Anda di halaman 1dari 17

Aerospace Science and Technology 12 (2008) 117 www.elsevier.

com/locate/aescte

Aircraft noise reduction technologies: A bibliographic review


D. Casalino a,,1 , F. Diozzi b,2 , R. Sannino b,3 , A. Paonessa c
a Rotorcraft Aerodynamics and Aeroacoustic Department, Italian Aerospace Research Center, via Maiorise, Capua, I-81043, Italy b Documentation Center, Italian Aerospace Research Center, via Maiorise, Capua, I-81043, Italy c Acoustics Department, Alenia Aeronautica, viale dellAeronautica, Pomigliano dArco, I-80038, Italy

Available online 22 October 2007

Abstract A bibliographical review of the main technologies employed for the mitigation of aircraft noise is presented. According to a componentbased approach, analytical and semi-empirical models of the aeroacoustic mechanisms involved in the noise generation from airframe and engine components are presented as a key element of the noise reduction technology. These models, developed in the past to investigate the inuence of some design parameters on the overall acoustic levels, are nowadays powerful design tools when employed in a multi-disciplinary optimization framework. In this spirit, the recent achievements in the numerical prediction of complex aeroacoustic phenomena through CFD/CAA techniques, not addressed in this work, provide a complementary approach to experiments to improve the accuracy of the available analytical and semiempirical models. The bibliographical style of the paper is guaranteed by a qualitative description of the underlying physical mechanisms and their mathematical idealization. The reader is therefore remanded to the cited works for a deeper analysis. 2007 Elsevier Masson SAS. All rights reserved.
Keywords: Aircraft noise; Airframe noise; Fan noise; Jet noise

1. Motivations Aircraft and aero-engine manufacturers are exposed to a continuously growing demand for quieter aircraft. This is due on one hand to the increased community expectation for quality of life, and on the other hand to the necessity to compensate both the growth in air trafc and the encroachment of airportneighboring communities. Since the theoretical work by Sir James Lighthill in 1952 [99] on sound generated aerodynamically that led to the rst predictive model of jet noise, many progresses have been made in the physical understanding of several aeroacoustic mechanisms and their mathematical representation. Recent achievements in the numerical modeling of complex phenomena, like
* Corresponding author.

E-mail addresses: d.casalino@cira.it (D. Casalino), f.diozzi@cira.it (F. Diozzi), r.sannino@cira.it (R. Sannino), apaonessa@aeronautica.alenia.it (A. Paonessa). 1 Senior research engineer. 2 Head. 3 Information specialist. 4 Head. 1270-9638/$ see front matter 2007 Elsevier Masson SAS. All rights reserved. doi:10.1016/j.ast.2007.10.004

the screech tone generation in supersonic jets [13,156], contribute to extend the knowledge domain beyond the empirical evidence of these phenomena. The growth in the theoretical description of many aeroacoustic mechanisms in the past fty years has been accompanied by a progressive reduction of aircraft noise. Since the Sixties the historical aircraft noise trend shows a reduction of about 20 EPNdB, mostly due to the progressive introduction into service of high-bypass turbofans and more effective nacelle acoustic treatments. Since the Eighties, however, the noise reduction trend has not been so signicant. Therefore, any further noise reduction is very difcult to be achieved without affecting the aircraft operating cost. Due to the progresses achieved in reducing the propulsive noise and due to the expected reduction with the entry into service of ultra high-bypass ratio turbofans and novel noise control devices, on modern civil aircraft the engine noise is expected to be comparable and even lower than the airframe noise generated by the high-lift devices and by the undercarriage. From these considerations it is clear that aircraft manufacturers have to comply with a new emerging necessity: the environmental concern, in terms of acoustic and chemical pollution, must be considered as a driving factor in the de-

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

sign and operation of any new commercial aircraft. In other words, the preliminary design of an aircraft is the result of re-dened tradeoffs between operating costs and environmental performances, in a multi-disciplinary framework involving aircraft operational factors, aircraft/engine performances, airframe/powerplant noise modeling, powerplant emission modeling, and operational costs. Since multi-disciplinary/multi-objective optimization processes relies on fast numerical methods, a great deal of interest have been devoted during the last years to the re-discovery and improvement of analytical models for both airframe and powerplant noise prediction. The works recently published by Antoine and Kroo [8,9] and by Lilley [100,104] provide elegant examples of how the noise requirement can drive the aircraft design and its operation. The main goal of the present paper is to provide a literary review on the subject of airframe, fan and jet noise theory. Emphasis is placed on fully analytical or semi-empirical models that are commonly used by the major aerospace research establishments. For each noise generation mechanism, an overview of the main noise mitigation techniques is also provided. This review constitutes a knowledge basis for the development of a library of methods for aircraft noise prediction to be exploited in the joined AleniaAeronautica/CIRA research programmes. 2. Airframe noise The aerodynamic noise generated by all the non-propulsive components of an aircraft is classied as airframe noise. For modern high-bypass engine powered commercial aircraft, the airframe noise represents the main contribution to the overall yover noise levels during landing approach phases, when the high-lift devices and the landing-gear are deployed. Five main mechanisms are recognized to contribute signicantly to the airframe noise: (i) the wing trailing-edge scattering of boundary-layer turbulent kinetic energy into acoustic energy, (ii) the vortex shedding from slat/main-body trailing-edges and the possible gap tone excitation through nonlinear coupling in the slat/ap coves, (iii) the ow unsteadiness in the recirculation bubble behind the slat leading-edge, (iv) the roll-up vortex at the ap side edge, (v) the landing-gear multi-scale vortex dynamics and the consequent multi-frequency unsteady force applied to the gear components. All these mechanisms have been addressed both experimentally and theoretically since the Seventies. The most complete review on the topic is due to Crighton [31], who contributed considerably to the mathematical modeling of several aeroacoustic mechanisms. Nowadays theoretical knowledge of the ow phenomena involved in the airframe noise generation, together with the achieved readiness level of several numerical methods in the CFD/CAA domain (not addressed in this paper), are favorable conditions for a signicant technological advancement in the eld of airframe noise control and reduction. As a consequence, a great effort is currently undertaken by the main aircraft manufacturers for developing novel high-lift devices and optimized landing-gear assemblies.

2.1. Analytical models 2.1.1. Trailing-edge noise In clean conguration, the main source of airframe noise is represented by the wing trailing-edge noise. A solid surface immersed in a turbulent ow has a dual inuence on the radiated acoustic eld. On one hand, it affects the structure of the ow eld and, consequently, of the aeroacoustic sources. On the other hand, it constitutes an acoustic impedance discontinuity which affects the scattering of the acoustic waves. The noise from a trailing-edge is a singular problem for which a wrong separation of these two effects can lead to wrong theoretical results. A trailing-edge in a uctuating ow eld generates an unsteady vortical wake. This phenomenon can be regarded as an unsteady boundary-layer separation due to the uid viscosity. An inviscid model of the vortex shedding process consists in prescribing an edge condition. Since the vortex shedding smears the singular behavior of the ow at the trailing-edge, a Kutta condition is commonly imposed at the trailing-edge, which requires that the ow velocity is nite at the edge. The physical relation between the smoothing effect due to the vortical wake and the Kutta condition is not completely clear. The experimental works of Archibald [10] and Satyanarayana and Davis [152], for example, show that the Kutta condition is only partially fullled at the trailing-edge of an airfoil in a high-frequency perturbed eld. A viscous ow, in fact, has a characteristic relaxation time over which the ow reacts to an imposed disturbance. If this relaxation time is greater than the characteristic period of the perturbation eld, the ow would not have enough time to fully satisfy the Kutta condition. From these preliminary considerations it follows that the aeroacoustic sources are signicantly affected by the hypothesis made on the behavior of the ow at a trailing-edge. This is a rst difculty in modeling the trailing-edge noise. Another difculty lies on the fact that an edge does not distinguish between an acoustic and a vortical disturbance. As a result, the unsteady pressure eld induced by the wake itself can couple with the vortex shedding process causing an increase of the noise levels. A further difculty is related to the refraction effect of a boundary shear ow. This imposes some restrictions on the applicability of an acoustic analogy model that requires a clear separation between the vortical aeroacoustic sources and the diffracting effects due to the edge. Depending on the strategy used to face these difculties, different theoretical results have been obtained in the past. Crighton [28], following the work of Orszag and Crow [128], investigated the interaction between an acoustic incident eld, an unstable shear layer and a at-plate trailing-edge. He showed that, at low Mach numbers, a Kutta condition induces a change in the ow velocity dependence of the acoustic intensity 4 to U 2 , provided that the vortex-sheet is unstable. from U Jones [89] investigated the diffraction of the acoustic eld generated by a source near the edge of a semi-innite at-plate. He concluded that the Kutta condition generates an intense beaming effect along the plane of the plate. Crighton and Leppington [34] did not obtained signicant effects on the acoustic

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

radiation due to the Kutta condition. Davis [35] investigated the effect of the Kutta condition on the at-plate problem and concluded that, without a Kutta condition, the far-eld noise ex3 acoustic power hibits a sin2 (/2) directivity pattern and a U law. Conversely, with a Kutta condition, a beaming effect along the wake takes place and the noise intensity level increases as U . Howe [83], in his comprehensive review work, pointed out that both Crightons [28] and Davis [35] analysis were erroneous since not compatible with a far-eld wave radiation condition. Howe [82] demonstrated that the imposition of the Kutta condition removes the ow singularity at a sharp edge. As a consequence, it always leads to a reduction of the noise levels. The wake intensity, in fact, is related, both in amplitude and phase, to the incident vortical disturbances and generates a sound eld which cancels the one generated by the incident turbulent eld. A different view of the trailing-edge problem consists in focusing separately on the acoustic scattering properties of the edge, and the uctuating pressure eld close to the edge. Ffowcs Williams and Hall [48] and Crighton and Leppington [32,33] examined the scattering problem in the context of Lighthills acoustic analogy theory. They related the acoustic far eld to the turbulent quadrupole sources. Chase [21,22] and Chandiramani [20] proposed a procedure to relate the far-eld acoustic spectrum to measurable statistical properties of the hydrodynamic pressure eld in proximity of the edge. These properties are formally synthesized by the wavenumber/frequency spectrum of the driving pressure eld. Rienstra [142] examined the half-plane scattering problem of an acoustic incident eld in the presence of a vortex-sheet shed from the trailing-edge. He showed that the vortex shedding process extracts energy from both the incident acoustic eld and the mean ow, thus reducing the far-eld noise intensity. However, the interaction between the trailing-edge and the unsteady pressure eld induced by the wake generates a noise radiation whose energy may exceed that absorbed by the wake. Crighton [29], using the method of matched asymptotic expansions, determined the noise radiated by a line-vortex convected past the edge of a semi-innite at-plate under the inuence of its image vortex. The intensity of the acoustic eld driven by the vortex-induced hydrodynamic eld was demonstrated to have a sin2 (/2) directivity pattern and a third power dependence on the ow velocity. The latter result was in agreement with the general result obtained by Ffowcs Williams and Hall [48] and Crighton and Leppington [32] according to which the effect of the interaction between a uctuating turbulent ow and the edge of a semi-innite plate is to increase the far-eld 3 . In fact, as demonstrated by acoustic intensity by a factor M Obermeier [127], the acoustic energy radiated by a vortex lament in free space or in the presence of an innite rigid plate follows a sixth power law. Amiet [3,6] related the acoustic spectrum to the wall pressure spectrum through an airfoil response function. In order to use a wall pressure distribution with the same characteristics it would have in the absence of the trailing-edge, he assumed that the turbulence was statistically stationary when convected past the trailing-edge. A concentrated dipole sources induced

