings
of the 32nd AIAA Applied Aerodynamics Conference, Atlanta, GA, USA, 16–20 June 2014; p. 2151.
27. Deters, R.; Kleinke, S.; Selig, M. Static testing of propulsion elements for small multirotor unmanned aerial vehicles. In
Proceedings of the 35th AIAA Applied Aerodynamics Conference, Denver, CO, USA, 5–9 June 2017; p. 3743.
28. Deters, R.; Dantsker, O.; Kleinke, S.; Norman, N.; Selig, M. Static performance results of propellers used on nano, micro, and mini
quadrotors. In Proceedings of the 2018 Applied Aerodynamics Conference, Atlanta, GA, USA, 25–29 June 2018; p. 4122.
29. Merchant, M.; Miller, L.S. Propeller performance measurement for low Reynolds number UAV applications. In Proceedings of
the 44th AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 9–12 January 2006; p. 1127.
30. Brandt, J.; Selig, M. Propeller performance data at low reynolds numbers. In Proceedings of the 49th AIAA Aerospace Sciences
Meeting including the New Horizons Forum and Aerospace Exposition, Orlando, FL, USA, 4–7 January 2011; p. 1255.
31. McCrink, M.; Gregory, J. Blade element momentum modeling of low-reynolds electric propulsion systems. J. Aircr. 2017,
54, 163–176. [CrossRef]
Fluids 2022, 7, 279 21 of 23
32. Gur, O.; Rosen, A. Optimizing Electric Propulsion Systems for Unmanned Aerial Vehicles. J. Aircr. 2009, 46, 1340–1353. [CrossRef]
33. Huff, D.; Henderson, B. Electric motor noise for small quadcopters: Part 1–Acoustic Measurements. In Proceedings of the 2018
AIAA/CEAS Aeroacoustics Conference, Atlanta, GA, USA, 25–29 June 2018; p. 2952.
34. Pagliaroli, T.; Moschetta, J.; Benard, E.; Nana, C. Noise signature of a MAV rotor in hover. In Proceedings of the 49th International
Symposium of Applied Aerodynamics, Lille, France, 24–26 March 2014; pp. 24–25.
35. Sinibaldi, G.; Marino, L. Experimental analysis on the noise of propellers for small UAV. Appl. Acoust. 2013, 74, 79–88. [CrossRef]
36. Gur, O.; Rosen, A. Design of a Quiet Propeller for an Electric Mini. J. Propuls. Power 2009, 25. 717–728. [CrossRef]
37. JanakiRam, D.; Scruggs, B. Investigation of performance, noise and detectability characteristics of small-scale remotely piloted
vehicle /RPV/ propellers. In Proceedings of the 7th Aeroacoustics Conference, Palo Alto, CA, USA, 15–20 November 1981;
Volume 19, pp. 1052–1060.:10.2514/6.1981-2005. [CrossRef]
38. Serre, R.; Chapin, V.; Moschetta, J.; Fournier, H. Reducing the noise of Micro–Air Vehicles in hover. In Proceedings of the
International Micro Air Vehicle Conference and Flight Competition, Toulouse, France, 18–21 September 2017; pp. 51–59.
39. Leslie, A.; Wong, K.; Auld, D. Broadband Noise Reduction on a mini-UAV Propeller. In Proceedings of the 14th AIAA/CEAS
Aeroacoustics Conference (29th AIAA Aeroacoustics Conference), Vancouver, BC, Canada, 5–7 May 2008. [CrossRef]
40. Nelson, P.; Morfey, C. Aerodynamic Sound Production. J. Sound Vib. 1981, 79, 263–289. [CrossRef]
41. Rozenberg, Y.; Roger, M.; Moreau, S. Rotating Blade Trailing-Edge Noise: Experimental Validation of Analytical Model. AIAA J.
2010, 48, 951–962. [CrossRef]
42. Massey, K.; Gaeta, R. Noise measurements of tactical UAVs. In Proceedings of the 16th AIAA/CEAS Aeroacoustics Conference,
Stockholm, Sweden, 7–9 June 2010; p. 3911.
