Sei sulla pagina 1di 22

CEAS Aeronautical Journal

https://doi.org/10.1007/s13272-019-00377-2

REVIEW PAPER

Aircraft noise generation and assessment

Combustion noise: modeling and prediction

C. K. W. Tam1   · F. Bake2 · L. S. Hultgren3 · T. Poinsot4

Received: 16 May 2017 / Revised: 12 November 2018 / Accepted: 11 February 2019


© This is a U.S. government work and its text is not subject to copyright protection in the United States; however, its text may be subject to foreign
copyright protection 2019

Abstract
This paper reviews both direct and indirect combustion noise. For convenience, they will simply be referred to as combus-
tion noise and entropy noise. Combustion noise has been studied for well over half a century. However, because of the large
number of parameters involved and the complexities inherent in the combustion processes, a widely accepted theory has yet
to be developed. For this reason, this review focuses primarily for direct combustion noise on experimental measurements,
semi-empirical relations and empirical but practical prediction methodologies. Important characteristic features and empiri-
cal correlations of combustion noise based on open flames and engine noise data are highlighted. Plausible generation of
entropy noise by the passage of entropy waves through a nonuniform mean flow was first predicted theoretically circa 1970s.
But it took forty years for its existence to be confirmed experimentally. Since then, there have been numerous publications
on this subject. They are the primary materials of this review. The fundamental experiment and noise generation mechanism
will be discussed first. Of great practical importance is whether there is significant generation of entropy waves inside an
engine. This issue and new methods for modeling and predicting internally generated engine entropy noise are items that are
examined at some length. Recent advances in computational methods especially in large eddy simulation lead us to envision
an important role to be played by high-fidelity numerical simulation in future combustion and entropy noise research and
prediction. A critical evaluation of a strategy for investigating and predicting the generation and propagation of these noise
components from the combustor of an engine to its exhaust is included in this review.

Keywords  Aeroacoustics · Propulsion noise · Combustion noise · Modeling · Prediction

This paper is part of a Special Issue on Aircraft Noise Generation


1 Introduction
and Assessment.
Core noise is projected to be an important component of
* C. K. W. Tam propulsive noise of the next generation aircraft (see Ref.
tam@math.fsu.edu
[1]). This is because traditionally dominant noise compo-
F. Bake nents such as jet noise and fan noise have been substantially
friedrich.bake@dlr.de
reduced in recent years by the introduction of high bypass
L. S. Hultgren ratio engines, by the use of geared fans and by incorporat-
Lennart.S.Hultgren@nasa.gov
ing innovative fan blade designs including sweep and lean
T. Poinsot technology and the use of advanced acoustics liners and
poinsot@cerfacs.fr
other noise suppressors. Another significant reason is that
1
Florida State University, Tallahassee, USA there is a strong push for green engines. Green engines are
2 designed for high efficiency and low emission. To achieve
German Aerospace Center (DLR), Berlin, Germany
3
these goals, these engines are to operate at a much higher
National Aeronautics and Space Administration (NASA),
temperature and at the fuel-lean limit. These changes, invari-
Cleveland, OH, USA
4
ably, will make the combustion processes highly unsteady
Institute de Mecanique des Fluides de Toulouse, Toulouse,
resulting in an increased emission of direct combustion noise
France

13
Vol.:(0123456789)
C. K. W. Tam et al.

and hot spots. When hot spots from the combustor are con- It is well known that combustion, generally, enhances
vected past a turbine or a flow constriction such as a nozzle turbulence. Should we include turbulence generated sound
in the flow path of the engine, sound will also be generated, as combustion noise? How significant is this noise com-
Refs. [2–4]. This is referred to as indirect combustion noise. ponent as compared to the noise associated with chemical
Thus, a green engine could potentially produce high levels reaction? These are difficult questions that do not seem to
of both direct and indirect combustion noise. have straightforward answers.
Over the years, there have been a number of compre- Combustion involves kinetic interaction at the molecular
hensive reviews on combustion noise. Among them are the level. So it is possible that some of the characteristics of
works of Mahan and Karchmer [5], Candel et al. [6], Duran combustion noise are controlled at the molecular level. For
et al. [7], Dowling and Mahmoudi [8], and Ihme [9]. The most of us, we implicitly assume that the dominant charac-
latest review article (by Ihme) was published in 2017. For teristics of combustion noise are formed at the continuum
this reason, we will include references from the extensive level and can be predicted using continuum equations. We
lists provided in these articles only when relevant to our are not sure if this assumption is correct.
discussion. In the literature, there are a number of illuminating theo-
Lately, because of renewed interest in the subject, there retical and experimental works on combustion noise, Bragg
has been a surge of publications in combustion noise (both [13], Smith and Kilham [14], Price et al. [10], Hurle et al.
direct and indirect). We will emphasize the new work, espe- [11], Thomas and Williams [15], Abugov and Obrezkov
cially modeling and prediction, observed characteristics of [16], Clavin and Siggia [17], Rajaram and Lieuwen [18],
combustion noise, and the physics and mechanisms of noise and Hirsh et al. [19], just to name a few. But one of the
generation. main difficulties is that kinetic time scales and continuum
This review is divided into three parts. Part 1 is on direct time scale are widely disparate. Moreover, hydrocarbon fuel
combustion noise. Part 2 is on indirect combustion noise. with complex molecules leads to multi-steps reactions. This
Part 3 is on high-fidelity simulation of combustion noise. makes any analytical calculation and even fully numerical
The reason behind these subdivisions is that the basic noise computation too involved to be feasible at this time. In addi-
generation mechanisms for direct and indirect combustion tion, in turbulent combustion, the effect of turbulence on
noise are quite different; although both are products of com- the transport of fuel and oxidizer is not so easy to quantify.
bustion. It is believed that by separating the two noise com- Because of all these reasons, there is no relatively well-
ponents would bring clarity and put each part of the review developed theory.
into focus. Furthermore, high-fidelity simulation of direct To avoid many of the difficulties mentioned above in the
and indirect combustion noise, either separately or simul- development of a theory of combustion noise, Strahle [20,
taneously, will play an increasingly important role in future 21] was the first to employ Lighthill’s Acoustic-Analogy
research and deserves a separate discussion. method to identify the combustion-noise source term by
recasting the governing equations in a certain form. Since
Strahle’s early work, this approach has gained a good deal
2 Part 1. Combustion noise (direct) of popularity. The Acoustic-Analogy formulation was origi-
nally developed by Lighthill [22, 23] in the early 1950s for
2.1 Introduction the identification of the sources of jet noise and as a frame-
work to describe its propagation to the far field. For turbu-
Direct combustion noise is generated when fuel is burnt. This lence-generated jet noise, Lighthill [22, 23] showed that the
suggests that to understand combustion noise, perhaps, we source in the Acoustic-Analogy theory is a quadropole. He
should first understand the processes of combustion. But this also established his celebrated U 8 velocity-scaling law for jet
is beyond the scope of this review. In addition, it is not clear noise by means of this method. However, during the inter-
we fully understand all the intricate details of the chemical vening years between then and now, high-quality narrow
processes in the combustion of hydrocarbon fuel at this time. band, hot and cold jet noise data at different Mach numbers
We recognize that there is still a lack of fundamental became available. As a result, more and more discrepancies
understanding on how direct combustion noise is generated. between the jet-noise quadrupole theory and experimental
Most investigators believe, with good reasons, that it has to measurements became known. About a decade ago, based
do with heat release during chemical reaction in the com- on extensive experimental data, Tam et al. [24] proposed a
bustion process (e.g. Price et al. [10], Hurle et al. [11], and two-source jet-noise model to replace the quadrupole rep-
Shivashankara et al. [12]). But it appears that no one has a resentation. They showed, using single far-field microphone
mechanistic explanation of how sound is generated because data, two far-field microphone cross-correlation data and
of the heat release. Presumably, sound is generated at the jet-turbulence far-field microphone correlation data, that
flame front. But the details are blurred. high-speed jet noise actually consists of two components.

13
Aircraft noise generation and assessment

These two noise components are generated by the fine scale


turbulence and the large turbulence structures of the jet flow.
They also provided far-field data that were at variance with
Lighthill’s velocity-scaling law. Their data show clearly that
the velocity exponent is not eight. It depends on the direc-
tion of radiation and the jet temperature. For a temperature
ratio 2.7 jet, the velocity exponent is around 6.5–7.5 in the
sideline direction and around nine for inlet angles larger than
130◦.
Another problem with the Acoustic-Analogy approach
is that there are several possible implementations. In the
combustion-noise investigation of Bailly et al. [25], the
Lighthill’s [22, 23], the Phillips’ [26] and the Lilley’s [27,
28] formulations were invoked, with each approach yield-
ing a different form of the combustion-noise source term.
So we have three different descriptions of the noise source.
This situation is rather perplexing and puzzling. It is unlike
many other physical problems for which the sources are well
defined. All these cause us to be unsure whether Acoustic Fig. 1  Spectral shape of combustion noise from open flames meas-
Analogy is an appropriate way to determine the sources of ured by Rajaram et  al.  [30]. Spectra are scaled by peak SPL and
laminar flame speed. Black dotted curve is the similarity spectrum,
combustion noise. Ref. [24]
In view of the many uncertainties in combustion-noise
theories, we believe it will be more beneficial to confine Part
1 of this article to review only experiments and empirical
and semi-empirical studies devoted to modeling and predic- that higher turbulence levels in flames give rise to higher
tion of combustion noise. combustion noise intensity but no effect on spectral shape.
In the studies of Kotake and Takamoto [29, 34], they pro-
2.2 Basic and simple engine noise experiments duced data to show that the shape of the noise spectrum and
frequency at the peak of the spectrum appeared not to be
Fuel-rich combustion is uneconomical and, therefore, dependent on burner size and shape and nor on the equiva-
impractical. For this reason, we will consider only fuel-lean lence ratio of the fuel. In a series of papers on the results
combustion noise. In addition, since most engines use hydro- of their experimental investigations on noise from open
carbon fuel, we will restrict our review to hydrocarbon fuel flames, Lieuwen and Rajaram [18, 39, 40], based on their
only. However, both liquid and gaseous fuels are included. own data, reported that the spectral shape of combustion
noise is independent of burner diameter, fuel equivalence
2.2.1 Combustion noise from open flames ratio, flow velocity and turbulence level. Figure  1 shows a
collapse of several of their noise spectra for a fixed-diameter
Open flames are the simplest source of combustion noise. burner into a single spectrum when the sound-pressure-level
It has been widely studied in the past e.g. Refs.  [5, 18, (SPL) is normalized by its peak value and the frequency F is
29–40] and others. Interests in the noise field from an open scaled by the laminar flame speed SL . An excellent summary
flame or an engine are generally confined to two quanti- of the various studies of combustion noise from open flames
ties, namely, the directivity and the spectra. It is known for is provided in the recent work of Rajaram and Lieuwen [45].
a long time that the noise from open flames is nearly iso- A major difficulty in studying combustion noise from
tropic e.g. Refs. [41–43]. Thus most studies on combustion open flames is that the phenomenon involves many vari-
noise from open flames concentrated on the noise spectrum. ables/parameters. They include parameters of the flame
Intensity is important but it can be computed by integrating holder geometry, the fluid flow parameters such as inflow
the spectrum. Variations of the noise spectrum with burner velocity, fuel and oxidizer densities, their temperatures and
diameter, turbulence level, flame temperature are reported in turbulence level, the chemical parameters of equivalence
Refs. [5, 18, 29–40]. Kilham and Kirmani [33], Putnam [44], ratio and whether the fuel is gaseous or liquid, premixed or
and Rajaram and Lieuwen [18] concentrated their research not. With so many parameters, it is not a simple task to cor-
effort on the effect of turbulence on premixed flame noise. relate them with the emitted noise to form simple relations
They used a grid to produce flow turbulence. They found experimentally.