by the turbulent ow near the trailing-edge was used to model the noise radiation. The airfoil response function used by Amiet was obtained under the assumption that, at high frequency, as relevant for a trailing-edge noise problem, the leading- and trailing-edge noise scattering can be taken into account separately in an iterative converging procedure [5]. This approach had been already applied by Landahl [97] to an innite-span airfoil embedded in a parallel gust of arbitrary wavelength, and later on by Adamczyk [1] to an innite-span swept wing interacting with an oblique gust. Amiet [5] extended Adamczyks analysis in order to account for a difference between the free-stream velocity and the convection velocity of the gust. In recent years trailing-edge broadband noise models of Amiet type, involving iterative leading- trailing-edge corrections, have been developed and validated against experimental data in practical cases representative of airframes, wing turbines, helicopter blades and fan cooling systems by Roger and Moreau [117,118, 147,148]. Howe [82] investigated the general problem of the noise from an airfoil interacting with a frozenly convected turbulent 2 1), he eddy. By supposing a small ow Mach number (M assumed an incompressible potential ow. The acoustic problem was then formulated in terms of Howes [81] acoustic analogy theory describing the noise generated by vorticity and entropy gradients. In the limit of a frozen convection hypothesis Howe showed that, when a line-vortex is convected past the edge of a half-plane, a Kutta condition results in a vortex shedding that exactly cancels the sound generated by the vortex-edge interaction. Furthermore, in the case of an acoustically compact airfoil, the effect of the vortex shedding is to cancel the eld diffracted by the trailing-edge. Since the fulllment level of the Kutta condition decreases as the frequency of the perturbation ow increases, the effectiveness of the vortical wake in reducing the radiated noise decreases at higher frequency. As a consequence, a trailing-edge in a turbulent ow acts as a source of high-frequency noise. Howe [83] revisited the problem of the trailing-edge noise and formulated a generalized theory. This includes, as special cases, the models developed by Ffowcs Williams and Hall [48], Crighton [29], Chase [21,22] and Chandiramani [20]. All these models were shown to give essentially the same results when properly interpreted. In particular, the so-called cardioid sin2 (/2) directivity law and the fth power scaling law of the acoustic intensity on the free-stream velocity were recovered. Moreover, Howe investigated the inuence of the Kutta condition on the noise levels by assuming a wake convection velocity w which differs from the convection velocity v of the vortical disturbances within the boundary-layer. He found that the effect of the Kutta condition is to reduce the sound pressure level by a factor (1 w/v)2 , which diverges when the Kutta condition is fully satised (w = v). A rst experimental validation of the trailing-edge noise theory by Howe was carried-out by Brooks and Hodgson [16] who measured the noise generated by a NACA-0012 airfoil in a low Mach number ow at several angles of attack and with different degrees of edge bluntness. The airfoil was provided of roughness trips on both its sides in order to ensure a well de-

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

veloped turbulent boundary-layer. In order to investigate the vortex shedding process in terms of wake convection velocity w , as proposed by Howe [83], Brooks and Hodgson carried-out coherence measurements between a cross hot wire just downstream of the edge and a wall pressure transducer close to the edge. No vortex shedding was detected. From this result and the veried consistency of the evanescent wave model proposed by Chase [22] in the edge region, they concluded that a vanishing wake velocity can be used in the trailing-edge noise model. 5 acoustic In addition a good agreement with the theoretical U 2 power law and the sin (/2) directivity pattern was found. 2.1.2. Flap side-edge noise One of the most effective sources of airframe noise at takeoff and landing conditions is due to the vortical ow around the side edge of a deployed ap. The physical mechanism of ap side-edge noise generation can be described as follows. The pressure jump across the upper and lower surfaces of the ap creates a recirculating ow around the side edge. The shear layer detaches at the side edge of the ap and rolls up to a single vortical structure. This is responsible for two noise generation mechanisms: the rst is due to a direct interaction between the shear layer uctuations and the sharp edge, the second is due to the induction of unstable oscillation modes in the vortical structure. In addition, vortex breakdown has been observed at high ap angles as an additional noise source mechanism. The problem of sound generation by the vortex roll-up about a deected side-edge has been investigated analytically by Hardin [73] and by Howe [84]. The model put forward by Hardin is based on the analogy according to which, the transient streamwise vortex roll-up can be seen, in a plane normal to the edge and the ow direction, as a two-dimensional vortex that is swept around the edge of half-plane under the effect of self induction. The model presents strong analogies with the trailing-edge noise model proposed by Crighton [29] and 5 acoustic power law and the therefore suggested the same U 2 same sin (/2) directivity pattern. A drawback of this analytical model is that, not only the hydrodynamic eld, but also the driven acoustic eld is assumed to be two-dimensional. A different model was proposed by Howe [84]. By using an acoustic analogy approach, he separated the sound generation mechanisms from the acoustic edge scattering problem. Thus he represented all the involved vortex dynamic processes in terms of a prescribed statistical surface pressure distribution, and focused on the Greens function of a simplied geometrical conguration. In particular, he obtained analytical expression for a nite chordwise slot in a wing with an otherwise straight trailing-edge, both in the low- and high-frequency limit. The most important result obtained by Howe is that the radiation efciency of the side-edge sources is greater for low values of the dimensionless frequency referred to the slot clearance and the spanwise ow velocity. This is because, at low frequency the sound generation mechanism is well represented by a dipoletype source, whereas at high frequency the sound generation mechanism is dominated by a monopole-type source that reproduces the mass ux through the slot.

A ap side-edge noise model derived on the basis of the aforementioned analyses have been recently validated against experimental data by Molin et al. [115]. 2.1.3. Slat/ap cove and trailing-edge noise While the dominant sound generated by a slat wing component is of broadband nature and due to the ow recirculating bubble behind the leading-edge and to turbulence convected past the trailing-edge through the gap, there exists a certain experimental evidence of tonal noise radiation. This is commonly attributed to the vortex shedding from the blunt trailing-edge. In certain cases, however, a surprisingly high tone have been observed, that disappears when the slat deection angle is slightly modied. This behavior was attributed by Khorramiet al. [92] to a feedback loop between the slat trailing-edge and the main wing surface which drives the vortex shedding frequency. A numerical demonstration of the gap whistle generation has been pursued by Tam and Pastouchenko [167]. The predicted tone frequencies was shown to agree fairly well with a simple algebraic model of feed-back loop based on the principle of causeeffect phase compatibility. 2.1.4. Landing-gear noise The sound generation mechanism from a landing-gear is due to the vortex-force generated by the quasi-periodic unsteady ow separation behind the different structural components. The mechanism is similar to the so-called Aeolian tones generation from circular cylinders [18,133], but complicated by the simultaneous shedding from bodies of different size and shape, and by the mutual vortex interactions. The resulting noise is of broadband nature, spanning over a wide interval in the audibility range. The dipolar character of the source mechanism results 6 acoustic power scaling law. However, due to acoustic inin U stallation effects, a destructive interference of sound generated by dipoles in the vertical plane, i.e. generated by vorticity vector components parallel to the reecting airframe panels, may occur, thus resulting in a quadrupole source degeneration and 8 acoustic power scaling law. Several semia corresponding U empirical models have been developed by aircraft manufactures both in the US and Europe. The Airbus model, for instance, has been recently applied to evaluate the impact of a low noise gear on the overall airframe noise [114]. A further installation aspect that should be addressed in landing-gear noise modeling is the excitation of some cavity mode in the wheel wells. Three different mechanisms of tonal noise response of a cavity ow exist. The rst one is due to a feedback cycle generated by the coupling between the waves generated when the shear layer vortical uctuations impinge on the downstream edge, and the vortex shedding from the upstream edge [23,75,76,85,86,149,165]. The second mechanism involves a volume mode uctuation of the recirculating bubble induced by a coupling with a standing wave across the width or along the length of the cavity. The third mechanism is a Helmholtz resonation occurring when the ux mass across the cavity neck balance the uid volume stiffness. A complete review of self-sustaining mechanisms in cavity ows is due to Rockwell and Naudascher [146]. Fortunately, realistic

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

cavities, in the presence of uncontrolled grazing ow regimes, seem to respond only in their depth mode [14,40]. The resulting acoustic eld can be easily modeled by extrapolating the internal cavity pressure levels at the depth mode frequency to the far eld with a monopole directivity pattern. 2.1.5. Component-based empirical and semi-empirical models The airframe noise source breakdown presented in this section constitutes the basis for a component-based comprehensive aircraft noise modeling. The typical approach consists in describing the acoustic far-eld spectrum as resulting from the uncorrelated superposition of different effects, each related to macroscopic ow quantities via ad-hoc scaling laws. Several parameters appear in this kind of comprehensive representation that must be determined through dedicated experimental campaigns or numerical simulations. The comprehensive aircraft noise model which is universally accepted, because of its wide experimental validation undertaken in the US, is that one developed by Fink [51,52]. A similar approach, but involving different component noise models, have been recently proposed and validated experimentally by Lilley [104]. Both the Finks and Lilleys model are based on tuning coefcients that must be evaluated experimentally. A drawback of Finks model is that the high-lift device noise does not account for the ap side-edge noise, which is demonstrated to predominate on the trailing-edge noise. Taking advantage of this newly accepted source component breakdown and of the recent achievements in phased microphone array measurement techniques, Guo et al. [68] developed a fully empirical airframe noise model. A different approach has been recently developed by the main aircraft manufacturers, although few papers have been published on the subject. It consists in performing Reynoldsaveraged NavierStokes simulations to obtain the characteristic velocity and length scales to be used in analytical models. An application to trailing-edge noise predictions in the context of multidisciplinary design and optimization has been published by Hosder et al. [80]. 2.2. Technologies for control and reduction 2.2.1. Trailing-edge noise The fundamental concept behind a trailing-edge noise mitigation is that of modifying the surface impedance close to the edge in order to reduce the impedance jump felt by eddies convected past the edge. This can be realized by using porous edges, by applying compliant brushes or by using sawtooth edges. The latter technique also reproduces the effect of a swept angle between the edge and the spanwise vortex components, resulting in a trailing-edge noise reduction according to theory [83]. Although the simultaneous use of sawtooth and porous edges may result in a more signicant noise mitigation, full scale experimental campaigns carried out in the US [15] showed that only the sawtooth trailing-edge reduces the EPNL by 2 dB. The sawtooth edge has been also used to reduce the tone noise generation from a slat trailing-edge.

2.2.2. Flap side-edge noise In order to reduce the ap side-edge noise, different ow control strategies have been proposed in the recent years. Passive control devices include ap-tip fences [159,176], porous treatment of the ap-tip region [7,137], serrated trailing-edge corners, the so-called noise weeder [122], and small vortex generators referred to as microtabs [122]. Active control techniques consist in blowing continuously air into the vortical structure in order to counteract the vortex roll-up, and to displace the vortical structure away from the solid surface [93,94]. Both passive and active devices have been shown to mitigate the ap side-edge noise. However, performing a comprehensive technological evaluation of the different devices is a quite hard task. 2.2.3. Slat noise The development and test of noise reduction devices for the high-lift devices, the slat in particular, are constrained by the fact that these devices must not reduce the lift coefcient of the wing. Three technologies for slat noise reduction have been proposed and tested, both in the US and Europe. The rst consists in applying brushes [24] or serrated tabs on the slat suction side near the trailing-edge [122]. This allows to mitigate the vortex shedding, resulting in a quite signicant reduction of the corresponding tonal noise peak. The second slat noise reduction technique consists in lling the volume of the recirculating bubble behind the leading edge, resulting in a quite signicant reduction of the broadband noise levels [39,122]. Finally, the third technology consists in using slat gap acoustic liners. This device has the advantage of not affecting the aerodynamic properties of the wing. The use of liners is legitimated by the fact that the high-frequency wave emitted from the trailing-edge can be attenuated via multiple reections on the two sides of the gap, during the wave travel from the trailing-edge to the lower gap opening. The slat gap liner noise mitigation properties have been recently demonstrated by Smith et al. [158]. 2.2.4. Landing-gear noise The reduction of bluff gear components and the passive control of vortex shedding via splitter plates behind the cylindrical struts are the solutions commonly adopted to reduce the landing-gear noise. In addition, a practical alternative to a structural re-design of the gear consists in using add-on fairings and covers. The noise reduction potentialities of several addon devices applied to a full-scale A340 landing gear have been demonstrated by Dobrzynski et al. [38]. 2.3. Research trends Three main trends in the airframe noise research can be observed. The rst consists in combining all the available single source models into comprehensive tools where the several tuning parameters are obtained either through dedicated experiments or advanced CFD simulations. These tools can be used to perform tradeoff studies during preliminary design stages, when the focus is on multidisciplinary design and optimization. The second research trend is the increasing use of CFD