43. Zawodny, N.; Jr, D.B.; Burley, C. Acoustic Characterization and Prediction of Representative, Small-Scale Rotary-Wing Unmanned
Aircraft System Components . In Proceedings of the 72nd American Helicopter Society (AHS) Annual Forum, West Palm Beach,
FL, USA, 17–19 May 2016.
44. Cabell, R.; Grosveld, F.; McSwain, R. Measured noise from small unmanned aerial vehicles. In Proceedings of the Inter-Noise and
Noise-Con Congress and Conference Proceedings, Hamburg, Germany, 21–24 August 2016; Volume 252, pp. 345–354.
45. Fattah, R.; Chen, W.; Wu, H.; Wu, Y.; Zhang, X. Noise measurements of generic small-scale propellers. In Proceedings of the 25th
AIAA/CEAS Aeroacoustics Conference, Delft, The Netherlands, 20–23 May 2019; p. 2498.
46. Brooks, T.; Jolly, J.R., Jr.; Marcolini, M. Helicopter Main-Rotor Noise: Determination of Source Contributions Using Scaled Model Data;
Technical Report; NASA: Washington, DC, USA, 1988.
47. Zawodny, N.; Boyd, D., Jr. Investigation of rotor-airframe interaction noise associated with small-scale rotary-wing unmanned
aircraft systems. J. Am. Helicopter Soc. 2017, 65 , 1–17. [CrossRef]
48. Farassat, F.; Succi, G. A review of propeller discrete frequency noise prediction technology with emphasis on two current methods
for time domain calculations. Top. Catal. 1980, 71 , 399–419. [CrossRef]
49. Intravartolo, N.; Sorrells, T.; Ashkharian, N.; Kim, R. Attenuation of Vortex Noise Generated by UAV Propellers at Low Reynolds
Numbers. In Proceedings of the 55th AIAA Aerospace Sciences Meeting, Grapevine, TX, USA, 9–13 January 2017. [CrossRef]
50. Williams, J.E.F.; Hawkings, D.L. Sound Generation by Turbulence and Surfaces in Arbritary Motion. Philos Trans. Royal Soc. A
1969, 264, 321–342. [CrossRef]
51. Succi, G. Design of Quiet Efficient Propellers; SAE Technical Paper; SAE International: Warrendale, PA, USA, 1979; p. 14. [CrossRef]
52. Schlinker, R.; Amiet, R. Helicopter Rotor Trailing Edge Noise. In Proceedings of the 7th Aeroacoustics Conference, Palo Alto, CA,
USA, 5–7 October 1981; p. 2001. [CrossRef]
53. Rozenberg, Y.; Roger, M.; Moreau, S. Fan Blade Trailing-Edge Noise Prediction Using RANS Simulations. J. Acoust. Soc. Am.
2008, 123, 5207–5212. [CrossRef]
54. Corcos, G. The structure of the turbulent pressure field in boundary-layer flows. J. Fluid Mech. 1964, 18, 353–378. [CrossRef]
55. Pang, E.; Cambray, A.; Rezgui, D.; Azarpeyvand, M.; Showkat-Ali, S. Investigation Towards a Better Understanding of Noise
Generation from UAV Propellers. In Proceedings of the 2018 AIAA/CEAS Aeroacoustics Conference, Atlanta, GA, USA, 25–29
June 2018. [CrossRef]
56. Miljković, D. Methods for attenuation of unmanned aerial vehicle noise. In Proceedings of the 2018 41st International Convention
on Information and Communication Technology, Electronics and Microelectronics (MIPRO), Opatija, Croatia, 21–25 May 2018;
pp. 0914–0919.
57. Betz, A. Schraubenpropeller mit geringstem Energieverlust. Mit einem Zusatz von l. Prandtl. Nachrichten Von Der Ges. Der Wiss.
GöTtingen-Math.-Phys. Kl. 1919, 1919, 193–217.