13
C. K. W. Tam et al.

Fig. 2  Noise spectra of Honey-


well APU RE220 at full power
at three angles. Spectra dis-
placed vertically by 10 dB for
clarity. Smooth curves are the
similarity spectrum, Ref. [24]

2.2.2 Combustion noise from simple engines formula relating the noise intensity, I, and the fuel consump-
tion rate, ṁ f  , can easily be derived,
Large turbofan engines, invariably, have many noise sources.
Many of them have intensities higher than combustion noise. I ∝ ṁ 2f ∕r2 , (1)
Combustion noise is more likely to be observed in the meas-
urements using smaller turboshaft engines. where r is the distance to the observer. The proportional-
Auxiliary power units (APU) are small turboshaft engines ity constant depends on the engine size and design. So far,
used to provide power to commercial aircraft when parked this simple formula has been validated by Honeywell data,
on a ramp before the main engine is turned on. Their jet Ref. [46]. More validations are needed. If it is true, it con-
exhaust velocity is around Mach 0.1 or smaller. So jet noise firms the proposition that noise intensity generated by com-
is exceedingly low. The dominant noise components of bustion is largely proportional to the rate of heat release.
APUs are combustion noise and turbine noise. Turbine noise In the open literature, engine combustion noise other than
is primarily tones. Therefore, they can readily be identified those of APUs is rarely reported. Most engine companies
and separated out. The dominant broadband noise is com- treat their noise data as proprietary. There is no question that
bustion noise. the proprietary issue has hampered the advances of the sci-
Over the years, Honeywell Aerospace Company has ence of combustion noise and the development of prediction
measured extensively the noise of their large, medium and model and formulas.
small size APUs at high, medium and low power settings,
Tam et al. [46]. In the Honeywell APU experiments, the 2.2.3 The shape of combustion noise spectrum
direction of jet exhaust from the tail pipe is the zero degree
angle. The directivity covers the angular sectors of ±85◦ . Because the phenomenon of combustion involves many vari-
The data set has accumulated to more than 100 spectra. All ables and processes, theoretical advances in understanding
measured spectra show a dominant broad peak, as shown and predicting combustion noise are difficult. Historically,
in Fig. 2. There are some minor peaks superimposed on for physical problems of such complexity an empirical
the main spectrum. These peaks are known to be tail-pipe approach often becomes a good first step. Since the 1950s,
resonances. Tam et al. [46] after examining the entire data significant progress has been made by many investigators
set found that all the spectra had the same spectral shape in their studies of combustion noise from open flames and
regardless of the size of engine, power setting and direction small turboshaft engines. However, combustion-noise exper-
of radiation. The spectral shape they found was identical to iments are not easy to perform and the range of variables
the large turbulence structures similarity spectrum of high- that may be included in the tests is limited. Nevertheless,
speed jet noise, Ref. [47]. The similarity spectrum is shown one common finding reported in the literature is that the
as the smooth spectrum in Figs. 1 and 2. spectral shape of combustion noise is relatively insensitive
Fuel efficiency for APU is nearly 99%. On assuming that to changes in experimental variables. Tam [48] examined
the acoustic pressure fluctuations produced by combus- the largest, to date, set of existing observations and proposed
tion are proportional to the heat release rate, the following that the combustion-noise spectrum is seemingly universal

13
Aircraft noise generation and assessment

and is described by the similarity noise spectrum of the large opposite azimuthal mode numbers, but identical radial mode
turbulence structures of jet flows (see Figs. 1 and 2). He numbers, can be taken to be of the same random amplitude.
supported the proposition by showing good comparisons The plane wave mode (0, 0) can always propagate, but the
between the similarity spectrum and numerous combustion other modes can only propagate if the frequency is higher
noise spectra from open flames, non-premixed flames at than a mode-dependent cut-off/cut-on value. If the frequency
low Mach number, can-type combustors, APUs and turbo- is less than this value, the mode is evanescent.
fan engines at low power. A universal spectral shape, upon Karchmer [51] and, more recently, Royalty and Schus-
verification by further experimentation, would be very use- ter  [52] have documented the modal structure of the
ful for identification of combustion noise in environments, unsteady pressure field in aircraft-engine non-premixed
where there are more than one noise source and for provid- combustors. The former [51] utilized a ducted full-scale
ing a validation test for potential combustion-noise theories. YF102 annular combustor in a rig experiment, whereas the
latter [52] analyzed measurements obtained in an actual
2.3 Modeling and prediction TECH977 research turbofan engine. These investigations
both showed that the acoustic pressure spectrum at the exit
Efficient low-order noise modeling and prediction techniques of a real combustor has a multi-peak nature/structure and
are essential for system-level community-noise projections that individual modes could be uniquely identified within
and engineering trades at the preliminary/conceptual design separate frequency bands. The plane-wave (0, 0) mode domi-
stages of aircraft-propulsion systems. In general, the semi- nates for the lowest frequencies up to the cut-on frequency
empirical methods for turbofan core/combustion-noise pre- of the (±1, 0) mode pair. For higher frequencies, the most
diction in use today have their roots in developments that recently cut-on mode, or mode pair, dominates the unsteady
mainly took place during the 1970s. These early advances in pressure field; that is, successively higher azimuthal modes
core/combustion-noise modeling are discussed in the review dominate with increasing frequency. Radial modes, n ≠ 0 ,
chapter by Mahan and Karchmer [5], and references therein. in general, do not appear to play a role due to their cut-off/on
The approaches used can broadly be divided into fundamen- frequency values being higher than the direct-combustion-
tal and applied research avenues. noise frequency range. Similar results were obtained by Kre-
From a fundamental-research perspective, it was com- jsa and Karchmer [53] for the core-nozzle unsteady pressure
monly postulated that the noise generated in an aircraft com- field of an AiResearch QCGAT turbojet engine. They found
bustor is closely related to that of an open flame. Theories that the plane-wave (0, 0) mode was dominant up to about
and models for open-flame noise were developed and vali- 800–900 Hz at the tail-pipe exit.
dated by correlating observations with variations in physical Karchmer [51] concluded that the basic source generat-
parameters. Successively more complex situations were then ing mechanism itself has a relatively smooth and feature-
considered such as the effects of ducting on flame noise and less spectral shape, but the geometry of the combustor is
finally combustor-rig experiments. References and a further extremely effective in promoting resonant modes and thus
discussion can be found in the comprehensive review article modifying the unsteady pressure spectrum in the combustor.
by Strahle [49] and in Hultgren et al. [50]. It was found that This conclusion was echoed by Schuster and Lieber [54],
the radiated sound from an open turbulent flame has a rela- who also suggested that the far-field spectrum can be
tively universal spectrum. On an 1/3-octave basis, it gradu- thought of as the product of a single-peak broadband com-
ally rises to a peak somewhere in the 300–600 Hz frequency bustion noise spectrum, that is related to the spectrum of
range and monotonically falls off thereafter. an open turbulent flame, and a spectral transfer function
In principle, to model the direct combustion noise ema- that represents the resonance and propagation effects in
nating from an aircraft engine, an understanding is needed the engine core and exhaust. They [54] observed that it is
of the acoustic pressure field inside a combustor, how it is unrealistic to expect that a single transfer relation will be
affected by changes in engine-operational parameters, and generally applicable to the wide variety of combustor, tur-
finally its further propagation and interaction with turbine bine and exhaust geometries. However, as pointed out by
stages through the engine-internal flow path. Inside actual Mahan and Karchmer [5], the higher modes are often cut-
combustors, operating either in rigs or engines, the confined off in the smaller diameter turbine and core nozzle and are
geometry leads to the existence of a multitude of acous- thus unable to propagate efficiently through the engine to
tic modes, (m, n), where m and n denote the azimuthal and the surroundings. As a consequence, it is generally accepted
radial mode numbers, respectively. These modes are driven that the direct combustion noise emanating from turbofan
by the unsteady heat release of the combustion and their engines is dominated by the plane-wave mode and, there-
amplitudes are generally statistically independent. How- fore, effectively has a single-peak spectrum. That is, spectral
ever, for the common situation, where there is no signifi- peaks associated with the non-plane-wave modes are in prac-
cant global swirl present in the combustor, then modes with tice reduced to such an extent that their amplitudes are well

13
C. K. W. Tam et al.

below the dominant plane-wave peak as well as the levels of In 1980, The Society of Automotive Engineers Interna-
other propulsion-noise components. For current generation tional (SAE) adopted the method developed by GE [59] as
engines, jet noise, in particular, dominates even the peak the SAE ARP876 [61] technical standard for the prediction
value of the direct combustion noise except at low engine- of noise from conventional combustors installed in gas-tur-
power settings. The situation can be different for auxiliary bine engines. This standard was the culmination of a roughly
power units (APUs) due to their low exit velocities and often ten-year effort with contributions from major American and
long tailpipes, however. The low exit velocities significantly European companies. The standard [61, Appendix D] also
reduce the importance of jet noise and makes combustion contains a background history of its development as well as
noise the dominant low-frequency exhaust noise source for an extensive list of relevant references. This semi-empirical
APUs. Schuster and Lieber [54] also found that APU far- direct-combustion-noise model [59, 61] still forms the ker-
field combustor-noise peak frequencies correlated strongly nel of the core-noise module in the NASA Aircraft Noise
with the exhaust duct length and mean exhaust duct sound Prediction Program (ANOPP) [62, 63] and it is referred
speed according to plane-wave radiation from an unflanged to therein as the SAE method. The NASA-ANOPP core-
circular pipe. noise module also contains several extensions to the GE/
A substantial amount of applied research, relating meas- SAE formulation. The GE/SAE, the PW, and the NASA-
ured real-engine noise levels to operating parameters, was ANOPP methods are further discussed in subsections below.
published in the open literature during the 1970s. This work, Tam [48] provides a further discussion of the spectral dis-
eg. [55–59], was mainly carried out by aircraft-engine com- tributions used in the methods as well as a suggestion for an
pany researchers in the United States, often with support improved universal far-field spectrum for direct combustion
from the Department of Transportation/Federal Aviation noise, see the discussion in Sect.  1.2.3 above.
Administration (DOT/FAA) or the National Aeronautics
and Space Administration (NASA). Largely informed by the 2.3.1 A common form of the semi‑empirical models
physical understanding obtained from the early fundamental
studies briefly described above, the General Electric Com- A common feature of most of the semi-empirical models
pany (GE) [59] and Pratt and Whitney (PW) [58] determined for direct combustion noise is that the far-field directivity
semi-empirical, engine-manufacturer-specific formulas for and spectral distribution are decoupled. In this case, the
the total radiated acoustic power, with model coefficients/ (dimensional) combustion-noise mean-square pressure in
constants determined using rig testing of isolated compo- each 1/3-octave band (b), in the absence of atmospheric
nents and full-engine static-test data. These models [58, 59] attenuation, is given by
also included simple frequency-independent formulas to
account for the turbine attenuation of the broadband noise 𝜌∞ c∞ 𝛱D(𝜃)S(fb )
< p�2 >(b) = , (2)
originating in the combustor. The decoupled far-field direc- 4𝜋rs2
tivity and spectral distribution were obtained empirically
from full-engine static tests. for a static-engine test, where rs is the distance between the
In general, each of these acoustic power prediction source and the observer and 𝜌∞ and c∞ are the ambient den-
tools [58, 59] showed good agreement with data from the sity and speed of sound at the observer location. 𝛱 is the
engine manufacturer that developed the method, but not nec- total radiated acoustic power by the source. D(𝜃) is a direc-
essarily with data from other companies [5, 60]. The GE tivity function that depends only on the polar angle 𝜃 and
predictions [59] were within 3–5 dB of their data and the satisfies the normalization condition
PW measurements [58] had a standard deviation of 1.7 dB 𝜋