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

simulations to design and optimize noise reduction devices. In addition, the numerical simulation contributes to increase the physical knowledge of the sound generation mechanisms and to the improvement of the source models. Finally, the third trend in the eld of airframe noise consists in developing advanced measurement techniques that permit to discriminate between concurring noise generation mechanisms, and to generate databases for the validation of predictive numerical tools. 3. Fan noise

by-pass nozzle. Since the rst two noise mitigation concepts requires analytical models that highlight the mutual inuence of all the design parameters, a brief overview of the some available analytical models is presented hereafter. In addition, since the optimization of both passive and active control devices relies on duct noise transmission models and far-eld radiation models, a short review of the available transmission/radiation analytical models is presented. Finally, an overview of the main technologies for fan noise reduction and control is drawn. 3.1. Analytical models

In turbofan aero-engines, noise is generated by the interaction between ow non-uniformities and rotating bladed and stator vanes. In modern high-bypass-ratio turbofans, the noise generated by the fan system exceeds the one generated by the compressor and by the turbine stages. Due to a coupling between the aeroacoustic excitation mechanisms and the duct acoustic modes, in subsonic blade tip conditions the rotor-locked constant dipole source drives an evanescent duct mode that is attenuated during the sound transmission through the duct. Only higher order blade-passagefrequency harmonics generated by the ow unsteadiness due to the rotorstator interaction mechanisms can be transmitted through the duct. Conversely, at supersonic blade tip conditions, the rotor-locked shock wave system generates propagative multiple pure tones at harmonics of the rotational shaft frequency, the so called buzz-saw noise. The fan-noise interactional sources are concentrated near the blade leading- and trailing-edges. Leading-edge sources on the rotor blades are generated by inow disturbances such as ingested atmospheric turbulence, duct boundary layer turbulence and interaction with ow distortion. Leading-edge sources on the stator vanes are generated by the quasi-periodic impingement of the rotor viscous wake. Trailing-edge sources on both the rotor blades and the stator vanes are generated by small scale vortical disturbances convected past the trailing-edge. The rotorstator interaction noise has both a tonal and a broadband spectral content. The tonal spectral components are due to the deterministic interaction between the coherent part of the rotor wakes and the stator vanes, whereas the broadband spectral components are generated by the turbulent vortical disturbances. A spectral broadening around the blade passage frequency harmonics is also observed, which is due to a random modulation of the wake/stator interaction process. The reduction of fan noise radiation to the far eld can be pursued by ve generical concepts: (i) mitigation of the interaction mechanisms through an optimal design of the rotor blades and the stator vanes, or via ow control techniques that reduces the velocity decit in the rotor wakes, (ii) tuning of the stator cascade parameters in order to reduce the aerodynamic response to an impinging gust, (iii) tuning of the rotor blades and stator vanes numbers in order to drive only few propagating (cut-on) duct modes, (iv) use of passive/active duct wall treatments in order to attenuate the sound during transmission through the duct, (v) manipulation of sound diffraction mechanisms at the inlet lip and exhaust nozzle via advanced nacelle devices, such as the negatively scarfed inlet or the serrated

3.1.1. Gustairfoil interaction models The unsteady pressure eld induced by a vortical ow convected past a blade generates an unsteady force on the blade and interaction noise in the far eld. Convected vortical disturbances that have a regular space-time structure are modeled as velocity gusts. For examples, the tip-vortices shed from helicopter blades that impinge with regularity on the following blades can be described as an oblique harmonic gust. The same approach is used to model the interaction between the rotor wake and the stator vanes in a fan system. Theories of blade gust interaction are also used to model the blade aerodynamic response to a turbulent stream. In this case a Fourier decomposition of the impinging vorticity eld is used to determine the blade response to the superposed Fourier gust components. A gustairfoil interaction is predominantly affected by the gust orientation with respect to the airfoil leading edge, and by the characteristic wavelength of the gust. The gust orientation depends on the value of two angles. The rst one is that formed by the vorticity vector and its projection onto the airfoil plane. The second angle is that formed by the projection of the vorticity vector onto the airfoil plane and the airfoil leading edge. The latter angle is usually referred to as skew angle. The effects of the gust skew angle and wavelength on the unsteady pressure eld are described by the model developed by Graham [63], according to which the space of solutions of a gustairfoil interaction problem can be divided into two sub-spaces, depending on the value of a gust parameter. This parameter is related to the free-stream Mach number, the gust wavelength and the skew angle. Each solution subspace can be represented by a model problem whose solution is known. Any solution of a bladevortex interaction problem can be related to the model solution of the corresponding subspace through appropriate similarity rules. The two model solutions are that of an incompressible oblique bladevortex interaction, and that of a two-dimensional compressible problem. The earliest incompressible gustairfoil interaction models where developed in the Twenties, when the aeroelastic phenomena associated with the increased ight speeds became a critical element of aircraft design. The rst gustairfoil interaction theory for incompressible ows was put forward by von Krmn and Sears [181]. This theory, based on the concepts of circulation theory [180], recovers the results predicted by Theodorsen [173] for a at-plate in a small amplitude sinusoidally oscillatory motion, and some results obtained by Kssner [96]. On the basis of von Krmn & Sears [181] unsteady

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

airfoil theory, Sears [154] derived an analytical expression for the unsteady lift induced by a vortical sinusoidal gust convected past a thin airfoil. Later on, Filotas [50] extended Sears analysis to an oblique sinusoidal gust. The mathematical treatment of a linearized gustairfoil interaction problem consists in splitting the velocity eld into the sum of a solenoidal (rotational) part and a potential (irrotational) part. The solenoidal part is a known function of the imposed upstream disturbances and represents a vortical wave decoupled by the steady-state aerodynamic eld. The potential part is an unknown function of both the mean ow and the vortical disturbances. The solenoidal and the potential parts are coupled by the boundary conditions on the airfoil surface. For a compressible ow the potential part of the unsteady velocity eld is governed by a constant coefcient, homogeneous, convective wave equation that reduces to a Laplace equation if the ow is supposed to be incompressible. Theoretical studies of the unsteady aerodynamic eld past an airfoil in a compressible ow were carried out by Possio [136] and by Amiet [4,5]. The rst one obtained an integral equation relating the pressure eld on the surface of a thin airfoil to a sinusoidally uctuating velocity eld around the airfoil. The second one proposed analytical procedures for the gustairfoil interaction problem in the high- and low-frequency limit. The time-average ow about a real airfoil with nite thickness, camber and angle of attack is no longer a parallel uniform ow. Goldstein and Atassi [62] developed a second order theory for the gustairfoil interaction problem that accounts for the dependence of the unsteady velocity eld on the mean potential ow around the airfoil. They showed that the vortical waves are distorted by the mean ow around the airfoil and that the distortion affects signicantly both the amplitude and the phase of the unsteady velocity eld. In the case of a thin airfoil with small angle of attack and camber, Goldstein and Atassis second order theory provides explicit analytical formulas for the unsteady lift induced by longitudinal and transverse gusts. A typical aeroacoustic approach takes advantage of the acoustic analogy theory in order to relate the acoustic far eld to the pressure distribution on the airfoil surface. Widnall [184] investigated the sound generated by the interaction between a two-dimensional airfoil and an obliquely incident vortex. The airfoil was assumed to be chordwise compact, but not necessarily spanwise compact. The vortex was supposed to remain stationary as it was convected past the airfoil. The velocity eld induced by the vortex was decomposed into Fourier components. These components were introduced into Filotas [50] airfoil response function in order to determine the uctuating pressure eld on the airfoil surface. Finally, the sound generated by the bladevortex interaction was determined by using the pressure eld induced on the airfoil surface. Amiet [2] described the acoustic eld generated by an airfoil immersed in a turbulent ow. He related the spectral behavior of the far pressure eld to the spectral properties of the incident turbulent ow. In the case of an airfoil in a low Mach number stream, the Sears function was used as airfoil response function. Amiet [4] demonstrated that a generalized PrandtlGlauert transformation can be used to reduce a small perturbation problem in a com-

pressible stream into a standard wave equation in a medium at rest. Furthermore, in the low-frequency limit, the secondorder time derivative in the transformed wave equation can be neglected leading to a Laplace equation. The solution of this Laplace equation can be matched to an outer compressible solution that provides the acoustic far eld. Amiet [5] proposed an analytical procedure to calculate the unsteady lift induced by a compressible high-frequency gust on a thin airfoil. The method was based on the assumption that, as shown by Landahl [97], the leading edge and the trailing-edge aerodynamic problems, at high frequencies, can be separately solved and matched in a converging iterative scheme. Amiet solved the leading edge and the trailing-edge problem in terms of Schwartzchild solution up to a second-order matched solution. Martinez and Widnall [108] used the rst two terms in the series of Adamczyks [1] iteration scheme in order to predict the pressure eld induced by an oblique high-frequency compressible gust on a rectangular thin blade. A three-dimensional surface pressure distribution was recovered by means of a spanwise Fourier superposition of two-dimensional solutions. Martinez and Widnall [109] extended their previous formulation to the case of a rotating blade encountering an oblique highfrequency gust. The pressure eld on the blade surface was build via a spanwise Fourier superposition of two-dimensional solutions, but having linearly increasing magnitudes along the blade span. A different aeroacoustic approach consists in solving the linearized Euler equations by means of singular perturbation methods that provide matched near- and far-eld solutions. Myers and Kerschen [123,124] developed analytical models for the compressible, high-frequency interaction between a gust convected by a subsonic mean ow, and a cambered airfoil at non-zero incidence. The mean ow distortion effects due to the airfoil angle of attack and camber where highlighted. Later on Evers and Peake [42] extended the formulation to a transonic mean ow. 3.1.2. Gustcascade interaction models The singular perturbation solutions of high-frequency gust airfoil interaction problems have been also applied to gust cascade interaction problems by Peake and co-workers. Peake and Kerschen [132] and Evers and Peake [43] investigated the sound generated by the interaction between convected vortical and entropic disturbances and a row of blade with small but non-zero camber and thickness, at a non-zero incidence angle. The multiple interactions between the sound and the cascade were included in the model, providing analytical expressions for the forward noise radiation. Analogously to the gustairfoil interaction problem, a signicant sensitivity of the far-eld noise to the ow distortion around each leading edge and to the blade geometry was predicted. However, by integrating the effects of a full incident turbulence spectrum, a quite less sensitive noise level was predicted. Therefore the authors argued that the tonal noise components generated by rotor/stator single harmonic gust interaction is signicantly affected by the blade geometry. Conversely, the broadband part of the noise spectrum is less signicantly affected by the blade geometry.