58. Patrick, H.; Finn, R.; Stich, C. Two and Three-Bladed Propeller Design For the Reduction of Radiated Noise. In Proceedings of the
3rd AIAA/CEAS Aeroacoustics Conference, Atlanta, GA, USA, 12–15 May 1997; pp. 934–950. [CrossRef]
59. Roncz, J. Propeller Development for the Rutan Voyager; Technical Report; SAE Technical Paper; SAE International: Warrendale, PA,
USA, 2018. [CrossRef]
60. Adkins, C.; Liebeckt, R. Design of Optimum Propellers. J. Propul. Power 1994, 10, 676–682. [CrossRef]
61. Sobieszczanski-Sobieski, J.; Haftka, R. Multidisciplinary aerospace design optimization—Survey of recent developments.
Struct. Optim. 1996, 14, 1–23. [CrossRef]
62. Gur, O.; Rosen, A. Optimization of Propeller Based Propulsion System. J. Aircr. 2009, 46, 95–106. [CrossRef]
Fluids 2022, 7, 279 22 of 23
63. Gur, O.; Rosen, A. Multidisciplinary Design Optimization of a Quiet Propeller. In Proceedings of the 14th AIAA/CEAS
Aeroacoustics Conference (29th AIAA Aeroacoustics Conference), Vancouver, BC, Canada, 5–7 May 2008; Volume 3073, pp. 5–7.
[CrossRef]
64. Wisniewski, C.; Byerley, A.; Van Treuren, K.; Hays, A. Experimentally testing commercial and custom designed quadcopter
propeller static performance and noise generation. In Proceedings of the 23rd AIAA/CEAS Aeroacoustics Conference, Denver,
CO, USA, 5–9 June 2017; p. 3711.
65. Cattanei, A.; Ghio, R.; Bongiovı, A. Reduction of the tonal noise annoyance of axial flow fans by means of optimal blade spacing.
Appl. Acoust. 2007, 68, 1323–1345. [CrossRef]
66. Clark, I.; Daly, C.; Devenport, W.; Alexander, W.; Peake, N.; Jaworski, J.; Glegg, S. Bio-inspired canopies for the reduction of
roughness noise. J. Sound Vib. 2016, 385 , 33–54. [CrossRef]
67. Peake, N. The aeroacoustics of the Owl. In Fluid-Structure-Sound Interactions and Control; Springer: Berlin/Heidelberg,
Germany, 2016. [CrossRef]
68. Jaworski, J.; Peake, N. Aerodynamic noise from a poroelastic edge with implications for the silent flight of owls. J. Fluid Mech.
2013, 723 , 456–479. [CrossRef]
69. Chong, T.; Vathylakis, A. On the aeroacoustic and flow structures developed on a flat plate with a serrated sawtooth trailing edge.
J. Sound Vib. 2015, 345 , 65–90. [CrossRef]
70. Avallone, F.; Van Der Velden, W.; Ragni, D.; Casalino, D. Noise reduction mechanisms of sawtooth and combed-sawtooth
trailing-edge serrations. J. Fluid Mech. 2018, 848 , 560–591. [CrossRef]
71. Howe, M.S. Noise produced by a sawtooth trailing edge. J. Acoust. Soc. Am. 1991, 90, 482–487. [CrossRef]
72. Ning, Z.; Hu, H. An Experimental Study on the Aerodynamics and Aeroacoustic Characteristics of Small Propellers. In
Proceedings of the 54th AIAA Aerospace Sciences Meeting, San Diego, CA, USA, 4–8 January 2016. [CrossRef]
73. Hasheminejad, S.M.; Chong, T.; Joseph, P.; Lacagnina, G. Airfoil Self-Noise Reduction Using Fractal-Serrated Trailing Edge. In
Proceedings of the 2018 AIAA/CEAS Aeroacoustics Conference, Atlanta, GA, USA, 25–29 June 2018. [CrossRef]
74. Ragni, D.; Avallone, F.; van der Velden, W.C.; Casalino, D. Measurements of near-wall pressure fluctuations for trailing-edge
serrations and slits. Exp. Fluids 2019, 60, 6. [CrossRef]
75. Arce León, C.; Merino-Martínez, R.; Ragni, D.; Avallone, F.; Snellen, M. Boundary layer characterization and acoustic measure-
ments of flow-aligned trailing edge serrations. Exp. Fluids 2016, 57, 182. [CrossRef]
76. Wei, Y.; Xu, F.; Bian, S.; Kong, D. Noise reduction of UAV using biomimetic propellers with varied morphologies leading-edge
serration. J. Bionic Eng. 2020, 17, 767–779. [CrossRef]
77. Chaitanya, P.; Narayanan, S.; Joseph, P.; Kim, J. Leading edge serration geometries for significantly enhanced leading edge noise
reductions. In Proceedings of the 22nd AIAA/CEAS Aeroacoustics Conference, Lyon, France, 30 May–1 June 2016; pp. 1–20.