∫0
from the predictions. Zuckermann [60] suggested that the D(𝜃) sin 𝜃d𝜃 = 2. (3)
need for distinct models may be caused by differences in
burner design philosophy. A common difficulty encoun- S(fb ) is a spectrum function satisfying
tered during the development of these methods was that the ∑
measured total far-field acoustic signature in static-engine S(fb ) = 1 (4)
tests normally had to be adjusted by subtracting the low- b

frequency jet noise using an appropriate model to reveal the and fb is the 1/3-octave-band center frequency. Note that
core noise. This is because in the absence of forward-flight the total radiated acoustic power at the distance rs from the
effects, combustor noise is not a dominant noise source even source
at low engine-power settings. This makes the quality of the
results somewhat dependent on the accuracy, particularly at ∑
< p�2 >(b)
∫A
(5)
b
low frequencies, of the jet-noise model used. 𝜌c
dA = 𝛱 ,

13
Aircraft noise generation and assessment

where dA = rs2 sin 𝜃d𝜃d𝜙 , with 𝜙 denoting the azimuthal takeoff can be used. Note that the acoustic transmission loss
angle. That is, in the absence of atmospheric attenuation, the is independent of both the frequency and the engine operat-
total radiated acoustic power 𝛱 is preserved as the acoustic ing condition. The spectrum and directivity functions are
waves propagate towards the observer. defined in SAE ARP876 [61]. In particular, the spectrum
The sound pressure level SPL(b) in an 1/3-octave fre- is the SAE single-peak in-flight jet-noise spectrum with
quency band, the overall sound pressure level OASPL, and the peak frequency fixed at 400 Hz. Equation (10), with
the overall power level OAPWL are given by K = KSAE , combined with (11) for the turbine-transmission
loss, will be referred to as the SAE formulation herein. The
SPL(b) =10 log10 (< p�2 >(b) ∕p2ref ), (6) SAE ARP876 standard [61, Appendix D] also contains a

OASPL =10 log10 ( b < p�2 >(b) ∕p2ref ) = 10 log10 [𝜌∞ c∞ 𝛱D(𝜃)∕4𝜋rs2 p2ref ], (7)

background history of its development as well as an exten-


OAPWL =10 log10 (𝛱∕𝛱ref ), (8) sive list of relevant references.
where pref = 2 × 10−5 Pa and 𝛱ref = 1 × 10−12 W if SI units
are used. Note that some authors use PWL to denote the 2.3.3 The Pratt and Whitney model
overall power level and also simply refer to it as the power
level. Guided by [61], here Also during the latter half of the 1970s, researchers at Pratt
PWL (b)
= OAPWL + 10 log10 S(fb ), (9) and Whitney [57, 58] developed a semi-empirical prediction
method for direct combustor noise. They derived models for
denotes the acoustic power radiated in all directions associ-
the total acoustic power level, turbine coupling/ transmission
ated with a given 1/3-octave frequency band (b), however,
losses, and peak frequency; and they empirically determined
and is referred to as the power-level spectrum.
model constants, the directivity pattern, and a universal nor-
malized spectral distribution using a range of burner-rig and
2.3.2 The GE/SAE model
full-scale static engine tests.
In this model, the total radiated acoustic power is given
In the GE/SAE direct-combustion-noise model [59, 61], the
by [5, 58]
total radiated acoustic power is given by
� √ �4 � �
⎡ ṁ c Tt,ci h F 2 ⎤
1
OAPWL = 10 log10 ⎢ Ac Pt,ci
2 2
1 + PR st Fc2 ⎥ + 132 + 10 log10 (FTA ), (12)
⎢ Nf Pt,ci Ac cp Tt,ci ⎥
⎣ ⎦

where Nf is the number of fuel nozzles in the combustor, Ac


( )2 ( )2 is the combustor cross-sectional area, hPR is the constant-
Tt,ce − Tt,ci Pt,ci
𝛱 = 10K∕10 c2o ṁ c × FTA , (10) pressure fuel heating value, cp is the specific heat at constant
Tt,ci Po pressure, Fc and Fst are the combustor fuel–air ratio and the
stoichiometric fuel–air ratio; and as earlier ṁ c , Pt,ci , and Tt,ci
where FTA is a turbine attenuation, or loss, factor and is given are the mass flow rate and total pressure and temperature at
by [56, 61]except that the acoustic power the combustor inlet. The frequency-independent, turbine-
( )−4 attenuation factor is given by
𝛥Tdes
FTA = FTA,GE = . (11) 4𝜁(𝓁∕𝜋D) 0.8𝜁
To FTA = FTA,PW = = , (13)
(1 + 𝜁)2 (1 + 𝜁)2
The constant K = KSAE = − 60.53 … [61]. ṁ c is the mass √
where 𝜁 = 𝜌te cte ∕𝜌ti cti ≈ (Pt,te ∕Pt,ti ) Tt,ti ∕Tt,te  , with the sub-
flow rate into the combustor, Tt,ci and Tt,ce are the total tem- scripts ’te’ and ’ti’ indicating turbine exit and inlet, is the
perature at the combustor inlet and exit, Pt,ci is the total ratio of the characteristic impedances across the turbine. 𝓁 is
combustor-inlet pressure, and Po and To are the reference the circumferential correlation length of the direct-combus-
(static) pressure and temperature. The reference state, tion noise at the combustor–turbine interface, D is the outer
denoted by the subscript ’o’, is the standard sea-level condi- diameter of this interface. Mathews and Rekos [58] found
tions ( Po = 101.325 kPa, To = 288.15 K, co = 340.294  m/s, that combustor-rig and transmission-loss-corrected engine
in SI units). 𝛥Tdes is the design-point temperature drop across data agreed when 𝓁𝜋∕D = 0.2 , which led to the final, and
the turbine. If this is not known, the temperature drop at

13
C. K. W. Tam et al.

simplified, result in (13). Note that the acoustic transmission Hultgren [65] used time-series data, obtained during the
loss depends on the engine operating condition in the PW NASA/Honeywell Engine Validation of Noise and Emis-
formulation. sion Reduction Technology program [66] (EVNERT), to
Unfortunately, the argument of the base-ten logarithm, assess the turbine transfer of direct combustion-noise and
i.e., the quantity inside the square brackets, in (12) is not to develop an update to the turbine-attenuation formula used
dimensionally correct. It should be nondimensional, but it in ANOPP. The program used the Honeywell TECH977
is not! This means that the constant ( = 132 , here) depends research turbofan engine, which is typical of a business-jet
on the choice of units for the independent variables. Conse- application in the 31–36 kN (7000–8000 lbf) thrust class.
quently, to use this formula (12), customary US engineering The true combustion-noise turbine-transfer function was
units must be used in the input, see, [58] for unit details. educed from the EVNERT data by applying a three-signal
The peak frequency involves burner design and geometry technique utilizing one sensor in the combustor and two
parameters and is given by at the turbine exit. Note that the true turbine gain factor is
( ) always underpredicted (i.e., attenuation is overestimated) by
Rh ṁ f 1 a directly measured one, using only two sensors because
fp = Kf PR , (14)
cp Pt,ci AL
des c c
of a positive bias error caused by the presence of pressure
fluctuations in the combustor that are uncorrelated with the
where R is the gas constant, ṁ f is the fuel mass flow rate, propagated direct-combustion noise [65]. The resulting gain
Lc is the combustor length, and the subscript ‘ des ’ indicates factors ( < 1 ) were compared with the corresponding con-
evaluation at the design point (which can be taken as takeoff stant values obtained from the GE and the simplified PW
conditions). acoustic-turbine-loss formulas, (11) and (13). Hultgren [65]1
found that the gain factor obtained from the simplified PW
2.3.4 The NASA–ANOPP combustion‑noise models formula (13) agreed better with the experimental results
for frequencies of practical importance. He also found that
The purpose of the NASA-developed modular computer replacing the GE combustor-noise turbine-attenuation func-
program ANOPP is “to predict aircraft noise with the best tion (11) in Hultgren and Miles [67] with the simplified PW
currently available methods” [64]. To maintain and enhance one (13) clearly improved the total-noise predictions and
the program, NASA has continued to contract with industry also improved the combustion-noise predictions. The lat-
to evaluate ANOPP against fly-over and real-engine data and ter comparison was not as conclusive as the former due to
to suggest improvements to its modules. These investments the inherent difficulty in extracting the combustion-noise
have over the years led to significant improvements in the component from the total noise signature over its whole
capability of ANOPP. frequency range. However, the former would not be true if
The GE/SAE direct-combustion-noise model [59, 61] was the combustion-noise component predictions had not been
the first method implemented in the ANOPP [62, 63] core- improved by the attenuation-formula change.
noise module GECOR and is referred to as the SAE option The ANOPP core-noise module was subsequently
therein. However, in contrast to the original SAE formula- updated to allow the user a choice of using either the GE
tion, the reference state, denoted by the subscript ’o’ in (10) or PW turbine transmission formulas for both the SAE and
and (11), is in ANOPP taken to be the actual ambient condi- SmE options, i.e., the two additional configurations ANOPP-
tions, which generally are different from standard sea-level SAE-PW and ANOPP-SmE-PW were added. The current
values. Equations (10) and (11), with K = KSAE and relaxed ANOPP module also contains a separate intermediate-nar-
reference state, will be referred to as the ANOPP-SAE-GE row-band method [54] to account for tail-pipe resonances.
formulation herein.
The ANOPP prediction capability was extended to a 2.3.5 European semi‑empirical models for aircraft noise
larger class of engines during the mid 1990s by updating prediction
the propulsion-noise modules in ANOPP to also cover small
turbofan engines such as typically used by smaller regional- Since the early 2000s (in particular the last ten years), there
transport and business aircraft. For the GECOR module, has been an upswing in European aircraft noise research
this led to the introduction of a new small-engine (SmE) and prediction-tool development. Historically, most of the
option. The SmE-option formulas are identical to those development of non-proprietary models and semi-empirical
of the SAE option, except that the acoustic power level is tools for the prediction of aircraft noise took place in the US,
reduced by 4 dB, i.e., the constant K = KSmE = − 64.53 …
in (10). Equation (10), with K = KSmE , combined with (11)
for the turbine-transmission loss, will be referred to as the 1
  Due to a typographical error, the formula in [65] corresponding to
ANOPP-SmE-GE formulation herein. Eq. (13) is inverted, but the computations therein are correct.

13
Aircraft noise generation and assessment

but this is now changing. Several efforts are underway (or the open literature, if a core/combustion-noise model has
at a mature stage), some of which are of the database type, been implemented in CEASAR-S, at least it appears not nor-
using flyover data from existing aircraft, and some of which mally used in predictions. It is often stated that CEASAR-
are for physics-based semi-empirical models. Some of the S “computes noise source levels coming from main noise
latter efforts are briefly discussed below. Even though these components of current jet aircraft: jet, fan, high-lift devices
efforts are very impressive and overall advance the state-of- and landing gear” [73], see also [72].
the-art, it appears that any advancement of semi-empirical
models for direct-combustion noise has yet to take place. 2.3.6 Other codes