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

Other gustcascade interaction analyses were carried out by Hanson and Horan [72] and by Hanson [69]. Later on, more exhaustive models were developed by Hanson for a tandem blade row conguration, with the aim of describing the reection/transmission effects and other unsteady coupling mechanisms. 3.1.3. Rotorstator reection/transmission models The rotor viscous wakes and the blade tip vortices that interact with the downstream stator vanes generate acoustic waves that should be transmitted through the rotor blade row before reaching the nacelle inlet. In addition, a portion of the acoustic energy is back reected by the rotor and should be transmitted through the stator vanes before reaching the nacelle exhaust. The occurrence of resonance phenomena due to the mutual acoustic scattering of a cascade tandem system has been investigated analytically by Woodley and Peake [186,187]. They showed that the a tandem cascade system can exhibit two types of resonance: (i) a cut-on/cut-off resonance associated with the gap between the rows, (ii) a resonance of the downstream row driven at low frequencies by a vorticity wave produced by trapped duct modes in the upstream row, and at higher frequencies by radiation modes between the blade rows. The mode trapping and rotor/stator unsteady coupling of the pressure and vortical waves have been also addressed by Hanson in two works. Firstly, the author investigated the tonal rotor/stator interaction in two-dimensional coupled cascades [71]. Later on, the author addressed the problem of broadband noise generation in a coupled rotor/stator system [70], by making use of the harmonic cascade theory put forward by Glegg [58]. The main conclusion of this latter analysis was that the unsteady coupling between the rotor and the stator blades enforces the broadband noise levels. 3.1.4. Nacelle noise transmission models The sound generated by rotor/stator interaction in subsonic rotor tip conditions, as well as by the rotor alone in supersonic rotor tip conditions, propagates through the nacelle duct in the presence of a non-uniform mean ow and acoustic treatment on the duct walls. In the case of innite constant area duct with constant impedance type boundary conditions and a uniform axial ow, the propagation of harmonic acoustic disturbances is governed by the convected Helmholtz equation whose solution, due to the geometry simplicity allowing separation of variables, can be expressed as a superposition of eigensolutions or modes [44, 135]. In the case of a slowly varying circular annular duct with ow and impedance walls, Rienstra [143] extended the modal solution through an elegant application of a perturbative analysis based on the multiple-scales approach. The key points of Rienstras analysis where: the use of an approximate axial mean ow solution, the use of the impedance wall condition obtained by Myers [125] for a non-uniform mean ow grazing a curved surface, and the generalization of the analytical solution to the transition from annular to hollow duct.

Although the multiples-scales approach applied to duct propagation models was already known before, Rienstras analysis represented a real breakthrough in the theory of duct acoustic. Starting from Rienstras work, in fact, several levels of sophistication have been reached in recent years, leading to a predictive analysis approach that in many practical situation is more robust and reliable than a numerical approach. An extension of the multiple-scales modal solution to a generic duct cross section has been presented by Rienstra [144]. This analysis includes the treatment of the cut-on/cut-off transition in a varying area duct. Extensions to sheared and swirling mean ow have been proposed by Cooper and Peake [26,27] and by Vilenski and Rienstra [64]. An additional research trend in the eld of duct acoustics is represented by the analytical modeling of the sound transmission in the presence of a spliced acoustic treatment. Tester et al. [171] have recently developed and veried numerically a model for the sound scattering by liner splices. They have quantied the inuence of the splice number and thickness on the re-distribution of the incident mode energy among scattered duct modes. The technological impact of the models developed by the aforementioned authors is not only due to their high sophistication level. An additional value is represented by the fact that most of these models have been veried through fully numerical solutions, and are presented in a easily reproducible way, as also relevant for the verication of numerical approaches [19, 145,161]. 3.1.5. Nacelle inlet/exhaust noise radiation models Inlet/exhaust nacelle radiation models have been developed in the past for the idealized conguration of a semi-innite unanged circular or annular duct of negligible lip thickness. These models are based on the WienerHopf technique applied at different levels of geometry/ow complexity. Two ows congurations are commonly addressed: the rst is the case of a constant internal/external ow, which is representative of an inlet ow in cruise conditions, the second is the case of two co-axial internal/external ows with discontinuous conditions, which is representative of an exhaust ow in generic ight conditions. However, the latter ow conguration is to be completed with a cylindrical vortex sheet shed from the exhaust lip and convected downstream. This results from the satisfaction of a Kutta condition that permits to remove the singularity of pressure at the lip due to the absence of viscosity in the uid model. For inlet noise radiations, several models have been proposed in the past [79,140,153], starting from the analytical treatment of an unanged duct in the absence of ow put forward by Levine and Schwinger [98]. The main interest of these models lies in the simple relationship between the angular location of the main lobe radiation peak and the cutoff ratio of the radiated duct mode. For exhaust noise radiations, two major contributions are due to Munt [121] and Rienstra [141]. Munt obtained an analytical solution for a discontinuous jet pipe ow supporting a shear layer developing from the pipe lip, whereas Rienstra

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

obtained an analytical solution for an annular exhaust with an innite centerbody and in the presence of a constant ow. These two geometry/ow congurations have been recently merged by Gabard and Astley [53] who obtained solutions for the idealized aero-engine bypass duct radiation problem. More recently, Demir and Rienstra [36] extended the formulation to account for a lined center body and a co-axial ow with discontinuous thermodynamic properties. 3.1.6. Liner impedance models In order to attenuate the fan noise transmitted along the nacelle ducts, the inner walls of aircraft engines are lined with panels of acoustic treatment. The acoustic liner may cover most of the available surface, both in the inlet and exhaust ducts, as resulting from an optimization procedure involving antagonist factors like the installation of anti-icing systems. The key parameter in the liner optimization process is the acoustic impedance. This is comprised of a real part, the resistance, and an imaginary part, the reactance. The rst step in the liner design consists in estimating the values of the liner resistance and reactance that ensure the maximum sound attenuation for a prescribed duct modal content, over the frequency range of interest. The second step consists in selecting a liner class that matches as close as possible the optimal resistance and reactance values for each frequency band of interest. The acoustic treatment used to line an aero-engine walls are of resonator type, consisting in one or two sandwich layers. A single degree of freedom (SDOF) panel is constituted of a porous sheet, a cellular separator such as honeycomb, and a solid backplate. A two degree of freedom (2DOF) panel is constituted of a porous sheet, two layers of honeycomb separated by a porous septum, and a solid backplate. Both SDOF and 2DOF liners are effective over narrow frequency ranges and must be tuned on one or two fan tones, respectively. Typically, the acoustic properties of this class of treatment do not depend on the amplitude of the incident acoustic wave (linear behavior) up to a high value of the incident sound pressure level at which the liner resistance starts exhibiting a dependence on the incident wave amplitude (nonlinear behavior). The rst recognition of this dual behavior is due to Melling [110] who argued that, in the linear regime, the micro ow in the orice is laminar and the dissipative (resistive) losses may be of Poiseuille type or Helmholtz type. In both cases the losses are due to viscous dissipation in the shear layer. This hypothesis have been partially conrmed through recent direct numerical simulations by Tam and Kurbatskii [166] who observed, in the linear regime, a jetlike ow close to the orice openings and a strongly oscillatory boundary layer. In the nonlinear regime, Melling argued that a turbulent jet takes place at the mouth of the resonator and the primary dissipation mechanism is turbulence. This mechanism was not conrmed by the numerical analysis by Tam and Kurbatskii who observed a vortex-shedding mechanism taking place at certain acoustic frequency bands, which is responsible for the conversion of acoustic energy into kinetic energy and further viscous dissipation into heat. Several analytical models describing the behavior of a liner have been developed in the past. A good review on the topic

is due to Zorumski and Tester [189] and, more recently, to Motsinger and Kraft [120]. At the present time, however, the bast practice in vogue is to predict the liner impedance by means of semi-empirical methods obtained by regression of experimental data. The recent works by Kraft et al. [95], Jones [88] and Hersh and Walker [78] provide examples of the semi-empirical approach. 3.2. Technologies for control and reduction 3.2.1. Fan assembly design The best practice in reducing the fan rotor/stator interaction noise consists in using the so-called Tyler and Sofrin selection rule [175]. This is based on the identication of suitable rotor blades and stator vane numbers, such that the resulting interaction tones are cut-off and therefore do not propagate. The model is based on an elegant manipulation of Fourier series for the prediction of the rotor/stator interaction tones, and simple properties of the duct eigenvalues. A second strategy to reduce the fan interaction noise consists in using the theoretical arguments arising from the gust cascade interaction models in order to reduce the unsteady force induced by the rotor wake on the stator vanes. A simple rule consists in exploiting the property according to which the unsteady lift on the blade decreases as the reduced frequency of the impinging gust increases. A simple way to reduce the reduced frequency is to increase the chord of the stator vanes [37]. Another method used to mitigate the unsteady force on the stator vanes consists in increasing the rotor/stator distance, but this causes performance degradations and weight increase. A quite best practice consists in designing swept and leaned stator vanes [41,188] so that the interaction between the rotor wake and the stator leading edge does not occur simultaneously along the whole span. In addition, the swept stator tip results in an increased rotor/stator distance, without any weight increase. A third strategy, which is indeed an extension of the previous one, consists in using the theoretical arguments arising from the rotorstator reection/transmission models in order to design a fan assembly in which acoustic coupling phenomena can be exploited in order to generate intrinsic antagonist wave phenomena. To the authors knowledge, no applicative work has been published on this topic. 3.2.2. Fan trailing-edge blowing The interaction between the viscous wakes of the rotor blades and the stator vanes is a primary source of rotorstator interaction noise. This can be reduced by lling the wake by blowing air into the wake from a slot in the rotor blade trailingedge. A proof of this concept has been put forward by Sutliff et al. [160] who observed that, by using a blowing rate of 1.8% of the fan mass ow rate, a signicant reduction (from 8 to 20 dB) of the tonal peaks at the rst three harmonics of the blade passage frequency can be attained both in the inlet and exhaust ducts.

10

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

3.2.3. Acoustic treatment As previously stated, a good review on the subject is that by Motsinger and Kraft [120]. Recent researches are mostly concerned with: (i) the use of zero-splice liners that enable full attenuation thanks to absence of acoustic mode scattering [56]; (ii) hybrid passive/active acoustic treatments [55]; (iii) the use of integrated anti-icing/liner systems at the nacelle lip [131]. 3.2.4. Active control Because of fuel efciency factors, the bypass ratio of civil aircraft engines is expected to be increased from less than 10 to about 15 in a near future. A drawback of ultra high bypass ratio aero-engines is the reduced efciency of the passive acoustic treatment. On one side, the increased rotor diameter results in lower blade passage tones, which are more difcult to be absorbed by means of a passive liner. On the other side, the reduced nacelle diameter-to-length ratio results in relatively shorter liners, with a consequent reduced absorbing efciency. These considerations have motivated the development and experimental proof of active control devices for fan tonal noise [107,174,185]. A technological assessment of different active noise control strategies is currently undertaken by the aero-engine manufacturers. 3.2.5. Negatively scarfed intake The negatively scarfed intake has the potential to reduce inlet radiated fan noise. On one side, the extended bottom line lip redirects a part of the acoustic energy upwards and away from the ground, as resulting from a combined effect of lip diffraction and mean ow refraction. On the other side, the longer lip allows the inclusion of additional acoustic liner. Experimental proofs of concepts have been carried out in recent years [11, 25,116], showing the acoustic benets of this novel technology. A broadened technological evaluation involving several aspects is currently undertaken by the major aircraft and aero-engine manufacturers. The asymmetric shape of a scarf inlet, in fact, is responsible for an azimuthal ow distortion along the duct [57] that may result in additional fan interaction noise, thus overwhelming the acoustic benets of the scarfed inlet. In addition, there are general factors as weight and other aerodynamic aspects that must be assessed in a multi-disciplinary evaluation process [157,182]. 3.2.6. Bypass exhaust lip serration Experimental campaigns [129] have demonstrated that the presence of serrations (or chevrons) on the nozzles of high bypass ratio turbofan engines yields a reduction in peak turbulence intensity in the bypass/freestream shear layer. This may result in a reduction of the secondary jet noise. Since, serrations are present on both primary (core) and secondary (bypass) nozzles, the effects of the serration on the bypass modal diffraction at the secondary nozzle are not completely clear. 3.3. Research trends Two major research trends in the eld of fan noise reduction and control can be observed. The rst consists in developing