78. Chaitanya, P.; Joseph, P.; Narayanan, S.; Vanderwel, C.; Turner, J.; Kim, J.W.; Ganapathisubramani, B. Performance and mechanism
of sinusoidal leading edge serrations for the reduction of turbulence-aerofoil interaction noise. J. Fluid Mech. 2017, 818 , 435-464.
[CrossRef]
79. Narayanan, S.; Chaitanya, P.; Haeri, S.; Joseph, P.; Kim, J.; Polacsek, C. Airfoil noise reductions through leading edge serrations.
Phys. Fluids 2015, 27, 025109. [CrossRef]
80. Lyu, B.; Azarpeyvand, M. On the noise prediction for serrated leading edges. J. Fluid Mech. 2017, 826, 205–234. [CrossRef]
81. Hersh, A.; Soderman, P.; Hayden, R. Investigation of acoustic effects of leading-edge serrations on airfoils. J. Aircr. 1974,
11, 197–202. [CrossRef]
82. Leslie, A.; Wong, C.; Auld, D. Experimental analysis of the radiated noise from a small propeller. In Proceedings of the 20th
International Congress on Acoustics, ICA, Sydney, NSW, Australia, 23–27 August 2010.
83. McAlpine, A.; Nash, E.; Lowson, M. On the generation of discrete frequency tones by the flow around an aerofoil. J. Sound Vib.
1999, 222, 753–779. [CrossRef]
84. Graham, L. The Silent FLight of Owl. Aeronaut J 1934, 38, 837–843. [CrossRef]
85. Lee, S. Reduction of Blade-Vortex Interaction Noise Through Porous Leading Edge. AIAA J. 1994, 32, 480–488. [CrossRef]
86. Revell, J. Trailing-Edge Flap Noise Reduction by Porous Acoustic Treatment. In Proceedings of the 3rd AIAA/CEAS Aeroacoustic
Conference, Atlanta, GA, USA, 12–14 May 1997; pp. 493–505. [CrossRef]
87. Sueki, T.; Takaishi, T.; Ikeda, M.; Arai, N. Application of porous material to reduce aerodynamic sound from bluff bodies. Fluid
Dyn. Res. 2010, 42 , 015004. [CrossRef]
88. Geyer, T.; Sarradj, E.; Fritzsche, C. Porous airfoils : Noise reduction and boundary layer effects. Int. J. Aeroacoust. 2010, 9, 787–820.
[CrossRef]
89. Sarradj, E.; Geyer, T. Noise Generation by Porous Airfoils. In Proceedings of the 13th AIAA/CEAS Aeroacoustics Conference
(28th Aeroacoustic Conference), Rome, Italy, 21–23 May 2007. [CrossRef]
90. Jiang, C.; Moreau, D.; Yauwenas, Y.; Fischer, J.; Doolan, C.; Gao, J.; Jiang, W.; McKay, R.; Kingan, M. Control of rotor trailing edge
noise using porous additively manufactured blades. In Proceedings of the 2018 AIAA/CEAS Aeroacoustics Conference, Atlanta,
GA, USA, 25–29 June 2018. [CrossRef]
91. Moreau, S.; Dignou, B.; Jaiswal, P.; Yakhina, G.; Pasco, Y.; Sanjose, M.; Alstrom, B.; Atalla, N. Trailing-edge noise of a flat plate
with several liner-type porous appendices. In Proceedings of the 2018 AIAA/CEAS Aeroacoustics Conference, Atlanta, GA, USA,
25–29 June 2018. [CrossRef]
Fluids 2022, 7, 279 23 of 23
92. Rubio Carpio, A.; Merino Martínez, R.; Avallone, F.; Ragni, D.; Snellen, M.; van der Zwaag, S. Experimental characterization of
the turbulent boundary layer over a porous trailing edge for noise abatement. J. Sound Vib. 2019, 443, 537–558. [CrossRef]
93. Ali, S.S.; Azarpeyvand, M.; da Silva, C.I. Experimental Study of Porous Treatments for Aerodynamic and Aeroacoustic Purposes.