ANOTEC Consulting SOPRANO Tool. Based on the descrip- All major engine and airframe companies have proprietary
tion found in Ref. [68], the SOPRANO tool for aircraft noise in-house noise-prediction codes with capabilities similar to
calculations, developed by ANOTEC Consulting during the those of ANOPP. Their confidential, semi-empirical, direct-
European SILENCE(R) research program, contains several combustion-noise prediction methods are most likely based
airframe- and propulsion-noise (including core) components on the models described herein. Because each company has
implemented from models available in the open literature a large database of test results (including noise-certification
at that time. The core noise is predicted using the SAE tests), their models can be tailored for the engine type (for
ARP876 method.2 Since then SOPRANO has been extended an open example of this see [74]), and maybe even specific
with additional modules mainly within various European aircraft, under consideration.
research and development projects, but no other core-noise The companies more than likely also have multiblade-row
module has been included.2 actuator-disk-type codes (eg. [8, 74, 75]) to determine the
transfer of direct noise and the generation of indirect com-
DLR PANAM Prediction Tool. The German Aerospace bustion noise by the turbine. In general, this type of code
Center (DLR) developed the Parametric Aircraft Noise needs somewhat detailed mean-line data for each turbine
Analysis Module (PANAM) overall-aircraft noise predic- stage to solve a matrix problem for the transmission/genera-
tion tool [68–70] to enable comparative design studies with tion of noise. Consequently, this type of codes might not be
respect to airport-community noise and to identify promis- used directly in first-cut exploratory studies, but rather later
ing low-noise technologies at early aircraft-design stages. when noise estimates are refined or to tune parameters in
Each included airframe and propulsion noise-source is simu- simpler semi-empirical models.
lated using a semi-empirical model. The engine-noise mod-
els are based on existing models in the literature. However, 2.4 Future work
it appears [69, 70] that only the two dominant propulsion-
noise sources for current generation aircraft, namely jet and Because of the complexities and the numerous kinetic steps
fan sources, are currently implemented in the engine-noise involved in the burning of hydrocarbon fuel, it is unlikely a
module. complete theory of combustion noise, even in the simplest
combustion environment, would be developed in the near
ONERA IESTA Project. The French Aerospace Lab ONERA future. In the mean time, there is a real need to understand
(Office National d’Etudes et de Recherches Arospatiales) the relation between combustion and combustion noise
developed a generic simulation platform, IESTA (Infrastruc- generation. For example, combustion takes place along the
ture for Evaluating Air-Transport Systems), for the purpose flame front. Most likely it is also the place combustion noise
of assessing future air-transport concepts [71]. The predic- is generated. A mechanistic theory of how sound is produced
tion of aircraft noise is a significant part of determining the at the flame front would shed light on the source of combus-
environmental impact (noise and emissions) of these con- tion noise. Combustion processes are random and stochastic.
cepts. This capability has been implemented in an acous- So the noise generation processes would also have stochastic
tic model labeled CARMEN [72], which consists of three characteristics. Thus by allowing, for instance, the location
modules: CEASAR-S containing acoustic source models, of the flame front to be stochastic would make a more real-
CEASAR-I modeling installation effects (reflections and istic noise theory.
scattering by aircraft surfaces, etc.), and SIMOUN propa- Currently, there is a lack of reliable combustion noise
gating the sound field to observers on the ground. The CAR- prediction methods. It seems that instead of pursuing
MEN methodology is similar to the ones used by NASA purely theoretical development, one can take a phenom-
ANOPP [62, 63] and DLR PANAM [72]. It is not clear from enological approach to combustion noise. As an illustra-
tion of this approach, one may first concentrate attention
on a few parameters involved in combustion and later on
2
  N. van Oosten, private communication, April 2017. another set of parameters. The idea is to look for the effects

13
C. K. W. Tam et al.

of variation of the selected group of parameters on the radi- vorticity waves with fluctuations in velocity only and the
ated noise including spectra and directivity. One may then acoustic waves with fluctuations in all flow variables. Now,
try to extract formulas from the measured data relating noise if entropy waves are convected into a non-uniform flow, the
intensity, spectra and directivity to fuel burnt rate and other flow can no longer support entropy disturbances by them-
parameters. This approach, in addition to possibly providing selves. The disturbances in the non-uniform flow cannot con-
noise prediction formulas, would also allow the identifica- sist of temperature and density components alone. They will
tion of key parameters that would influence the generation now have pressure and velocity components too. In other
and radiation of combustion noise. Ultimately, combustion words, pure entropy blobs cannot exist in non-uniform flows.
noise radiated from an engine or combustor is important to This suggests that there is coupling of the three wave modes
the industry. By knowing the most influential parameter for leading to pressure fluctuations and hence noise radiation.
combustion noise generation, it would definitely be helpful The idea of mode coupling is, without doubt, correct.
in the design of low noise combustors. Nevertheless, it is somewhat of an abstract concept. To
explain the noise generation mechanism in more simple
physical terms Tam and Parrish [86] performed a numerical
3 Part 2. Entropy noise (indirect combustion experiment in which a narrow Gaussian entropy pulse is sent
noise) through a convergent–divergent (C–D) nozzle. The reason
for using a single narrow pulse is that in such an experiment
3.1 Introduction there will only be one possible noise source. Therefore, there
is no room for ambiguity as to where and how the sound is
Unsteady combustion processes in the combustor of an generated. Figure 3 shows the distribution of pressure and
engine produce direct combustion noise as well as entropy density fluctuations in a subsonic C–D nozzle at six differ-
(hot and cold spots) and vorticity waves. These waves are ent instances as the entropy pulse makes its way through the
convected downstream by the mean flow. Indirect combus- nozzle. In each figure, the entropy pulse is represented by
tion noise or entropy noise is generated when the hot and a dotted line, with corresponding density scale on the left
cold spots are strongly accelerated past a high-pressure tur- border, and the generated pressure wave is represented by
bine or through the non-uniform flow in the ducting and noz- a solid line, with corresponding pressure scale on the right
zle of the engine flow path. For a long period of time, indi- border. At station 1, the pulse is in the upstream uniform
rect combustion noise was merely a theoretical concept. Its duct, so the pressure is zero everywhere, since the waves
existence lacked experimental confirmation. In the follow- are uncoupled. At station 2, the pulse has just entered the
ing, we will first describe the fundamental experiment that convergent part of the nozzle, and the pressure just ahead of
proves the existence of indirect combustion noise/entropy the pulse drops to a negative value. This causes the entropy
noise by Bake et al. [4, 76–80]. The experiment used a spe- pulse to radiate a rarefaction wave into the downstream
cially designed test facility. We will then review the status region so as to maintain pressure balance at the pulse. At
of our understanding on indirect combustion noise for the station 3, the entropy pulse approaches the nozzle throat,
present generation engines. This is followed by a review on x = 0 . Here, the pressure just in front of the pulse becomes
the state of art in modeling and prediction of entropy noise. more negative. So a stronger rarefaction wave is radiated
We would like to note that just as entropy noise is gen- ahead. At station 4, the entropy pulse has passed through
erated when entropy waves are convected past a highly the nozzle throat into the divergent part of the nozzle. The
non-uniform flow, the convection of vorticity waves past pressure just ahead of the entropy pulse is now positive.
such a flow also generates noise. This was demonstrated Again, to maintain pressure balance, a compression wave is
by Kings and Bake [81] and Kings et al. [82] in a vorticity radiated ahead of the pulse. There is a switch from radiating
pulse experiment using a Vorticity Wave Generator test rig. rarefaction waves to compression waves when the entropy
Following this experiment, Kings et al. [83] repeated their pulse crosses the nozzle throat. At station 5, the radiated
investigation successfully using broadband small scale vorti- sound pressure remains positive or it is still a compression
city waves as input. Further validation of vorticity noise was wave. Finally, at station 6, the pulse reaches the downstream
provided by an experiment of Fischer et al. [84]. uniform region of the duct. As expected, there is no more
What is the mechanism responsible for the generation sound radiation.
of entropy noise? Since the beginning of entropy noise Notice that the half-width of the entropy pulse expands
research, the generation mechanism has, generally, been and then contracts as the pulse is convected through the con-
attributed to mode coupling. The idea is that a uniform flow vergent and then the divergent part of the nozzle. This pro-
can support three types of independent small amplitude dis- vides a clue on how sound is generated. Figure  4 illustrates
turbances, Chu and Kovasznay [85]. They are the entropy the sound generation processes in the convergent part of the
waves with fluctuations in temperature and density only, the nozzle. In this part of the nozzle, the mean flow velocity

13
Aircraft noise generation and assessment

Fig. 3  Instantaneous density and pressure distributions inside a C–D nozzle at six instances during the passage of an entropy pulse through the
nozzle. Dotted line represents density (left scale), solid line represents pressure (right scale). Nozzle throat is at x = 0

Fig. 4  Indirect combustion
noise generation processes in
the convergent part of a C–D
nozzle. Flow is from left to
right. Entropy pulse enters from
the left end of nozzle

13
C. K. W. Tam et al.

Fig. 5  Sketch of the Entropy


Wave Generator (EWG) test
facility

increases toward the throat (see the first inset below the noz- addition, due to the lack of data post-processing capabil-
zle in Fig. 4). Since the entropy wave pulse is convected ity at that time, clear identification of indirect combustion
downstream by the mean flow, this makes the front part of noise was not possible. Strahle and Muthukrishnan [89] in
the pulse move faster than the back part of the pulse. In turn, a combustion noise experiment attempted to correlate the
this difference in speed causes the pulse to expand in width measured total combustion noise power with their theory.
as indicated in the second inset of Fig. 4. The mass of the They employed a combustor test rig with an open outlet.
pulse is constant, so the density in the front part of the pulse With an open outlet, only weak velocity gradients were able
decreases. Now, the flow is adiabatic, because viscous dis- to develop inside the test facility. Because of the lack of
sipation and thermal conduction are negligible. It follows strong velocity gradient indirect combustion noise was not
that a drop in density will lead to a drop in pressure (see the observed. Muthukrishnan et al. [90] performed a coherence
third inset of Fig. 4). When the pressure in the front part of analysis involving different sensor signals inside and outside
the pulse decreases, pressure balance between the pulse and the combustion chamber. They reported that they were able
the gas in front requires the emission of a rarefaction wave. to separate the combustion noise components of their aero-
Once the entropy pulse crosses over to the divergent side engine combustor test rig. The result showed a dominat-
of the nozzle, the processes are reversed. In this part of the ing broadband entropy noise contribution in the total noise
nozzle, the width of the entropy pulse contracts, because the spectrum. Similar experiment was reported by Guédel and
front part of the pulse now moves slower than the back part Farrando [91] using a Turbomeca helicopter engine. They
of the pulse. As a result, the density in the front part of the used a three-signal coherence technique for noise separation.
pulse increases and a compression wave is emitted. This is They found a low-frequency sound source located between
shown in Fig. 3. the combustion chamber and the low-pressure turbine.
To provide further support for the above-proposed sound The first successful experiment proving the existence of
generation processes, the case of a similar pulse solution entropy noise was carried out in 2009 by Bake et al. [92,
but for a supersonic nozzle was also investigated. Unlike a 93]. A specially designed Entropy Wave Generator (EWG)
subsonic nozzle, the mean flow velocity of a supersonic noz- test facility was used in their experiment. The test setup is
zle increases downstream over the entire length of the C–D sketched in Fig. 5. It consists of a tube in which entropy
nozzle. Now, it was found that the entropy pulse radiated waves were produced by an electric heating module using
rarefaction waves in the do“wnstream direction throughout thin platinum wires. Downstream of the EWG module, the
its entire passage through the nozzle. This is completely con- flow passed through a C–D nozzle. Acoustic pressure sig-
sistent with the sound generation mechanism proposed in nals were measured by wall-mounted microphones in the
Fig. 4 and thus confirms its validity. tube section downstream of the nozzle. The test facility
was designed to allow a broad parametric study including a
3.2 Fundamental experiments variation of the amplitude of temperature fluctuations, the
C–D nozzle Mach number and the convection/propagation
The first experimental investigations aiming to confirm distance between the heating module and the C–D nozzle.
the existence of indirect combustion noise in the 1970s When the EWG is turned on for a short period of time (e.g.
were conducted by Zukoski and Auerbach [87] and Bohn 10 or 100 ms), an entropy wave is generated. This pulse is
[88]. They used a facility made up of a long channel con- convected downstream by the mean flow. If, indeed, indi-
nected to a nozzle. Temperature fluctuations were generated rect combustion noise is generated during the passage of the
by an electric heater. Unfortunately, the amplitude of the entropy pulse through the nozzle, the time of arrival of the
induced temperature was very small, approximately 1 K. In sound pulse can be measured accurately by the microphones