analytical models of growing sophistication level, both for the noise generation, and for the nacelle transmission and external radiation. The verication of these models is supported by the numerical solution of the ow governing equations. The second research trend consists in developing advanced passive/active control techniques through combined experimental and numerical investigations. The estimation of the cost/benet balance for each noise reduction device is the key point of a difcult technological assessment. 4. Jet noise The progressive introduction into service of rst generation turbofan engines during the Fifties and the Sixties contributed to a 20 EPNdB reduction of aircraft noise certication levels. Successive noise improvements due to second generation turbofans were achieved through a progressive increase of the bypass ratio, balanced by a mitigation of the increased fan noise through improved acoustic treatments. In modern high-bypass-ratio turbofan engines, the jet still contributed to about a half of the overall acoustic energy during take-off. For this reason, the jet noise remains the most fervent aeroacoustic research area, since the establishment of this topic in the Fifties. The main difculty in the prediction of jet noise is due to the lack of a general theory describing the turbulent uctuations in a jet ow. For this reason, a rst principle jet noise theory is not yet available. However, the basic mechanisms involved in the sound generation were established from the beginning, thanks to the several implications of Lighthills acoustic analogy [99]. In subsonic jets, the small-scale turbulence is believed to be the dominant source of noise. Even though large-scale coherent turbulent structures and instability waves have been observed in a wide range of Reynolds numbers, these structures are not effective aeroacoustic sources. However, they play a crucial indirect role on the jet noise generation, by enhancing the turbulent mixing and the consequent jet spreading. In supersonic jets, even though a direct empirical evidence is difcult to be achieved, large-scale coherent turbulent structures and instability waves are believed to be very effective aeroacoustic sources [163]. The universal character of the small-scale turbulent uctuations contributes to the success of predictive models based on the acoustic analogy theory. This models describe the mechanisms of mixing noise generation in subsonic jets. The same mechanisms are present in supersonic jets, but in this case other phenomena predominate. The convection of instability waves or coherent turbulent structures at supersonic speed is responsible for the generation of Mach waves. In addition, when shock cells are present in the jet plume, the interaction between shear layer vortical disturbances and the shocks generate a broadband shock associated noise. Finally, upstream propagating shock associated pressure disturbances can excite instability waves in the shear layer through receptivity mechanisms at the nozzle lip. As a consequence, a feedback coupling can occur with tonal noise selection and generation of the so-called screech noise. Analytical models and semi-empirical predictive methods are

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

11

available for each of these jet noise generation mechanisms. These models are briey reviewed in the following subsections. 4.1. Analytical models 4.1.1. Mixing noise From the publication of the rst aeroacoustic theory by Lighthill in 1952 [99] until the earlier Seventies, most of the analytical work on jet noise was focused on Lighthills acoustic analogy. The main result of Lighthills theory was the so-called eighth power law of the jet noise intensity versus the jet velocity. This law was conrmed through extensive experimental campaigns [183]. Further analytical improvements of Lighthills theory were mainly concerned with the extension to moving jets [46], the modeling of the turbulence correlation tensors [47,138] and consequent removal of the Doppler singularity, the analysis of the mean ow convective effects [60,139], and the description of the so-called Mach wave radiation occurring when vortical disturbances are convected at supersonic phase speed [49]. Several reviews on this subject can be found in the literature [30,103]. In the earlier Seventies, Lush [106] observed that the predictions of Lighthill were not uniformly valid. This indicated the necessity of alternative models for sound generation and propagation within the jet plume. Phillips [134] argued that the discrepancies between Lighthills model and the measurements at low observation angles away from the jet axis was due to mean ow refraction effects. Hence he re-arranged the governing mass and momentum equations in the form of a convected wave equation with a specied source term. Although exact in the sense that no assumptions were made in his derivation, the application of Phillips theory led to the undesirable property that some sound propagation terms were included in the source term. Lilley [101] remedied to this by deriving a third order convective wave equation for transversely sheared mean ows in which all the linear propagation terms are removed from the source term. Several reviews on this subject can be found in the literature [61,102]. Tester and Morfey [172] obtained high-frequency solutions of the Lilleys wave equation for a sheared jet and determined semi-empirical solutions of the radiated acoustic eld. This approach is the basis for modern semi-empirical jet noise models described hereafter, which are based on Reynolds-averaged NavierStokes mean ow solutions. A commonly accepted feature of the mixing noise from high-speed jet is that it consists of two components. The rst one is associated with large-scale vortical uctuations in the jet plume and is predominant in the aft quadrant close to the jet axis. The second component is associated with ne-scale turbulence and is predominant in the forward quadrant and at near-normal angles to the jet-axis. Due to the more universal character of small-scale turbulent uctuations, this second mixing noise component can be predicted on the base of statistical turbulence models and acoustic analogy formulations, as discussed in Section 4.1.5. Recently, Tam and Auriault [164], Tam et al. [168,169] and Fedorchenko [45] have formulated alternative theories of mix-

ing noise generation in jet sheared ows, by evoking some predictive discrepancies of Lighthills theory. Their conjectures about the acoustic analogy have stimulated an animate debate among several researchers [119]. 4.1.2. Mach wave radiation Vortical disturbances frozenly convected through a sheared jet ow do not radiate sound until their phase velocity is supersonic. The phase velocity of a compact turbulent eddy is well represented by the eddy convection velocity. The pressure uctuations driven by turbulent structures convected at supersonic phase speed interfere constructively in the direction of the angle between the normal to the Mach cone attached to the eddies and the direction of the convective motion, usually denoted as sonic angle. As described by Ffowcs Williams and Maidanik [49], the acoustic power associated with the Mach wave radiation scales as the cube of the jet velocity. As a consequence, a saturation process occurs when the jet velocity is increased from low subsonic values, at which the acoustic intensity scales as the eight power of the jet velocity, to high supersonic values, at which the acoustic intensity scales as the cube of the jet velocity. 4.1.3. Shock associated noise When a supersonic jet nozzle operates at adapted ambient conditions, no shock cells are generated in the plume. In this case, the only noise generation mechanisms is associated with the small-scale turbulence mixing noise and Mach wave radiation. The mixture of these two mechanisms is commonly referred to as supersonic turbulent mixing noise [162]. Two commonly observed features of this noise generation mechanism are: (i) the high directionality, with downstream predominant radiation between 25 and 45 from the jet ow direction, (ii) a broadband power spectral density spanning over about two frequency orders, with a very smooth peak separating a variation law as the square of the frequency on the left, from a variation law as the inverse of the frequency on the right. When a supersonic jet is not operated at perfectly expanded conditions, a quasi-periodic shock cell pattern appears in the jet plume. The interaction between the vortical disturbances in the jet stream and the shock fronts generates additional noise, referred to as shock associated noise. The main property of this noise component is that its spectral peak frequency decreases away from the downstream ow direction. In addition, the associated acoustic power level exhibits an opposite directivity behavior with respect to the mixing noise. As a consequence, the acoustic radiation in the jet rear arc, at observation angles greatest than 90 , is commonly dominated by the shock associated noise. This behavior has been described by a simple and elegant phased point-source array model by Harper-Bourne and Fisher [74]. They assumed that the source of shock associated noise is a synchronized array of periodic point monopoles located at a distance equal to the length of the shock cells. The phases of adjacent monopoles are assumed to be correlated by the time taken for turbulent eddies to be convected from one point source to the next. The resulting acoustic waves interfere constructively along observation angles which depend on

12

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

the acoustic frequency, according to a behavior analogous to that of a directional antenna. Starting from the basic concept of a phased array of sources, Tam and Tanna [170] developed a predictive analytical model of the broadband shock associated noise. By employing sophisticated mathematical techniques and elegant use of statistical models of vortical instability waves, they obtained a compact semi-empirical formula for the acoustic power spectral density that provides favorable comparisons with experimental data. 4.1.4. Screech tones The noise spectrum of chocked jets is characterized by the presence of discrete harmonics referred to as screech tones. The fundamental tone arises at the same frequency all along the observation arc. This frequency appears to be a lower bound for the central frequency peak of the shock associated noise. This suggested a relationship between screech tones and shock associated noise. In fact, as illustrated by Tam [162], the screech tone frequency corresponds to the phased array frequency at an observation angle of 180 . The shock associated waves propagating in the upstream direction are those enforced by a feedback loop involving instability waves generated in the shear layer close to the nozzle. The occurrence of harmonics of the fundamental screech tone is attributed to nonlinear propagation effects in the jet plume. 4.1.5. RANS-bases semi-empirical models The idea of extracting from a RANS aerodynamic solution the statistical turbulent quantities required by an acoustic analogy model to predict the far-eld noise radiated by a jet is very old. One of the earliest RANS-based mixing-noise model is implemented in the NASA code MGB, from the initials of its inventors Mani, Gliebe and Balsa [12]. The MGB code uses a RANS mean ow solution to dene local length scale, time scale, and source strength parameters for a source model. This is based on a simplied Lighthill quadrupole source term with an approximate high-frequency solution of Lilleys equation for propagation to the far eld. MBG predictions have been heavily used by industry and NASA for the last twenty years and, over time, its accuracy has been increased. More recently, Khavaran and co-workers [90,91] have removed many of the limitations of the method by reformulating both the propagation and noise source models into the jet-noise code MGBK. Methods of this kind are usually referred to as semi-empirical, since the RANS turbulence model itself is semi-empirical. Tam and Auriault [164] have proposed a RANS-based prediction scheme for the ne-scale turbulence mixing noise component of jet noise. In their formulation a modied k turbulence model provides parameters for a semi-empirical spacetime correlation function of turbulent uctuations. The aeroacoustic source term, however, is not based on Lighthills analogy. By analogy with the gas kinetic theory, they postulated a relationship between the turbulence kinetic energy and the uctuating pressure. An adjoint eld was then used to project the near-eld source onto the far eld. Morris and Farassat [119] argued that the Tam and Auriaults model provides better results than previous Lighthills type models, not because of some aw

in the acoustic analogy approach, but thanks to a better modeling of the turbulence statistics. RANS-based methods provide good predictions for generic axi-symmetric or rectangular nozzles, but they are often unable to account for nozzle design changes, such as the addition of chevrons [105]. In addition, they do not perform well at the low- and high-frequency ends of the acoustic spectrum. The fundamental limitation of the RANS-based approaches is in the non-universal character of the turbulent uctuations and to the difculty in computing the space-time velocity correlations by means of RANS approaches. 4.2. Technologies for control and reduction In order to mitigate the noise generation from high-speed jets, several passive and active ow control techniques have been developed and tested in the years. Passive control is accomplished by modications of the nozzle shape (e.g. serration, beveling, tabs, etc.). Active control is accomplished by adding mass or energy to the ow in order to excite ow instabilities or affect the ow through the generation of new ow structures (e.g. streamwise vortices). Active control is further divided into two categories: open-loop and closed-loop. In the open-loop control, actuation take place based on a predetermined law. In the closed-loop control, real-time information from sensors in the ow are used to drive the actuation process. Only open-loop techniques have been successfully applied to jet noise mitigation. The basic concept behind passive control techniques is the enhancement of jet mixing through the generation of streamwise vorticity. The same concept is applied in open-loop control operated with steady mass or energy injections into the ow, or with pulsating injections at frequencies which are much lower than any instability frequency of the ow. Conversely, when active control is operated at frequencies in the range of ow instability frequencies, the noise mitigation mechanism is intimately related to the aeroacoustic efciency of the excited vortical jet mode, which ultimately acts as an attractor of turbulent kinetic energy. The most popular passive and open-loop active ow control techniques for jet noise mitigation are overlooked in the following sections. 4.2.1. Nozzle tabs and chevrons The rate of mixing of the jet and the surrounding uid can be increased by the presence of small tabs on the nozzle [67, 151]. Unfortunately, the increased mixing, while reducing lowfrequency noise, generates excessive high-frequency noise that may overwhelm any acoustic benet. In addition, tabbed nozzles always result in thrust loss. As an alternative to tabbed nozzles, serrated nozzles edge, or chevrons, have been recently proposed. These devices are the current state of the art in jet-noise mitigation technology for medium- and high-bypass turbofan engines [17,77,129,155]. Analogously to tabs, the triangular serrations in the nozzle trailing-edge induce streamwise vorticity into the shear layer that leads to increased mixing and reduced length of the jet