In Proceedings of the 23rd AIAA/CEAS Aeroacoustics Conference, Denver, CO, USA, 5–9 June 2017; p. 3358. [CrossRef]
94. Herr, M.; Rossignol, K.; Delfs, J.; Lippitz, N.; Mößner, M. Specification of Porous Materials for Low-Noise. In Proceedings of the
24th AIAA/CEAS Aeroacoustic Conference, Atlanta, GA, USA, 16–20 June 2014; pp. 1–19. [CrossRef]
95. Joseph, P.F. Poro-Serrated Trailing-Edge Devices for Airfoil Self-Noise. AIAA J. 2015, 53, 3379–3394. [CrossRef]
96. Chong, T.; Dubois, E.; Vathylakis, A. Aeroacoustic and flow assessments of the poro-serrated trailing edges. In Proceedings of the
22nd AIAA/CEAS Aeroacoustics Conference, Lyon, France, 30 May–1 June 2016. [CrossRef]
97. Chong, T.; Dubois, E. Optimization of the poro-serrated trailing edges for airfoil broadband noise reduction. J. Acoust. Soc. Am.
2016, 140, 1361–1373. [CrossRef] [PubMed]
98. Ghaffarivardavagh, R.; Nikolajczyk, J.; Anderson, S.; Zhang, X. Ultra-open acoustic metamaterial silencer based on Fano-like
interference. Phys. Rev. B Condens. Matter 2019, 99, 024302. [CrossRef]
99. Újsághy, O.; Kroha, J.; Szunyogh, L.; Zawadowski, A. Theory of the Fano resonance in the STM tunneling density of states due to
a single Kondo impurity. Phys. Rev. Lett. 2000, 85, 2557. [CrossRef]
100. Ning, Z.; Hu, H. An Experimental Study on the Aerodynamic and Aeroacoustic Performances of a Bio-Inspired UAV Propeller.
In Proceedings of the 35th AIAA Applied Aerodynamics Conference, Denver, CO, USA, 5–9 June 2017. [CrossRef]
101. Bodling, A.; Agrawal, B.; Sharma, A.; Clark, I.; Alexander, W.; Devenport, W. Numerical Investigations of Bio-Inspired Blade
Designs to Reduce Broadband Noise in Aircraft Engines and Wind Turbines. In Proceedings of the 55th AIAA Aerospace Sciences
Meeting, Grapevine, TX, USA, 9–13 January 2017; p. 0458. [CrossRef]
102. Noda, R.; Nakata, T.; Ikeda, T.; Chen, D.; Yoshinaga, Y.; Ishibashi, K.; Rao, C.; Liu, H. Development of bio-inspired low-noise
propeller for a drone. J. Robot. Mechatron. 2018, 30, 337–343. [CrossRef]
103. Huang, X.; Sheng, L.; Wang, Y. Propeller synchrophase angle optimization of turboprop-driven aircraft—An experimental
investigation. J. Eng. Gas Turbines Power 2014, 136, 112606. [CrossRef]
104. Jones, J.; Fuller, C. An Experimental Investigation of the Interior Noise Control Effects of Propeller Synchrophasing; Technical Report;
NASA: Washington, DC, USA, 1986.
105. Jones, J.D.; Fuller, C. Noise control characteristics of synchrophasing. II-Experimental investigation. AIAA J. 1986, 24, 1271–1276.
[CrossRef]