13
Aircraft noise generation and assessment

housed in the downstream section of the nozzle. By vary- of the combined direct and indirect combustion noise. In
ing systematically the distance between the heating element his work, a sequence of low pass filters was employed. The
and the C–D nozzle, it is possible to show unambiguously result showed that time delays measured in each direction is
that sound is produced inside the nozzle at the time when fairly consistent with one another within a range of operating
the entropy pulse is passing through the nozzle. In this way, conditions. He also demonstrated the feasibility of separat-
Bake et al. [92, 93], for the first time, succeeded in providing ing a mixture of direct and indirect combustion noise. The
direct and irrefutable evidence of the generation of indirect problem of noise-source separation was also investigated by
combustion noise. Hultgren and Miles [67].
Since this seminal demonstration of the generation of Bake et al. [110] studied the generation of entropy noise
indirect combustion noise, there has been a deluge of publi- and vorticity noise by either injecting entropy or vorticity
cations on this subject. Using the EWG test rig results, some fluctuations into the flow upstream of a high-pressure turbine
of these [94–98] give an enhanced model description for rig. Because the flow in such a rig is highly complex, this
entropy noise—basically succeeding the theory of Marble experiment is especially challenging. In addition, more than
and Candel [3]. The compactness and linearity assumption one type of noise sources are involved. In spite of the intrin-
have been examined [98, 99] yielding for example improved sic complexity of the problem, Bake et al. [110] were able
phase predictions. Furthermore, numerical simulations have to conclude that the dominant noise generated in the early
been conducted to assess the model capabilities [100] as well stages of the turbine is vorticity noise. However, noise gen-
as investigating [101–103] the two-dimensional extension erated in the later stages of the turbine was largely entropy
of the actuator-disk theory of Cumpsty and Marble [104] noise.
for turbomachinery blade rows. Other publications revealed In the framework of the European funded TEENI project
the significance of entropy noise for combustion instabil- (Turboshaft Engine Exhaust Noise Identification) coordi-
ity [105] even when accounting for entropy wave dispersion nated by Turbomeca, comprehensive acoustic measurements
and dissipation effects [99]. Tam et al. [106] demonstrated in and around a helicopter engine (Ardiden 1H-1 type) have
the contribution of entropy noise to the noise emission of been conducted, Pardowitz et al. [111]. The engine was
auxiliary power units (APUs). Finally Knobloch et al. [107] extensively equipped with acoustic sensors not only at the
provided additional test cases for entropy noise generation combustor but also at the turbine stages and at the exhaust.
in nozzle flows by presenting experimental results from the Due to the fact that jet noise is low for this engine, it was
Hot Acoustic Test Rig (HAT). anticipated that the core noise source contributions can be
quantified with better accuracy than for turbofan engines.
3.3 Engine noise experiments Classical coherence analysis methodologies [112–116] such
as the coherent-output-power technique, the signal enhance-
Demonstrating that indirect combustion noise can be gener- ment technique and the five-microphone method were used
ated in a specially designed experimental facility is impor- to study the source and propagation characteristics of the
tant by itself. However, it is also important to show that a various noise components. In addition, advanced analysis
significant amount of indirect combustion noise is actually methods, which combine correlation methods and mode
generated in an engine. One of the difficulties in doing so decomposition techniques [117, 118] were applied for the
is that an engine has many noise sources. So it is natural to first time to a real engine. The methods enabled the quan-
start by trying to separate the noise components inside and tification of the different noise source contributions to the
outside of an engine. total sound power. They also yielded a detailed database
Miles [108] used a coherence function technique, involv- for validation of numerical prediction methods. In addition,
ing a combustor-internal pressure sensor and far-field these methods allow an assessment of the contributions of
microphones, to perform noise separation to investigate the intensities of various noise sources in different frequency
the core noise sources of a two-shaft turbofan engine. The bands.
source location technique was based on adjusting the time The NASA funded study of Schuster et al. [119] con-
delay between a combustor pressure signal and a far field centrated on measuring several pieces of vital information
microphone signal by maximizing the coherence. It was relevant to indirect combustion noise generation inside an
discovered that the cross-spectral density band 0–200 Hz engine. Their testbed is a Honeywell TECH977 research
was dominated by indirect combustion noise generated by turbofan engine. Near the exit of the combustor, the tem-
entropy waves interacting with the turbine. The signal in perature is extremely high. This hostile environment makes
the 200–400 Hz frequency range was attributed mostly to any measurement particularly difficult. Schuster et al. [119]
direct combustion noise. Miles  [109] later extended his measured successfully the temperature and pressure fluctua-
work using a generalized cross-correlation function tech- tion spectra at a location near the combustor exit (upstream
nique to study the change in propagation time to the far field of the high-pressure turbine), in the inter-turbine duct, and

13
C. K. W. Tam et al.

Fig. 6  Measured temperature
fluctuation spectrum at the exit
of the combustor of a Honey-
well TECH977 engine, Schuster
et al. [119]. Orange color line is
the measured spectrum. Black
color line is the stochastic
model simulation spectrum (see
Sect. 3.4)

Fig. 7  Comparison of two
pressure spectra, one measured
upstream (black) and the other
downstream (green) of the high
pressure turbine. Origin of
noise increase is unknown at
this time. Data from Schuster
et al. [119]

13
Aircraft noise generation and assessment

downstream of the low turbine exit. Figure 6 shows their the need to first solve the combustor problem. Solving the
measured temperature fluctuation spectrum near the com- combustor problem is by no means an easy task. It is, in
bustor exit. This is one of the first, if not the very first, of this many ways, more difficult than computing the generation
type of measurements available in the open literature. The of indirect combustion noise during the passage of entropy
spectrum is dominated by relatively low frequency compo- waves through a high-pressure turbine. For this reason, most
nents. Figure  7 shows a superposition of two pressure spec- recently some investigators, Tam and Parrish [121] and
tra. These spectra were measured using specially designed Papadogiannis et al. [122] tried to follow an alternative strat-
semi-infinite-tube pressure sensors, with purge flow. The egy. Their idea is to use a simplified entropy wave field as
black spectrum is measured at a location upstream of the input to the nozzle or turbine flow computation. In both stud-
high-pressure turbine. The green spectrum is measured ies, a sinusoidal entropy wave train was used as input. This,
downstream of the high-pressure turbine. The significance of course, is an oversimplification of a real entropy wave
of Fig. 7 is that, by comparing the two spectra, it becomes field so their computation only yields qualitative results.
clear that there is an increase in noise across the high-pres- Earlier, Tam in his computational aeroacoustics
sure turbine. The tantalizing question is ‘what generates this book [123, Appendix F] presented an energy conserving
extra noise?’ There are a number of possibilities. But it is method for discretizing a given pressure spectrum and then
possible that it is indirect combustion noise. The challenge using the discretized information to create computationally
is how to show that it is! a stochastic pressure wave field in the time domain. This
method, when properly implemented, automatically ensures
3.4 Modeling and prediction that the stochastic wave field created numerically has the
same spectrum as the originally given pressure spectrum.
Prediction of indirect combustion noise began with the pio- Recently, this method was modified by Tam et al. [124] to
neering work of Marble and Candel [3] and Cumpsty and generate a stochastic entropy wave field inside a Honeywell
Marble [104, 120]. These early works have proven to be TECH977 engine. The spectrum in black shown in Fig. 6 is
very influential. They introduced the idea of the short noz- the spectrum computed by them using the modified Tam’s
zle approximation also known as the actuator-disk theory. method. The prescribed spectrum of their computed entropy
This approximation made it possible to make quantitative wave field was the Schuster et al. [119] spectrum shown in
prediction of the generation and transmission of indirect orange color in Fig. 6. It is obvious that there is excellent
combustion noise through nozzles and ducts. For this rea- agreement. The Tam method was originally formulated for
son, it is still used by some investigators today. However, the plane waves. Most recently, the method has been extended
short nozzle approximation is primarily a one-dimensional to produce a fully three-dimensional stochastic entropy-
approximation. It applies only to low frequency sound with wave field in the time domain. In the extended method, two
wavelength longer than the nozzle length. For sometime, two-point spatial cross-correlation functions in two orthogo-
the full Euler equations plus the energy equation have been nal directions in the lateral plane (plane perpendicular to
solved computationally, at least, for geometrically simple the flow direction) are additional input functions. These
laboratory experiments. The situation in real turbomachinery cross-correlation functions allow a user to specify the lat-
still presents a challenge. However, it is believed that for eral dimensions of the entropy blobs in the entropy wave
indirect combustion noise prediction, as well as for turbine field. In other words, if the extended method is used, a user
attenuation of direct combustion noise, an effort should be can create a random broadband entropy wave field with a
made to use computational methods. This will ultimately prescribed spectral shape and intensity in temperature fluc-
allow a high-accuracy prediction over a much larger range tuations and distributions of the lateral dimensions of the
of frequencies than the actuator-disk methods provide. entropy blobs through the specification of the lateral spatial
To predict indirect combustion noise from an engine, con- cross-correlation functions.
ventional thinking suggests that one should start with the Now, if the temperature fluctuation spectrum and the two-
prediction of the creation of hot and cold spots inside the point cross-correlation functions are measured at a location
combustor. Then the computation should follow the down- slightly downstream of the combustor exit, it is possible
stream convection of these entropy spots by the mean flow. through the use of the stochastic entropy wave field genera-
Upon exiting the combustor, the flow carrying the entropy tion method to convert the measured data into a time domain
disturbances enters and makes its way through the high- inflow boundary condition for the computation of indirect
pressure turbine. There is a general consensus that it is in combustion noise generation inside an engine or across a
the passage through the high-pressure turbine that signifi- high-pressure turbine. Such a boundary condition effectively
cant amount of indirect combustion noise is generated. If replaces the engine combustor. In this way, the need to solve
this prediction strategy is followed then the prediction of the combustor problem can be entirely circumvented. A sig-
indirect combustion noise from an engine is burdened with nificant benefit of this approach is also that the computed

13
C. K. W. Tam et al.

broadband noise is essentially indirect combustion noise. noise produced by current generation aircraft engines. This
There is no need to carry out a noise component separation does not mean that engines do not produce indirect combus-
operation. tion noise. This is, perhaps, because the intensity is rela-
This alternative strategy was employed by Tam tively low and hence it is difficult to identify indirect com-
et al. [124] in their attempt to show that the noise increase bustion noise from all the other noise components emitted
across the high-pressure turbine inside the Honeywell from an engine. In addition, we need better noise component
TECH977 engine (see, Fig. 7) is indirect combustion noise. separation methods to identify indirect combustion noise.
To avoid full modeling of the turbines, they replaced the
turbines by body forces with a magnitude such that the same 3.5 Future work
pressure drops across the turbines as measured are repro-
duced. There is no tunable parameter in their computation. Since the breakthrough experiment by Bake et. al. [4] dem-
The computed results yielded a spectrum that matched well onstrating that indirect combustion noise or entropy noise
(within 3 dB) with the green spectrum in Fig. 7 up to about was real, many in the aeroacoustics community started to
1 kHz frequency. Above 1 kHz, the computed spectrum recognize and become concerned that core noise could be
drops off rapidly. Tam et al. attributed the high-frequency the next noise barrier for commercial aviation. This worry
drop-off to the lack of details in their turbine model. There- is compounded by the prospect of replacing present day jet
fore, a confirmation of the generation of indirect combus- engines by green engines in the not too distant future. Cur-
tion noise in an engine is still lacking though some progress rently, an open question is how important entropy noise is
seems have been made. among the various noise components radiating from exist-
Many investigators use numerical simulations to study ing jet engines. There have been a number of experimental
and to predict indirect combustion noise generation inside an investigations on this issue. But, thus far, no definitive con-
engine. A popular way is to employ a readily available com- clusion is available. Research on entropy noise from present-
mercial CFD code that the investigator had used before. But day and future green engines is unquestionably a top priority.
one must be cautioned that this may not be a good practice. Just as important is to find possible ways to suppress this
The magnitude of sound inside an engine (except for reso- noise component. However, up to now there does not seem
nance) is, in general, many orders of magnitude smaller than to have been much work done on entropy noise suppression
that of the fluid flow. The numerical noise of a commercial either theoretically or experimentally. Needless to say, fresh
code may be tolerable for flow computation, but it could be ideas are very much needed.
comparable in level to that of indirect combustion noise. In
addition, most commercial CFD codes have strong built-in
numerical damping (because of numerical stability consider- 4 Part 3. High‑fidelity simulations (direct
ations). There is usually no simple way to assess the amount and indirect combustion noise)
of damping or a way to alter the damping rate. Such damping
could be strong enough to wipe out a big part of internally Predicting combustion noise is often done using semi-
generated acoustic waves. To compute indirect combustion empirical methods, based on extrapolations and simplifica-
noise in an engine, it is natural to take the computational tions. The recent revolution due to the new performances
domain to be the entire flow path inside the engine. This of Large Eddy Simulation methods allows now to imagine
choice, invariably, create inflow and outflow boundaries that that in future a fully deterministic simulation chain may be
require proper boundary conditions. The inflow boundary developed to predict combustion noise from first principle
condition has to play a dual role. That is, it must generate the computations only [9, 125, 126]. The challenge is clearly
incoming entropy waves and at the same time allow the out- a difficult one: it requires the simulation of the unsteady
going disturbances created inside the engine to exit without flow in combustion chambers and in turbine stages, followed
significant reflections. If an aeroacoustic code instead of a by a jet noise simulation. Moreover, all these computations
CFD code is employed for the time marching computation, must be coupled: the LES of the combustion chamber must
a split-variable method (see Tam [123, Chapter 9]) could capture acoustic (direct noise) and entropy as well vorticity
be used. At the outflow boundary, an absorbing boundary waves (indirect noise) and feed them to a tool computing
condition is needed to avoid contaminating the computed the propagation of these waves in the rotor/stator stages;
solution by wave reflected back by the outflow boundary finally these perturbations must be introduced in a simula-
back into the computational domain. Acoustic boundary tion of jet noise and propagated to the far field. A few groups
conditions of these types are usually not a part of a com- are trying to build such a chain using high-fidelity solvers.
mercial CFD code. ’High-fidelity’ approaches in this context must be under-
At this time, there is no clear experimental evidence that stood as strategies based on first principles and as little mod-
indirect combustion noise plays an important role in the eling as possible. This requires unsteady solvers which can