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

13

plume. However, since the penetration into the ow is lower than that occurring with tabs, the mixing enhancement occurs with a minimal engine performance penalty. Experimental tests show a complex dependence of the noise benets from a series of geometrical parameters, such as the number of chevrons and the level of penetration of the chevrons into the ow. The single and mutual inuence of these parameters on the jet noise generation mechanisms is still unclear. Recently, non-uniform circumferential chevron distributions and shapes have been successfully used in order to exploit the circumferential non-uniformity of the ow in realistic installed engine congurations [111113]. 4.2.2. Nozzle beveling and axial misalignment Beveled primary nozzles have been proposed by Viswanathan [177179] for both single and dual-stream jets, as a way to reduce the jet noise at both subsonic and supersonic operating conditions. Although the mechanisms according which the noise reduction occurs, the beveled noise exhibits strong azimuthal variations of the far-eld noise levels with respect to the axi-symmetric nozzle. Signicant reductions of the noise levels below the longer nozzle lip have been observed. In addition, the acoustic power spectral density is signicantly affected all along the observation arc, revealing an energy re-distribution among the involved noise generation mechanisms. An analogous concept based on the forcing of non-axisymmetric ows has been proposed by Papamoschou [130]. This concept consists in tilting downward, by a few degrees, the secondary ow with respect to the primary ow. The dualstream axial misalignment results in a lower convective Mach number of the turbulent structures in the shear layer. Noise reduction at all frequencies have been documented, manly in the aft radiation arc. 4.2.3. Distributed nozzle The distributed exhaust nozzle concept is based on the idea that the spectral content of jet noise can be dramatically changed by splitting the exhaust plume through an array of smaller jet plumes. Each plume generates a higher frequency noise, resulting in a global noise benet, due to the more effective atmospheric absorption at higher frequencies. In addition, if the distributed nozzle assembly is properly designed, the plume coalescence occurs at lower velocity with respect to the single plume exhaust, thus resulting in lower low-frequency noise. An example of distributed exhaust nozzle is reported by Gaeta et al. [54]. Other congurations are described in the review by Gliebe et al. [59]. 4.2.4. Microjet injection The air, water and aqueous microjet injection is extensively used in the jet noise research area. Water, in particular, is widely used to reduce large pressure uctuations occurring in supersonic jets, as well as the blast wave generated at the ignition of the solid propellant in the rocket motor of a space launcher. Experimental studies clearly show that the water injection signicantly affects the shock cell structures, with consequent benets on the shock associated noise. Several examples of water

microjet injection to supersonic aero-engines jet plumes have been published recently [65,126], showing reductions in both shock associated noise and Mach wave radiation. The use of passive air microjet injection or synthetic jet actuation show less signicant noise benets [66]. Moreover, open-loop control through unsteady rotating microjet injections may even result in increased noise levels [87]. 4.3. Research trends The recent achievements in the numerical simulation of realistic high-speed jets and the associated noise is bringing a new vigor in the investigation of the physical mechanisms involved in the jet noise generation [13]. These simulations, based on the direct solution of ow governing equations ltered at the size of the computational mesh, generate huge data bases enabling statistical analyses and different kinds of noise-ow cross correlations. The research in the jet noise mitigation is mainly focusing on the experimental investigation of the inuence of parameters involved in passive/active ow control devices. In addition new passive/active control devices have been recently introduced. The use of plasma actuators distributed inside the nozzle to excite azimuthal instability modes in the jet is one of the active control technique recently proposed [150]. 5. Summary In this paper we have used a bibliographical approach to describe qualitatively the underlying mechanisms involved in the generation of aerodynamic noise in modern aircraft for civil transportation. Analytical or semi-empirical models of these mechanisms are presented as a key element for the assessment of noise mitigation technologies. After a great deal of theoretical work during the Sixties and the Seventies, and the development of advanced numerical techniques in the last twenty years, the recent trend is to use the numerical solution of the governing equations in order to develop and improve fast predictive tools. Due to the growing diffusion of industrial software that allow to easily integrate different numerical analyses, tradeoff studies and multi-objective/multi-disciplinary optimizations are a common practice during the preliminary design of novel aircraft. This is particularly useful, for instance, to assess the inuence of a noise mitigation device on the aircraft operating cost. A review of the main established technologies for airframe-, fanand jet-noise reduction and those currently under evaluation is also presented. Although many scientic and technological elements have not been addressed, we believe that this work may be useful for a rapid access to bibliographic information in the eld of aircraft noise reduction. References
[1] J.J. Adamczyk, The passage of an innite swept airfoil through an oblique gust, Journal of Aircraft 11 (5) (1974) 281287. [2] R.K. Amiet, Acoustic radiation from an airfoil in a turbulent stream, Journal of Sound and Vibration 41 (4) (1975) 407420.

14

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

[3] R.K. Amiet, Noise due to turbulent ow past a trailing edge, Journal of Sound and Vibration 47 (3) (1976) 387393. [4] R.K. Amiet, Low-frequency approximations in unsteady small perturbation subsonic ows, Journal of Fluid Mechanics 75 (3) (1976) 545552. [5] R.K. Amiet, High frequency thin-airfoil theory for subsonic ow, AIAA Journal 14 (8) (1976) 10761082. [6] R.K. Amiet, Effect of the incident surface pressure eld on noise due to a turbulent ow past a trailing edge, Journal of Sound and Vibration 57 (2) (1978) 305306. [7] D. Angland, X. Zhang, L. Chow, N. Molin, Measurements of ow around a ap side-edge with porous edge treatment, AIAA Paper 2006-213. [8] N.E. Antoine, I.M. Kroo, Aircraft optimization for minimal environmental impact, Journal of Aircraft 41 (4) (2004) 790797. [9] N.E. Antoine, I.M. Kroo, Framework for aircraft conceptual design and environmental performance studies, AIAA Journal 43 (10) (2005) 2100 2109. [10] F.S. Archibald, Unsteady Kutta condition at high values of the reduced frequency parameter, Journal of Aircraft 12 (1975) 545550. [11] N.J. Baker, C.L. Bewick, Noise test of a negatively scarfed inlet are, AIAA Paper 2001-2139. [12] T.F. Balsa, P.R. Gliebe, Aerodynamics and noise of coaxial jets, AIAA Journal 15 (11) (1977) 15501558. [13] J. Berland, C. Bogey, C. Bailly, Large eddy simulation of screech tone generation in a planar underexpanded jet, AIAA Paper 2006-2496. [14] P.J.W. Block, H. Heller, Measurements of fareld sound generation from a ow-excited cavity, NASA TM X-3292. [15] A.J. Bohn, M.D. Shovlin, Upper surface blowing noise of the NASA Ames quiet short-haul research aircraft, Journal of Aircraft 18 (10) (1981) 826832. [16] T.F. Brooks, T.H. Hodgson, Trailing edge noise prediction from measured surface pressures, Journal of Sound and Vibration 78 (1) (1981) 69117. [17] B. Callender, E. Gutmark, Far-eld acoustic investigation into chevron nozzle mechanisms and trends, AIAA Journal 43 (1) (2005) 8795. [18] D. Casalino, M. Jacob, Prediction of aerodynamic sounds from circular rods via spanwise statistical modeling, Journal of Sound and Vibration 262 (4) (2003) 815844. [19] D. Casalino, M. Roger, M. Jacob, Prediction of sound propagation in ducted potential ows using Greens function discretization, AIAA Journal 42 (4) (2004) 736744. [20] K.L. Chandiramani, Diffraction of evanescent waves, with applications to aerodynamically scattered sound and radiation from unblaffed plates, Journal of the Acoustical Society of America 55 (1) (1974) 1929. [21] D.M. Chase, Sound radiated by turbulent ow off a rigid half-plane as obtained from a wavevector spectrum of hydrodynamic pressure, Journal of the Acoustical Society of America 52 (3) (1972) 10111023. [22] D.M. Chase, Noise radiated from an edge in turbulent ow, AIAA Journal 13 (8) (1975) 10411047. [23] L. Chatellier, J. Laumonier, Y. Gervais, Theoretical and experimental investigations of low Mach number turbulent cavity ows, Experiments in Fluids 36 (5) (2004) 728740. [24] L.C. Chow, K. Mau, H. Remy, Landing gears and high lift devices airframe noise research, AIAA Paper 2002-2408. [25] L.R. Clark, R.H. Thomas, R.P. Dougherty, F. Farassat, C.H. Gerhold, Inlet shape effects on the far-eld sound of a model fan, AIAA Paper 1997-1589. [26] A.J. Cooper, N. Peake, Propagation of unsteady disturbances in a slowly varying duct with mean swirling ow, Journal of Fluid Mechanics 445 (2001) 207234. [27] A.J. Cooper, N. Peake, Upstream-radiated rotorstator interaction noise in mean swirling ow, Journal of Fluid Mechanics 532 (2005) 219250. [28] D.G. Crighton, Radiation properties of a semi-innite vortex sheet, Proceedings of the Royal Society of London A 330 (1972) 185193. [29] D.G. Crighton, Radiation from vortex lament motion near a half-plane, Journal of Fluid Mechanics 51 (1972) 357362. [30] D.G. Crighton, Basic principles of aerodynamic noise generation, Prog. Aerospace Sci. 16 (1) (1975) 3196.

[31] D.G. Crighton, Airframe noise, in: H.H. Hubbard (Ed.), Aeroacoustics of Flight Vehicles: Theory and Practice, vol.1: Noise Sources, NASA Langley Research Center, Hampton, VA, 1991, pp. 391447. [32] D.G. Crighton, F.G. Leppington, Scattering of aerodynamic noise by a semi-innite compliant plate, Journal of Fluid Mechanics 43 (1970) 721 736. [33] D.G. Crighton, F.G. Leppington, On the scattering of aerodynamic noise, Journal of Fluid Mechanics 46 (1971) 577597. [34] D.G. Crighton, F.G. Leppington, Radiation properties of the semi-innite vortex sheet: The initial value problem, Journal of Fluid Mechanics 64 (1974) 393414. [35] S.S. Davis, Theory of discrete vortex noise, AIAA Journal 13 (1975) 375380. [36] A. Demir, S.W. Rienstra, Sound radiation from an annular duct with jet ow and a lined centerbody, AIAA Paper 2006-2718. [37] J.H. Dittmar, R.P. Woodward, Fan stage redesigned to decrease stator lift uctuation noise, Journal of Aircraft 14 (8) (1977) 746750. [38] W. Dobrzynski, L.C. Chow, P. Guion, D. Shiells, Research into landing gears airframe noise reduction, AIAA Paper 2002-2409. [39] W. Dobrzynski, K. Nagakura, B. Gehlhar, A. Buschbaum, Airframe noise studies on wings with deployed high-lift devices, AIAA Paper 19982337. [40] W.M. Dobrzynski, H.H. Heller, Are wheel-well related aeroacoustic sources of any signicance in airframe noise, AIAA Paper 1977-1270. [41] B. Elhadidi, H.M. Atassi, Passive noise control by blade lean and sweep, AIAA Paper 2006-2999. [42] I. Evers, N. Peake, Noise generation by high-frequency gusts interacting with an airfoil in transonic ow, Journal of Fluid Mechanics 411 (2000) 91130. [43] I. Evers, N. Peake, On sound generation by the interaction between turbulence and a cascade of airfoils with non-uniform mean ow, Journal of Fluid Mechanics 463 (2002) 2552. [44] W. Eversman, Theoretical model for duct acoustic propagation and radiation, in: H.H. Hubbard (Ed.), Aeroacoustics of Flight Vehicles: Theory and Practice, vol. 2: Noise Control, NASA Langley Research Center, Hampton, VA, 1991, pp. 101163. [45] A.T. Fedorchenko, On some fundamental aws in present aeroacoustic theory, Journal of Sound and Vibration 232 (4) (2000) 719782. [46] J.E. Ffowcs Williams, Some thoughts on the effects of aircraft motion and eddy convection on the noise of air jets, Univ. Southampton Aeronaut. Astronaut. Rept. 155. [47] J.E. Ffowcs Williams, The noise from turbulence convected at high speed, Proceeding of The Royal Society of London, Series A: Mathematical and Physical Sciences 255 (1963) 469503. [48] J.E. Ffowcs Williams, L.H. Hall, Aerodynamic sound generation by turbulent ow in the vicinity of a scattering half-plane, Journal of Fluid Mechanics 40 (1970) 657670. [49] J.E. Ffowcs Williams, G. Maidanik, The Mach wave eld radiated by supersonic turbulent shear ows, Journal of Fluid Mechanics 21 (1965) 641657. [50] L.T. Filotas, Theory of airfoil response in a gusty atmosphere part i aerodynamic transfer function, Inst. for Aerospace Studies, Univ. of Toronto, UTIAS Rep. No. 139, ASFOR 69-2150TR. [51] M.R. Fink, Airframe noise prediction method, Federal Aviation Administration FAA-RD-77-29. [52] M.R. Fink, Noise component method for airframe noise, Journal of Aircraft 16 (10) (1979) 659665. [53] G. Gabard, R.J. Astley, Theoretical model for sound radiation from annular jet pipes: Far- and near-eld solutions, Journal of Fluid Mechanics 549 (2006) 315341. [54] R.J. Gaeta, K.K. Ahuja, D.B. Schein, W.D.J. Solomon, Large jet-noise reductions through distributed nozzles, AIAA Paper 2002-2456. [55] M. Galland, B. Mazeaud, N. Sellen, J. Perisse, Performance in wind tunnel of hybrid active/passive absorbent panels, AIAA Paper 2005-2895. [56] F. Gantie, H. Batard, Zero splice intake technology and acoustic benets, AIAA Paper 2006-2455. [57] C.H. Gerhold, L.R. Clark, R.T. Biedron, Control of inow distortion in a scarf inlet, AIAA Paper 2002-2432.