13
Aircraft noise generation and assessment

Fig. 8  Fully deterministic chain to compute combustion noise: LES is the turbine stages using actuator-disk theory to obtain both direct and
performed in the combustion chamber; acoustic, vorticity and entropy indirect noise in the outlet nozzle; a far field propagation tool is then
waves are measured at the combustor outlet plane and propagated into used to propagate noise to the far field

capture turbulent structures and numerical methods which flow within chambers [130–132]. The chemical reactions
offer almost zero dissipation and limited dispersion effects controlling methane/air or kerosene/air flames can now
to capture all waves propagating in the combustion cham- be introduced in 3-D codes using tabulation methods or
ber. The cost of such simulations is typically two orders of reduced chemistry approaches [133], while the flame/tur-
magnitude larger than conventional, commercial CFD codes: bulence interaction is handled using sophisticated multiscale
massively parallel computations using thousands of cores approaches zooming in the flame front to resolve its struc-
are mandatory. Note that when flames are unconfined, such ture [130]. These simulations capture the unsteady motions
computations are already possible [127, 128], because noise of the flame front, where heat release takes place, lead-
is produced directly by unsteady combustion and only direct ing to strong flow divergence and direct noise production.
noise is important: no entropy noise has to be considered, However, even though LES can be used for the combustion
because there is no turbine. Figure  8 displays an example chamber to predict all unsteady activity due to turbulence
of combustion noise prediction tool developed and applied and combustion with precision, the situation is different in
to helicopter and aircraft noise, Ref. [129]. the turbine, where it is much more difficult to use a high-
Using high-fidelity CFD methods for all components of fidelity method to predict noise propagation and generation
Fig. 8 (combustion chamber, turbine, jet) remains impossible by indirect mechanisms. In addition, the noise generated in
today. In the combustion chamber itself, LES has become the chamber and propagated through the turbine, must then
the reference tool in the last ten years: recent progress in be added to the computation of the jet noise before propa-
chemistry description and in turbulent subgrid scale models gating to the far field. A reasonable compromise (displayed
allows now to perform simulations of the unsteady reacting in Fig. 8) is to use LES in the combustion chamber and

Fig. 9  TEENI engine used to validate the computation chain of Fig. 8 [129]

13
C. K. W. Tam et al.

extract acoustic, vorticity and entropy waves at the outlet of


the chamber, upstream of the first turbine stages. An ana-
lytical theory based on actuator-disk assumptions, can then
be used to compute direct and indirect noise through the
turbine stage. This requires the knowledge of the flow cross
sections in the turbine and of the enthalpy jumps at each
rotor but this is much cheaper than a full 3-D LES of the
whole turbine. A limitation of actuator-disk theory is that the
Marble and Cumpsty approach [104] used to describe wave
propagation through a turbine stage is valid only for low
frequencies (compact limit), where jump conditions can be
written on both sides of the stage. Extending this theory to
higher frequencies remains a challenge: this has been done
already for simple nozzles, where the Marble and Candel
Fig. 10  Far-field microphones used for the TEENI engine [129] technique [3] valid too only under the compact assumption,

Fig. 11  Acoustic pressure from


140◦ to 180◦  in the far-field.
The squares correspond to the
LES results without entropy
filtering and the circles to LES
with filtering. Solid line: experi-
mental data [129]

13
Aircraft noise generation and assessment

has been extended to all one-dimensional waves by Durán 4. Bake, F., Kings, N., Fisher, A., Röhle, I.: Experimental inves-
and Moreau [134] using a Markus expansion. For a turbine tigation of the entropy noise mechanism in aero-engines. Int. J.
Aeroacoust. 8(1–2), 125–142 (2009a)
stage with rotation effects, this remains to be done. 5. Mahan, R.J., Karchmer, A.: Combustion and core noise. In:
The most complete example of the application of the Hubbard, H.H. (ed) Aeroacoustics of flight vehicles: theory and
tool described in Fig. 8 was given in 2016 [129] for a heli- practice, volume 1, chapter 9, pages 483–517. NASA Reference
copter engine, where such a prediction chain was used and Publication 1258, WRDC Technical Report 90-3052 (1991)
6. Candel, S., Durox, D., Ducruix, S., Birbaud, A.-L., Noiray, N.,
compared to measurements performed by Safran Helicopter Schuller, T.: Flame dynamics and combustion noise: progress
Engines within the European TEENI project, where sound and challenges. Int. J. Aeroacoust. 8(1–2), 1–56 (2009)
levels were measured both inside and outside a real engine. 7. Duran, I., Moreau, S., Nicoud, F., Livebardon, T., Bouty, E.,
Figure  9 displays the engine geometry and shows where Poinsot, T.: Combustion noise in modern aero-engines. Aer-
osp. Lab J. 7–05, 1–11 (2014). https​://doi.org/10.12762​/2014.
pressure probes were used to measure sound within the com- AL07-05
bustor and the turbine. In addition, sound was also measured 8. Dowling, A.P., Mahmoudi, Y.: Combustion noise. Proc. Com-
in the far field, as shown in Fig. 10. bust. Inst. 35, 65–100 (2015)
As soon as the objective is to compute the noise of a full 9. Ihme, M.: Combustion and engine-core noise. Ann. Rev. Fluid
Mech. 49, 277–310 (2017)
annular chamber, the question of azimuthal modes and of 10. Price, R.B., Hurle, I.R., Sugden, T.M.: Optical studies of the
the necessity of performing a full 360◦ LES appears. Here, generation of noise in turbulent flames. In: Twelfth Symposium
both single sectors and a full 360◦ LES were performed. (International) on Combustion: at the University of Poitiers,
Results showed a limited impact of the azimuthal modes and France, July, 1968, pages 1093–1102. Proceedings of the com-
bustion institute (1969)
a cancellation effect of noise between sectors—each burner 11. Hurle, I.R., Price, R.B., Sugden, T.M., Thomas, A.: Sound
generates noise which is not phased with the others so that emission from open turbulent premixed flames. Proc. R. Soc. A
the overall noise is not the simple addition of individual 303(1475), 409–427 (1968)
burner noise. A simple filtering method was developed to 12. Shivashankara, B.N., Strahle, W.C., Handley, J.C.: Evaluation of
combustion noise scaling laws by an optical technique. AIAA J.
account for this cancellation effect. This filtering effect is 13(5), 623–627 (1975a)
mainly applied to entropy waves, because indirect noise is 13. Bragg, S.: Combustion noise. J. Inst. Fuel 36(1), 12–16 (1963)
the dominating noise generation mechanisms in this engine. 14. Smith, T.J.B., Kilham, J.K.: Noise generation by open turbulent
This is a specific case and in general, this filtering effect will flames. J. Acoust. Soc. Am. 35(5), 715–724 (1963)
15. Thomas, A., Williams, G.T.: Flame noise: sound emission
have to be considered too for direct noise. Noise levels were from spark-ignited bubbles of combustible gas. Proc. R. Soc. A
compared within the turbine and in the far field showing 294(1439), 449–466 (1966)
that the overall simulation chain of Fig. 8 has the potential 16. Abugov, D.I., Obrezkov, O.I.: Acoustic noise in turbulent flames.
to predict combustion noise in real engines (Fig. 11), i.e., a Combust. Expos. Shock Waves 14(5), 606–612 (1978)
17. Clavin, P., Siggia, E.D.: Turbulent premixed flames and sound
method based on first principles only is a reasonable objec- generation. Combus. Sci. Technol. 78(1–3), 147–155 (1991)
tive. Note, however, that Fig. 8 also reveals that the model 18. Rajaram, R., Lieuwen, T.: Parametric studies of acoustic radia-
used for the damping of the entropy waves in the turbine tion from premixed flames. Combust. Sci. Technol. 175, 2269–
stages has a strong effect on the results—this is obviously 2298 (2003)
19. Hirsh, C., Wäsle, J., Winkler, A., Sattelmayer, T.: A spectral
a key point of this approach which can be improved only model for the sound pressure from turbulent premixed combus-
by performing full LES of the entropy wave propagation tion. Proc. Combust. Inst. 31(1), 1435–1441 (2007)
through the stages. 20. Strahle, W.C.: On combustion generated noise. J. Fluid Mech.
49, 399–414 (1971)
Acknowledgements  L. S. Hultgren was supported by the NASA 21. Strahle, W.C.: Some results in combustion generated noise. J.
Advanced Air Vehicles Program, Advanced Air Transport Technol- Sound Vib. 23(1), 113–125 (1972)
ogy Project, Aircraft Noise Reduction Subproject. 22. Lighthill, M.J.: On sound generated aerodynamically. I. General
theory. Proc. R. Soc. Lond. A 211(1107), 564–587 (1952). https​
://doi.org/10.1098/rspa.1952.0060
23. Lighthill, M.J.: On sound generated aerodynamically. II. Tur-
References bulence as a source of sound. Proc. R. Soc. Lond. A 222(1148),
1–32 (1954). https​://doi.org/10.1098/rspa.1954.0049
1. Hultgren, L.S.: A comparison of combustor-noise models. AIAA 24. Tam, C.K.W., Viswanathan, K., Ahuja, K.K., Panda, J.: The
Paper 2012-2087 (NASA/TM-2012-217671), 18th AIAA/CEAS sources of jet noise: experimental evidence. J. Fluid Mech. 615,
Aerocoustics Conference, Colorado Springs, Colorado (2012) 235–292 (2008)
2. Candel, S.M.: Analytical studies of some acoustics problems of 25. Bailly, C., Bogey, C., Candel, S.: Modelling of sound generation
jet e. PhD thesis, California Institute of Technology (Also DOT- by turbulent reacting flows. Int. J. Aeroacoust. 9(4–5), 461–490
TST-76-104 1976) (1972) (2010)
3. Marble, F.E., Candel, S.M.: Acoustic disturbances from gas non- 26. Phillips, O.M.: On the generation of sound by supersonic turbu-
uniformities convected through a nozzle. J. Sound Vib. 55(2), lent shear layers. J. Fluid Mech. 9(1), 1–28 (1960)
225–243 (1977) 27. Lilley, G.M.: The generation and radiation of supersonic jet
noise. Technical Report AFAPL-TR-72-53, Air Force Aero-
Propulsion Laboratory, (1972)

13
C. K. W. Tam et al.

28. Lilley, G.M.: The source of aerodynamic noise. Int. J. Aeroa- 48. Tam, C.K.W.: The spectral shape of combustion noise. Int. J.
coust. 2(3–4), 241–254 (2003) Aeroacoust. 14(3–4), 431–456 (2015). https​://doi.org/10.1016/j.
29. Kotake, S., Takamoto, K.: Combustion noise: effects of the shape jsv.2015.04.010
and size of burner nozzle. J. Sound Vib. 112(2), 345–354 (1987) 49. Strahle, W.C.: Combustion noise. Prog. Energy Combust. Sci.
30. R. Rajaram, Preetham, and T. Lieuwen. Frequency scaling of 4(3), 157–176 (1978)
turbulent premixed flame noise. AIAA Paper 2005-282. In: 11th 50. Hultgren, L.S., Miles, J.H., Jorgenson, P.C.E.: Engine system and
AIAA/CEAS Aeroacoustics Conference, Monterey, California, core noise. In: Dahl, M.D. (ed). Assessment of NASA’s Aircraft
(2005) Noise Prediction Capability, chapter 3, pages 35–62. NASA/
31. Shivashankara, B.N., Strahle, W.C., Handley, J.C.: Combustion TP-2012-215653, (2012)
noise radiated by open turbulent flames. In Nagamatsu, H.T., 51. Karchmer, A.M.: Acoustic modal analysis of a full scale annular
O’Keefe, J.V., Schwartz, I.R. (ed). Aeroacoustics: Jet and Com- combustor. AIAA Paper 1983-0760 (NASA-TM-83334). In: 8th
bustion Noise; Duct Acoustics, volume 37 of Progress in Astro- AIAA Aeroacoustics Conference, Atlanta, Georgia, (1983)
nautics and Aeronautics, pages 277–296. AIAA, (1975) 52. Royalty, C.M., Schuster, B.: Noise from a turbofan engine with-
32. Kumar, R.N.: Further experimental results on the structure and out a fan from the engine validation of noise and emission reduc-
acoustics of turbulent jet flames. In: Schwartz, I.R, Nagamatsu tion technology (EVNERT) program. AIAA Paper 2008-2810.
H.T., Strahle W. C. (eds) Aeroacoustics: Jet Noise, Combustion In: 14th AIAA/CEAS Aeroacoustics conference, Vancouver,
and Core Engine Noise, volume 43 of Progress in Astronautics British Columbia, (2008)
and Aeronautics, pages 483–507. AIAA, (1976) 53. Krejsa, E.A., Karchmer, A.M.: Acoustic modal analysis of the
33. Kilham, J.K., Kirmani, N.: The effect of turbulence on premixed pressure field in the tailpipe of a turbofan engine. Technical
flame noise. In: Seventeenth Symposium (International) on Com- Report NASA-TM-83387, NASA, (1983)
bustion, Leeds, England, 1979, pages 327–336. Proceedings of 54. Schuster, B., Lieber, L.: Narrowband model for gas turbine
the Combustion Institute, (1979) engine combustion noise prediction. AIAA Paper 2006-2677.
34. Kotake, S., Takamoto, K.: Combustion noise: effects of the veloc- In: 12th AIAA/CEAS Aerocoustics conference, Cambridge,
ity turbulence of unburned mixture. J. Sound and Vib. 139(1), Massachusetts, (2006)
9–20 (1990) 55. Motsinger, R.: Prediction of engine combustor noise and cor-
35. Ohiwa, N., Tanaka, K., Yamaguchi, S.: Noise characteristics of relation with T64 engine low frequency noise. Technical Report
turbulent diffusion flames with coherent structure. Combust. Sci. R72AEG313, General Electric Co., (1972)
Technol. 90(1–4), 61–78 (1993) 56. Emmerling, J.J., Kazin, S.B., Matta, R.K.: Core engine noise con-
36. Knott, P.R.: Noise generated by turbulent non-premixed flames. trol program, Volume III, Supplement 1—Prediction methods.
AIAA Paper 1971-0732. In: AIAA/SAE 7th Propulsion Joint Technical Report FAA-RD-74-125 III-I (AD A030376), FAA,
Specialist Conference, Salt Lake City, Utah, (1971) (1976)
37. Strahle, W.C.: A review of combustion generated noise. AIAA 57. Mathews, D.C., Rekos, N.F., Jr, Nagel, R.T.: Combustion noise
Paper 1973-1023. In: AIAA Aero-Acoustics Conference, Seattle, investigation. Technical Report FAA-RD-77-3, FAA, (1977)
Washington, (1973) 58. Mathews, D.C., Rekos Jr., N.F.: Prediction and measurement of
38. Hassan, H.A.: Scaling of combustion-generated noise. J. Fluid direct combustion noise in turbopropulsion systems. J. Aircraft
Mech. 66(3), 445–453 (1974) 14(9), 850–859 (1977)
39. Lieuwen, T., Rajaram, R.: Acoustic radiation from premixed 59. Ho, P.Y., Doyle, V.L.: Combustion noise prediction update.
flames subjected to convective flow disturbances. AIAA Paper AIAA Paper 1979-0588. In: 5th AIAA Aerocoustics Conference,
2002-0480. In: 40th AIAA Areospace Sciences Meeting, Reno, Seattle, Washington, (1979)
Nevada, (2002) 60. Zuckerman, R.S.: Core engine noise reduction: Definition and
40. Rajaram, R., Lieuwen, T.: Effect of approach flow turbulence trends. AIAA Paper 1977-1273. In: 4th AIAA Aeroacoustics
characteristics on sound generation from premixed flames. AIAA Conference, Atlanta, Georgia, (1977)
Paper 2004-0461. In: 42nd AIAA Aerospace Sciences Meeting 61. Society of Automotive Engineers International. Gas turbine jet
and Exhibit, Reno, Nevada, (2004) exhaust prediction. Technical Standard SAE ARP876 Rev. E,
41. Smith, T.J.B.: Combustion Noise. PhD thesis, University of (2006)
Leeds, (1961) 62. Zorumski, W.E.: Aircraft noise prediction program theoreti-
42. Shivashankara, B.N.: An Experimental Study of Noise Produced cal manual, part 1. Technical Report NASA-TM-83199-PT-1,
by Open Turbulent Flames. PhD thesis, Georgia Institute of NASA, (1982)
Technology, (1973) 63. Zorumski, W.E.: Aircraft noise prediction program theoreti-
43. Lieuwen, T., Mohan, S., Rajaram, R., Preetham. Acoustic radia- cal manual, part 2. Technical Report NASA-TM-83199-PT-2,
tion from weakly wrinkled premixed flames. Combustion and NASA, (1982)
Flame, 144(1-2):360–369, (2006) 64. Gillian, R.E.: Aircraft noise prediction program user’s manual.
44. Putnam, A.A.: Combustion roar of seven industrial gas burners. Technical Report NASA-TM-84486, NASA, (1982)
J. Inst. Fuel 49(400), 135–138 (1976) 65. Hultgren, L.S.: Full-scale turbofan-engine turbine-transfer func-
45. Rajaram, R., Lieuwen, T.: Acoustic radiation from turbulent pre- tion determination using three internal sensors. AIAA Paper
mixed flames. J. Fluid Mech. 637, 357–385 (2009) 2011-2912 (NASA/TM-2012-217252). In: 17th AIAA/CEAS
46. Tam, C.K.W., Pastouchenko, N.N., Mendoza, J., Brown, D: Com- Aeroacoustic Conference, Portland, Oregon, (2011)
bustion noise of auxiliary power units. AIAA Paper 2005–2829. 66. Weir, D.S.: Engine validation of noise and emission reduction
In: 11th AIAA/CEAS Aeroacoustics Conference, Monterey, technology phase I. Technical Report NASA/CR-2008-215225,
California, (2005) NASA. Honeywell Report No. 21-13843, Honeywell Aerospace,
47. Tam, C.K.W., Golebiowski, M., Seiner, J.M.: On the two compo- Phoenix, Arizona (2008)
nents of turbulent mixing noise from supersonic jets. Technical 67. Hultgren, L.S., Miles, J.H.: Noise-source separation using inter-
Report AIAA Paper 1996-1716. In: 2nd AIAA Aeroacoustics nal and far-field sensors for a full-scale turbofan engine. AIAA
Conference, State College, Pensylvania, (1996) Paper 2009-3220 (NASA/TM–2009-215834). In: 15th AIAA/
CEAS Aeroacoustic conference, Miami, Florida, (2009)