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

15

[58] S.A.L. Glegg, The response of a swept blade row to a three-dimensional gust, Journal of Sound and Vibration 227 (1999) 2964. [59] P.R. Gliebe, J.F. Brausch, R.K. Majjigi, R. Lee, Jet noise suppression, in: H.H. Hubbard (Ed.), Aeroacoustics of Flight Vehicles: Theory and Practice, vol. 2: Noise Control, NASA Langley Research Center, Hampton, VA, 1991, pp. 207269. [60] M. Goldstein, B. Rosenbaum, Effect of anisotropic turbulence on aerodynamic noise (Lighthill theory mathematical model for axisymmetric turbulence), Journal of the Acoustical Society of America 54 (1973) 630 645. [61] M.E. Goldstein, Aeroacoustics, McGraw-Hill, New York, 1976. [62] M.E. Goldstein, H.M. Atassi, A complete second-order theory for the unsteady ow about an airfoil due to periodic gust, Journal of Fluid Mechanics 74 (4) (1976) 741765. [63] J.M.R. Graham, Similarity rules for thin airfoils in non-stationary subsonic ows, Journal of Fluid Mechanics 43 (4) (1970) 753766. [64] G.G. Vilenski, S.W. Rienstra, On hydrodynamic and acoustic modes in a ducted shear ow with wall lining, Journal of Fluid Mechanics 583 (4) (2007) 4570. [65] B. Greska, A. Krothapalli, Jet noise reduction using aqueous microjet injection, AIAA Paper 2004-2971. [66] B. Greska, A. Krothapalli, N. Burnside, W.C. Horne, High-speed jet noise reduction using microjets on a jet engine, AIAA Paper 2004-2969. [67] C.E. Grosch, J.M. Seiner, M.Y. Hussaini, T.L. Jackson, Numerical simulation of mixing enhancement in a hot supersonic jet, Physics of Fluids 9 (4) (1997) 11251143. [68] Y.P. Guo, K.J. Yamamoto, R.W. Stoker, Component-based empirical model for high-lift system noise prediction, Journal of Aircraft 40 (5) (2003) 914922. [69] D.B. Hanson, Theory for broadband noise of rotor and stator cascades with inhomogeneous inow turbulence including effects of lean and sweep, NASA CR-2001-210762. [70] D.B. Hanson, Broadband noise of fans with unsteady coupling theory to account for rotor and stator reection/transmission effects, NASA CR2001-211136. [71] D.B. Hanson, Mode trapping in coupled 2d cascades aerodynamic and acoustic results, AIAA Paper 1993-4417. [72] D.B. Hanson, K.P. Horan, Turbulence/cascade interaction spectra of inow, cascade response, and noise, AIAA Paper 1998-2319. [73] J.C. Hardin, Noise radiation from the side edges of aps, AIAA Journal 18 (5) (1980) 549552. [74] M. Harper-Bourne, M.J. Fisher, The noise from shocks waves in supersonic jets. noise mechanisms, AGARD-CP-131 (1974) 11-1-11-13. [75] H.H. Heller, D.B. Bliss, The physical mechanism of ow-induced pressure uctuations in cavities and concepts for their suppression, AIAA Paper 1975-491. [76] H.H. Heller, D.G. Holmes, E.E. Covert, Flow induced pressure oscillations in shallow cavities, Journal of Sound and Vibration 18 (4) (1971) 545553. [77] B. Henderson, K. Kinzie, J. Whitmire, A. Abeysinghe, The impact of uidic chevrons on jet noise, AIAA Paper 2005-2888. [78] A.S. Hersh, B.E. Walker, J.W. Celano, Helmholtz resonator impedance model, part 1: Nonlinear behavior, AIAA Journal 41 (5) (2003) 795808. [79] G.F. Homicz, J.A. Lordi, A note on the radiative directivity patterns of duct acoustic modes, Journal of Sound and Vibration 41 (3) (1975) 283 290. [80] S. Hosder, J. Schetz, B. Grossman, W. Mason, Airframe noise modeling appropriate for multidisciplinary design and optimization, AIAA Paper 2004-0698. [81] M.S. Howe, Contributions to the theory of aerodynamic sound, with application to excess jet noise and the theory of the ute, Journal of Fluid Mechanics 71 (1975) 625673. [82] M.S. Howe, The inuence of vortex shedding on the generation of sound by convected turbulence, Journal of Fluid Mechanics 76 (4) (1976) 711 740. [83] M.S. Howe, A review of the theory of trailing-edge noise, Journal of Sound and Vibration 61 (3) (1978) 437465. [84] M.S. Howe, On the generation of side-edge ap noise, Journal of Sound and Vibration 80 (4) (1982) 555573.

[85] M.S. Howe, Edge, cavity and aperture tones at very low Mach numbers, Journal of Fluid Mechanics 330 (1996) 6184. [86] M.S. Howe, Low Strouhal number instabilities of ow over apertures and wall cavities, Journal of Acoustic Society of America 102 (1997) 772 780. [87] M.K. Ibrahim, R. Kunimura, Y. Nakamura, Mixing enhancement of compressible jets by using unsteady microjets as actuators, AIAA Journal 40 (4) (2002) 681688. [88] M.G. Jones, An improved model for parallel-element liner impedance prediction, AIAA Paper 1997-1649. [89] D.S. Jones, Aerodynamic sound due to a source near a half-plane, Journal of the Institute of Mathematics and its Applications 9 (1972) 114122. [90] A. Khavaran, J. Bridges, Modelling of turbulence generated noise in jets, AIAA Paper 2004-2983. [91] A. Khavaran, E.A. Kresja, On the role of anisotropy in turbulent mixing noise, AIAA Paper 98-2289. [92] M.R. Khorrami, M.E. Berkman, M. Choudhari, Unsteady ow computations of a slat with a blunt trailing edge, AIAA Journal 38 (11) (2000) 20502058. [93] L. Koop, K. Ehrenfried, A. Dillmann, Reduction of ap side-edge noise by active ow control, AIAA Paper 2002-2469. [94] L. Koop, K. Ehrenfried, A. Dillmann, Reduction of ap side-edge noise: Passive and active ow control, AIAA Paper 2004-2803. [95] R.E. Kraft, J. Yu, H.W. Kwan, Acoustic treatment impedance models for high frequencies, AIAA Paper 1997-1653. [96] H.G. Kssner, Zusammenfassender bericht ber den instationrem auftrieb von geln, Luftfahrtforsch 13 (1936) 410424. [97] M. Landahl, Unsteady Transonic Flow, Pergamon Press, New York, 1961. [98] H. Levine, J. Schwinger, On the radiation of sound from an unanged circular pipe, Physical Review 73 (1948) 383406. [99] M.J. Lighthill, On sound generated aerodynamically: 1. General theory, Proceeding of The Royal Society of London A A211 (1107) (1952) 564 578. [100] G. Lilley, A quest for quiet commercial passenger transport aircraft for take-off and landing, AIAA Paper 2004-2922. [101] G.M. Lilley, On the noise from jets, AGARD CP-131. [102] G.M. Lilley, Noise from turbulent shear ows, in: H.H. Hubbard (Ed.), Aeroacoustics of Flight Vehicles: Theory and Practice, vol. 1: Noise Sources, NASA Langley Research Center, Hampton, VA, 1991, pp. 291 310. [103] G.M. Lilley, Jet noise classical theory and experiments, in: H.H. Hubbard (Ed.), Aeroacoustics of Flight Vehicles: Theory and Practice, vol. 1: Noise Sources, NASA Langley Research Center, Hampton, VA, 1991, pp. 211290. [104] G.M. Lilley, The prediction of airframe noise and comparison with experiment, Journal of Sound and Vibration 239 (4) (2001) 849859. [105] W. Lord, J. Feng, Physics of jet noise suppression, NASA/CP-2001211152 (2000) 617630. [106] P.A. Lush, Measurements of subsonic jet noise and comparison with theory, Journal of Fluid Mechanics 46 (3) (1971) 477500. [107] R. Maier, J. Zillman, A. Roure, M. Winninger, L. Enghardt, U. Tapken, W. Neise, H. Antoine, E. Bouty, Active control of fan tone noise from aircraft engines, AIAA Paper 2001-2220. [108] R. Martinez, S. Widnall, Unied aerodynamic-acoustic theory for a thin rectangular wing encountering a gust, AIAA Journal 18 (6) (1980) 636 645. [109] R. Martinez, S. Widnall, An aeroacoustic model for high-speed, unsteady bladevortex interaction, AIAA Journal 21 (9) (1983) 12251231. [110] T.H. Melling, The acoustic impedance of perforates at medium and high sound pressure levels, Journal of Sound and Vibration 29 (8) (1973) 1 65. [111] V.G. Mengle, R. Elkoby, L. Brusniak, R.H. Thomas, Reducing propulsion airframe aeroacoustic interactions with uniquely tailored chevrons: 1. Isolated nozzles, AIAA Paper 2006-2467. [112] V.G. Mengle, R. Elkoby, L. Brusniak, R.H. Thomas, Reducing propulsion airframe aeroacoustic interactions with uniquely tailored chevrons: 2. Installed nozzles, AIAA Paper 2006-2434.