13
Aircraft noise generation and assessment

68. Bertsch, L., Guérin, S., Looye, G., Pott-Pollenske, M.: The para- 86. Tam, C.K.W., Parrish, S.A.: On the generation of indirect com-
metric aircraft noise analysis module - status overview and recent bustion noise. AIAA Paper 2014-3315. In: 20th AIAA/CEAS
applications. AIAA Paper 2011-2855. In: 17th AIAA/CEAS Aer- Aeroacoustics Conference, Atlanta, Georgia, (2014)
oacoustic conference, Portland, Oregon, (2011) 87. Zukoski, E.E., Auerbach, J.M.: Experiments concerning the
69. Bertsch, L., Dobrzynski, W., Guérin, S.: Tool development for response of supersonic nozzles to fluctuating inlet conditions. J.
low-noise aircraft design. J. Aircraft 47(2), 694–699 (2010) Eng. Power 98(1), 60–64 (1976)
70. Bertsch, L., Isermann, U.: Noise prediction toolbox used by the 88. Bohn, M.S.: Responce of a subsonic nozzle to acoustic and
DLR aircraft noise working group. Paper, 42nd International entropy disturbances. J. Sound Vib. 52(2), 283–297 (1977)
Congress and Exposition on Noise Control Engineering (Inter- 89. Strahle, W.C., Muthukrishnan, M.: Correlation of combustor rig
Noise 2013), Innsbruck, Austria, (2013) sound power data and theoretical basis of results. AIAA J. 18(3),
71. Brunet, M., Malbéqui, P., Ghedhaifi, W.: A novel approach to air 269–274 (1980)
transport system environmental impact evaluation through physi- 90. Muthukrishnan, M., Strahle, W.C., Neale, D.H.: Separation of
cal modelling and simulation. Paper, 26th International Congress hydrodynamic, entropy, and combustion noise in a gas turbine
of the Aeronautical Sciences (ICAS’08), (2008) combustor. AIAA J. 16(4), 320–327 (1978)
72. Sanders, L., Malbéqui, P., LeGriffon, I.: Capabilities of IESTA- 91. Guédel, A., Farrando, A.: Experimental study of turboshaft
CARMEN to predict aircraft noise. ICSV23 Paper 255. In: 23rd engine core noise. J. Aircraft 23(10), 763–767 (1986)
International Congress on Sound and Vibration (ICSV23), Ath- 92. Bake, F., Kings, N., Fischer, A., Röhle, I.: Indirect combustion
ens, Greece, (2016) noise: investigations of noise generated by the acceleration of
73. Brunet, M., Chaboud, T., Huynh, N., Malbéqui, P., Ghedhaifi, W.: flow inhomogeneities. Acta Acus. United Acust. 95, 461–469
Environmental impact evaluation of air transport systems through (2009b)
physical modelling and simulation. AIAA Paper 2009-6936. In: 93. Bake, F., Richter, C., Mühlbauer, B., Kings, N., Röhle, I., Thiele,
9th AIAA Aviation Technology, Integration, and Operations F., Noll, B.: The entropy wave generator (EWG): a reference case
Conference (ATIO), Hilton Head, South Carolina, (2009) on entropy noise. J. Sound Vib. 326(3–5), 574–598 (2009)
74. Gliebe, P., Mani, R., Shin, H., Mitchel, B., Ashford, G., Salamah, 94. Leyko, M., Nicoud, F., Poinsot, T.: Comparison of direct and
S., Connell, S.: Aeroacoustic prediction codes. Technical Report indirect combustion noise mechanisms in a model combustor.
NASA/CR-2000-210244, NASA, (2000) AIAA J. 47(11), 2709–2716 (2009)
75. Pickett, G.F.: Core engine noise due to temperature fluctuations 95. Leyko, M., Moreau, S., Nicoud, F., Poinsot, T.: Wave transmis-
convecting through turbine blade rows. AIAA Paper 1975-0528. sion and generation in turbine stages in a combustion-noise
In: 2nd AIAA Aero-Acoustics Conference, Hampton, Virginia, framework. AIAA Paper 2010-4032. In: 16th AIAA/CEAS
(1975) Aerocoustics Conference, Stockholm, Sweden, (2010)
76. Bake, F., Michel, U., Röhle, I., Richter, C., Thiele, F., Liu, M., 96. Durán, I., Moreau, S.: Analytical and numerical study of the
Noll, B.: Indirect combustion noise generation in gas turbines. entropy wave generator experiment on indirect combustion noise.
AIAA Paper 2005-2830. In: 11th AIAA/CEAS Aeroacoustics AIAA Paper 2011-2829. In: 17th AIAA/CEAS Aeroacoustic
Conference, Monterey, California, (2005) Conference, Portland, Oregon, (2011)
77. Bake, F., Michel, U., Röhle, I.: Investigation of entropy noise in 97. Goh, C.S., Morgans, A.S.: Phase prediction of the responce of
aero-engine combustors. Paper GT2006-90093, ASME Turbo choked nozzles to entropy and acoustic disturbances. J. Sound
Expo 2006, Barcelona, Spain, (2006) Vib. 330(21), 5184–5198 (2011)
78. Bake, F., Michel, U., Röhle, I.: Experimental investigation of the 98. Giauque, A., Huet, M., Clero, F.: Analytical analysis of indirect
fundamental entropy noise mechanism in aero-engines. AIAA combustion noise in subcritical nozzles. J. Eng. Gas Turbines
Paper 2007-3694. In: 13th AIAA/CEAS Aeroacoustics Confer- Power, 134(11):111202–1–8, (2012)
ence, Rome, Italy, (2007) 99. Morgans, A.S., Goh, C.S., Dahan, J.A.: The dissipation and shear
79. Bake, F., Michel, U., Röhle, I.: Fundamental mechanisms of dispersion of entropy waves in combustor thermoacoustics. J.
entropy noise in aero-engines: experimental investigation. Paper Fluid Mech. (JFM Rapids), 733:R2–1–11, (2013)
GT2007-27300, ASME Turbo Expo 2007, Montreal, Canada, 100. Leyko, M., Moreau, S., Nicoud, F., Poinsot, T.: Numerical and
(2007) analytical modelling of entropy noise in a supersonic nozzle with
80. Bake, F., Michel, U., Röhle, I.: Investigation of entropy noise a shock. J. Sound Vib. 330(16), 3944–3958 (2011)
in aero-engine combustors. J. Eng. Gas Turbines Power 129(2), 101. Durán, I., Moreau, S.: Study of the attenuation of waves propa-
370–376 (2007c) gating through fixed and rotating turbine blades. AIAA Paper
81. Kings, N., Bake, F.: Indirect combustion noise: noise generation 2012-2133. In: 18th AIAA/CEAS Aerocoustics Conference,
by accelerated vorticity in a nozzle flow. Int. J. Spray Combust. Colorado Springs, Colorado, (2012)
Dyn. 2(3), 253–266 (2010) 102. Durán, I., Moreau, S.: Numerical simulation of acoustic and
82. Kings, N., Enghardt, L., Bake, F.: Indirect combustion noise: entropy waves propagating through turbine blades. AIAA Paper
Experimental investigation of the vortex sound generation in 2013-2102. In: 19th AIAA/CEAS Aeroacoustic Conference, Ber-
accelerated swirling flows. AIAA Paper 2013-2100. In: 19th lin, Germany, (2013)
AIAA/CEAS Aeroacoustic Conference, Berlin, Germany, (2013) 103. Mishra, A., Bodony, D.J.: Evaluation of actuator disk theory
83. Kings, N., Enghardt, L., Bake, F.: Broadband indirect noise gen- for predicting indirect combustion noise. J. Sound Vib. 332(4),
eration by accelerated vorticity. AIAA Paper 2015-2820. In: 21st 821–838 (2013)
AIAA/CEAS Aeroacoustic Conference, Dallas, Texas, (2015) 104. Cumpsty, N.A., Marble, F.E.: The interaction of entropy fluctua-
84. Fischer, A., Bake, F., Röhle, I.: Broadband entropy noise phe- tions with turbine blade rows; a mechanism of turbojet engine
nomena in a gas turbine combustor. Paper GT2008-50263, noise. Proc. R. Soc. Lond. A 357, 323–344 (1977a)
ASME Turbo Expo 2008, Berlin, Germany, (2008) 105. Goh, C.S., Morgans, A.S.: The infuence of entropy waves on the
85. Chu, B.-T., Kovasznay, L.S.G.: Non-linear interactions in a thermoacoustic stability of a model combustor. Combust. Sci.
viscous heat-conducting compressible gas. J. Fluid Mech. 3, Technol. 185(2), 249–268 (2013)
494–514 (1958) 106. Tam, C.K.W., Parrish, S.A., Xu, J., Schuster, B.: Indirect com-
bustion noise of auxiliary power units. J. Sound Vib. 332, 4004–
4020 (2013)

13
C. K. W. Tam et al.

107. Knobloch, K., Werner, T., Bake, F.: Entropy noise generation AIAA Paper 2015-2819. In: 21st AIAA/CEAS Aeroacoustic
and reduction in a heated nozzle flow. AIAA Paper 2015-2818. Conference, Dallas, Texas, (2015)
In: 21st AIAA/CEAS Aeroacoustic Conference, Dallas, Texas, 120. Cumpsty, N.A., Marble, F.E.: Core noise from gas turbine
(2015) exhausts. J. Sound Vib. 54(2), 297–309 (1977b)
108. Miles, J.H.: Time delay analysis of turbofan engine direct and 121. Tam, C.K.W., Parrish, S.A.: Noise of high-performance aircraft
indirect combustion noise sources. J. Propuls. Power 25(1), at afterburner. J. Sound Vib. 352, 103–128 (2015)
218–227 (2009) 122. Papadogiannis, D., Wang, G., Moreau, S., Duchaine, F., Gic-
109. Miles, J.H.: Separating direct and indirect engine combustion quel, L., Nicoud, F.: Assessment of the indirect combustion noise
noise using the correlation function. J. Propuls. Power 26(5), generated in a transonic high-pressure turbine stage. J. Eng. Gas
1144–1152 (2010) Turbines Power, 138(4):041503–1–8, (2016)
110. Bake, F., Gaetani, P., Persico, G., Neuhaus, L., Knobloch, K.: 123. Tam, C.K.W.: Computational aeroacoustics: a wavenumber
Indirect noise generation in a high pressure turbine stage. AIAA approach. Cambridge University Press, Cambridge (2012)
Paper 2016-3004. In: 22nd AIAA/CEAS Aeroacoustic Confer- 124. Tam, C.K.W., Li, Z., Schuster, B.: An investigation on indirect
ence, Lyon, France, (2016) combustion noise generation in a turbofan engine. AIAA Paper
111. Pardowitz, B., Tapken, U., Knobloch, K., Bake, F., Bouty, E., 2016-2746. In: 22nd AIAA/CEAS Aeroacoustic Conference,
Davis, I., Bennett, G.: Core noise—identification of broadband (2016)
noise sources of a turbo-shaft engine. AIAA Paper 2014-3321. 125. Leyko, M., Durán, I., Moreau, S., Nicoud, F., Poinsot, T.: Simula-
In: 20th AIAA/CEAS Aeroacoustics Conference, Atlanta, Geor- tion and modelling of the waves transmission and generation in a
gia, (2014) stator blade row in a combustion-noise framework. J. Sound Vib.
112. Chung, J.Y.: Rejection of flow noise using a coherence function 333(23), 6090–6106 (2014)
method. J. Acoust. Soc. Am. 62(2), 388–395 (1977) 126. Durán, I., Moreau, S., Poinsot, T.: Analytical and numerical study
113. Hsu, J.S., Ahuja, K.K.: A coherence-based technique to separate of combustion noise through a subsonic nozzle. AIAA J. 51(1),
internal mixing noise from farfield measurements. AIAA Paper 42–52 (2013)
1998-2296. In: 4th AIAA/CEAS Aeroacoustic Conference, Tou- 127. Schlimpert, S., Koh, S.R., Pausch, K., Meinke, M., Schröder,
louse, France, (1998) W.: Analysis of combustion noise of a turbulent premixed slot
114. Minami, T., Ahuja, K.K.: Five microphone method for sepa- jet flame. Combust. Flame 175, 292–306 (2017)
rating two different noise sources from farfield measurements 128. Ihme, M., Pitsch, H., Bodony, D.J.: Radiation of noise in turbu-
contaminated by extraneous noise. AIAA Paper 2003-3261. In: lent non-premixed flames. Proc. Combust. Inst. 32, 1545–1554
9th AIAA/CEAS Aeroacoustic Conference, Hilton Head, South (2009)
Carolina, (2003) 129. Livebardon, T., Moreau, S., Gicquel, L., Poinsot, T., Bouty, E.:
115. Davis, I., Bennett, G.J.: Experimental investigations of coherence Combining les of combustion chamber and an actuator disk the-
based noise source identification techniques for turbomachinery ory to predict combustion noise in a helicopter engine. Combust.
applications - classic and novel techniques. AIAA Paper 2011- Flame 165, 272–287 (2016)
2830. In: 17th AIAA/CEAS Aeroacoustic Conference, Portland, 130. Poinsot, T., Veynante, D.: Theoretical and Numerical Combus-
Oregon, (2011) tion. elearning.cerfacs.fr/combustion, third edition, (2012)
116. Davis, I., Bennett, G.J.: Spatial noise source identification of 131. Pitsch, H.: Large-eddy simulation of turbulent combustion. Ann.
tonal noise in turbomachinery using the coherence function on a Rev. Fluid Mech. 38, 453–482 (2006)
modal basis. AI 2011-2825. In: 17th AIAA/CEAS Aeroacoustic 132. Poinsot, T.: Prediction and control of combustion instabilities in
Conference, Portland, Oregon, (2011) real engines. Proc. Combust. Inst. 36(1), 1–28 (2017)
117. Enghardt, L., Holewa, A., Tapken, U.: Comparison of different 133. Jaravel, T., Ribner, E., Cuenot, B., Bulat, G.: Large eddy simu-
analysis techniques to decompose a broad-band ducted sound lation of a model gas turbine burner using reduced chemistry
field in its mode constituents. AIAA Paper 2007-3520. In: 13th with accurate pollutant prediction. Proc. Combust. Inst. 36(3),
AIAA/CEAS Aeroacoustics Conference, Rome, Italy, (2007) 3817–3825 (2017)
118. Jürgens, W., Tapken, U., Pardowitz, B., Kausche, P., Bennett, 134. Durán, I., Moreau, S.: Solution of the quasi-one-dimensional
G.J., Enghardt, L.: Technique to analyze characteristics of tur- linearized Euler equations using flow invariants and the Magnus
bomachinery broadband noise sources. AIAA Paper 2010-3979. expansion. J. Fluid Mech. 723, 190–231 (2013b)
In: 16th AIAA/CEAS Aerocoustics Conference, Stockholm,
Sweden, (2010)
119. Schuster, B., Gordon, G., Hultgren, L.S.: Dynamic temperature
and pressure measurements in the core of a propulsion engine.

13

Potrebbero piacerti anche