16

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

[113] V.G. Mengle, R. Elkoby, L. Brusniak, R.H. Thomas, Reducing propulsion airframe aeroacoustic interactions with uniquely tailored chevrons: 3. Jetap interactions, AIAA Paper 2006-2435. [114] N. Molin, J. Piet, L. Chow, M. Smith, W. Dobrzynski, C. Seror, Prediction of low noise aircraft landing gears and comparison with test results, AIAA Paper 2006-2623. [115] N. Molin, M. Roger, S. Barre, Prediction of aircraft high-lift device noise using dedicated analytical models, AIAA Paper 2003-3225. [116] F. Monttagaud, S. Montoux, Negatively scarfed intake: Design and acoustic performance, AIAA Paper 2005-2944. [117] S. Moreau, M. Roger, Effect of airfoil aerodynamic loading on trailing edge noise sources, AIAA Journal 43 (1) (2004) 4152. [118] S. Moreau, M. Roger, Competing broadband noise mechanisms in lowspeed axial fans, AIAA Journal 45 (1) (2007) 4857. [119] P.J. Morris, F. Farassat, Acoustic analogy and alternative theories for jet noise prediction, AIAA Journal 40 (4) (2002) 671680. [120] R.E. Motsinger, R.E. Kraft, Design and performance of duct acoustic treatment, in: H.F. Nelson (Ed.), Aeroacoustics of Flight Vehicles: Theory and Practice, vol. 2: Noise Control, NASA, Hampton, VA, 1991, pp. 165206. [121] R. Munt, The interaction of sound with a subsonic jet issuing from a semi-innite cylindrical pipe, Journal of Fluid Mechanics 83 (1977) 609 640. [122] M. MvChoudhari, D.P. Lockard, M.G. Macaraeg, B.A. Singer, C.L. Streett, G.R. Neubert, R.W. Stoker, J.R. Underbrink, M.E. Berkman, M.R. Khorrami, S.S. Sadowski, Aeroacoustic experiments in the Langley low-turbulence pressure tunnel, NASA TM-2002-211432. [123] M.J. Myers, E.J. Kerschen, Inuence of camber on sound generation by airfoils interacting with high-frequency gusts, Journal of Fluid Mechanics 353 (1995) 221259. [124] M.J. Myers, E.J. Kerschen, Inuence of incidence angle on sound generation by airfoils interacting with high-frequency gusts, Journal of Fluid Mechanics 292 (1997) 271304. [125] M.K. Myers, On the acoustic boundary condition in the presence of ow, Journal of Sound and Vibration 71 (3) (1980) 429434. [126] T.D. Norum, Reduction in multi-component jet noise by water injection, AIAA Paper 2004-2976. [127] F. Obermeier, Berechnung aerodynamisch erzeugter shallfelder mittels der methode der matched asymptotic expansion, Acustica 18 (1967) 238240. [128] S.A. Orszag, S.C. Crow, Instability of a vortex sheet leaving a semiinnite plate, Studies in Applied Mathematics 49 (1970) 167181. [129] G. Page, J. McGuirk, M. Hossain, N. Hughes, M. Trumper, A computational and experimental investigation of serrated coaxial nozzles, AIAA Paper 2002-2554. [130] D. Papamoschou, New method for jet noise reduction in turbofan engines, AIAA Journal 42 (11) (2004) 22452253. [131] C.A. Parente, Combined acoustic and anti-ice engine inlet liner, US Patent Issued on November 24 by Northrop Grumman Corporation (No. 962863 led on 1997-11-03). [132] N. Peake, E.J. Kerschen, Inuence of mean loading on noise generated by the interaction of gusts with a at-plate cascade: Upstream radiation, Journal of Fluid Mechanics 347 (1997) 315346. [133] O.M. Phillips, The intensity of aeolian tones, Journal of Fluid Mechanics 1 (1956) 607624. [134] O.M. Phillips, On the generation of sound by supersonic turbulent shear layers, Journal of Fluid Mechanics 9 (1960) 128. [135] A.D. Pierce, Acoustics. An Introduction to its Physical Principles and Applications, McGraw-Hill, 1981. [136] C. Possio, Lazione aerodinamica sul prolo oscillante in un uido compressibile a velocit iposonora, LAerotecnica 18 (4) (1938) 441458. [137] J.D. Revell, H.L. Kuntz, F.J. Balena, W.C. Horne, B.L. Storms, R.P. Dougherty, Trailing-edge ap noise reduction by porous acoustic treatment, AIAA Paper 97-1646. [138] H.S. Ribner, The generation of sound by turbulent jets, Advances in Applied Mechanics 8 (1964) 103182. [139] H.S. Ribner, Quadrupole correlations governing the pattern of jet noise, Journal of Fluid Mechanics 38 (1969) 124.

[140] E.J. Rice, M.F. Heidmann, T.G. Sofrin, Modal propagation angles in a cylindrical duct with ow and their relation to sound radiation, AIAA Paper 79-0183. [141] S. Rienstra, Acoustic radiation from a semi-innite annular duct in a uniform subsonic mean ow, Journal of Sound and Vibration 94 (1984) 267288. [142] S.W. Rienstra, Sound diffraction at a trailing edge, Journal of Fluid Mechanics 108 (1981) 443460. [143] S.W. Rienstra, Sound transmission in slowly varying circular and annular lined ducts with ow, Journal of Fluid Mechanics 380 (1999) 279296. [144] S.W. Rienstra, Sound propagation in slowly varying lined ow ducts of arbitrary cross-section, Journal of Fluid Mechanics 495 (2003) 157173. [145] S.W. Rienstra, W. Eversman, A numerical comparison between the multiple-scales and nite-element solution for sound propagation in lined ow ducts, Journal of Fluid Mechanics 437 (2001) 367384. [146] D. Rockwell, E. Naudascher, Review: Self-sustaining oscillations of ow past cavities, J. Fluids Eng. 100 (1978) 152165. [147] M. Roger, S. Moreau, Broadband self-noise from loaded fan blades, AIAA Journal 42 (3) (2004) 536544. [148] M. Roger, S. Moreau, Bach-scattering correction and further extensions of Amiets trailing edge noise model, part 1: Theory, Journal of Sound and Vibration 286 (3) (2005) 477506. [149] J.E. Rossiter, Wind tunnel experiments of the ow over rectangular cavities at subsonic and transonic speeds, Aeronautical Research Council Reports and Memoranda 3438. [150] M. Samimy, J.H. Kim, I. Adamovich, J. Kastner, Toward noise mitigation in high speed and high Reynolds number jets using plasma actuators, AIAA Paper 2006-2703. [151] M. Samimy, K.B.M.Q. Zaman, M.F. Reeder, Effect of tabs on the ow and noise eld of an axisymmetric jet, AIAA Journal 31 (4) (1993) 609 619. [152] B. Satyanarayana, S. Davis, Experimental studies of unsteady trailingedge condition, AIAA Journal 16 (1978) 125129. [153] A.V. Saule, E.J. Rice, Far-eld multimodal acoustic radiation directivity, NASA TM-73839. [154] W.R. Sears, Some aspects of non-stationary airfoil theory and its practical application, Journal of the Aeronautical Sciences 8 (3) (1941) 104108. [155] J. Seiner, L. Ukeiley, B. Jansen, Noise reduction technology for f/a-18 e/f aircraft, AIAA Paper 2004-2972. [156] H. Shen, C.K.W. Tam, Three-dimensional numerical simulation of the jet screech phenomenon, AIAA Journal 40 (1) (2002) 3341. [157] A.R. Smith, R. Thorne, C. Gouttenoire, T. Surply, P. Chanez, Aerodynamic aspects of application of negative scarf intake to high bypass ratio civil turbofans, AIAA Paper 2005-4205. [158] M. Smith, L. Chow, N. Molin, Attenuation of slat trailing edge noise using slat gap acoustic liners, AIAA Paper 2006-2666. [159] B.L. Storms, J.C. Ross, W.C. Horne, J.A. Hayes, R.P. Dougherty, J.R. Underbrink, D.F. Scharpf, P.J. Moriarty, An aeroacoustic study of an unswept wing with a three-dimensional high-lift system, NASA TM1998-112222. [160] D.L. Sutliff, D.L. Tweedt, E.B. Fite, E. Envia, Low-speed fan noise reduction with trailing edge blowing, AIAA Paper 2002-2492. [161] C.K.W. Tam, K. Ju, E.W. Chien, Scattering of acoustic duct modes by axial liner splices, AIAA Paper 2006-2459. [162] C.K.W. Tam, Jet noise generated by large-scale coherent motion, in: H.H. Hubbard (Ed.), Aeroacoustics of Flight Vehicles: Theory and Practice, vol. 1: Noise Sources, NASA Langley Research Center, Hampton, VA, 1991, pp. 311390. [163] C.K.W. Tam, Supersonic jet noise, Annual Review of Fluid Mechanics 27 (1995) 1743. [164] C.K.W. Tam, L. Auriault, Jet mixing noise from ne-scale turbulence, AIAA Journal 37 (2) (1999) 145153. [165] C.K.W. Tam, P.J.W. Block, On the tones and pressure oscillations induced by ow over rectangular cavities, Journal of Fluid Mechanics 89 (1978) 373399. [166] C.K.W. Tam, K.A. Kurbatskii, Microuid dynamics and acoustics of resonant liners, AIAA Journal 38 (8) (2000) 13311339. [167] C.K.W. Tam, N. Pastouchenko, Gap tones, AIAA Journal 39 (8) (2001) 14421448.

D. Casalino et al. / Aerospace Science and Technology 12 (2008) 117

17

[168] C.K.W. Tam, N.N. Pastouchenko, Fine-scale turbulence noise from dualstream jets, AIAA Journal 44 (1) (2006) 90101. [169] C.K.W. Tam, N.N. Pastouchenko, K. Viswanathan, Fine-scale turbulence noise from hot jets, AIAA Journal 43 (8) (2005) 16751683. [170] C.K.W. Tam, H.K. Tanna, Shock associated noise of supersonic jets from convergent-divergent nozzles, Journal of Sound and Vibration 81 (3) (1982) 337358. [171] B.J. Tester, N.J. Baker, A.J. Kempton, M.C. Wright, Validation of an analytical model for scattering by intake liner splices, AIAA Paper 20042906. [172] B.J. Tester, C.L. Morfey, Developments in jet noise modelling theoretical predictions and comparisons with measured data, Journal of Sound and Vibration 46 (1) (1976) 79103. [173] T. Theodorsen, General theory of aerodynamic instability and the mechanism of utter, NACA Advance Restricted Report (also NACA Tech. Rep. No. 496, 1949). [174] R.H. Thomas, R.A. Burdisso, S.A. Lane, Time-averaged active controller for turbofan engine fan noise reduction, Journal of Aircraft 33 (3) (1996) 524531. [175] J.M. Tyler, T.G. Sofrin, Axial ow compressor noise studies, Transactions of the Society of Automotive Engineers 70 (1962) 309332. [176] H.M.M. van der Wal, P. Sijtsma, Flap noise measurements in a closed wind tunnel with a phased array, AIAA Paper 2001-2170. [177] K. Viswanathan, Nozzle shaping for reduction of jet noise from single jets, AIAA Journal 43 (5) (2005) 10081022. [178] K. Viswanathan, Elegant concept for reduction of jet noise from turbofan engines, Journal of Aircraft 43 (3) (2006) 616626.

[179] K. Viswanathan, Noise of dual-stream beveled nozzles at supercritical pressure ratios, Journal of Aircraft 43 (3) (2006) 627638. [180] T. von Krmn, J.M. Burgers, General Aerodynamic Theory, Perfect Fluids. Aerodynamic Theory, vol. II, Durand, Stanford, 1934. [181] T. von Krmn, W.R. Sears, Airfoil theory for non-uniform motion, Journal of the Aeronautical Sciences 5 (10) (1938) 379390. [182] D.S. Weir, B. Bouldin, J.M. Mendoza, Static and ight aeroacoustic evaluations of a scarf inlet, AIAA Paper 2006-2462. [183] R. Westley, G.M. Lilley, An investigation of the noise eld from a small jet and methods for its reduction, Tech. Rep. 53. College of Aeronautics, Craneld, England. [184] S. Widnall, Helicopter noise do to bladevortex interaction, Journal of Acoustic Society of America 51 (1) (1971) 354365. [185] M.J. Wilkinson, P.F. Joseph, Comparison of active control strategies for the active control of buzz-saw tones, AIAA Paper 2004-2850. [186] B.M. Woodley, N. Peake, Resonant acoustic frequencies of a tandem cascade. part 1. Zero relative motion, Journal of Fluid Mechanics 393 (1999) 215240. [187] B.M. Woodley, N. Peake, Resonant acoustic frequencies of a tandem cascade. part 2. Rotating blade rows, Journal of Fluid Mechanics 393 (1999) 241256. [188] R.P. Woodward, D.M. Elliott, C.E. Hughes, J.J. Berton, Benets of swept-and-leaned stators for fan noise reduction, Journal of Aircraft 38 (6) (2001) 11301138. [189] W.E. Zorumski, B.J. Tester, Far-eld multimodal acoustic radiation directivity, NASA TM-X-73951.

Anda mungkin juga menyukai