A homotopy method for determining the eigenvalues of locally or non-locally reacting acoustic liners in flow ducts

Similar documents
inter.noise 2000 The 29th International Congress and Exhibition on Noise Control Engineering August 2000, Nice, FRANCE

THE ACOUSTIC AND INSTABILITY WAVES OF JETS CONFINED INSIDE AN ACOUSTICALLY LINED RECTANGULAR DUCT FANG Q. HU. Department of Mathematics and Statistics

Computational Aeroacoustics

Sound diffraction by the splitter in a turbofan rotor-stator gap swirling flow

Proceedings of Meetings on Acoustics

HELMHOLTZ RESONATORS FOR DAMPING COMBUSTOR THERMOACOUSTICS

Development of the numerical method for calculating sound radiation from a rotating dipole source in an opened thin duct

AEROELASTIC ANALYSIS OF SPHERICAL SHELLS

Numerical solution of hyperbolic heat conduction in thin surface layers

Transmission loss of rectangular silencers using meso-porous and micro-perforated linings

NUMERICAL PREDICTION OF PERFORATED TUBE ACOUSTIC IMPEDANCE

Proceedings of Meetings on Acoustics

Advance in sound absorption and propagation properties of porous metals under harsh environment conditions

Cut-on, cut-off transition of sound in slowly varying flow ducts

The Acoustic Measurement Capabilities in the Acoustic Linear Research Laboratory at University of Cincinnati

On the dynamic stability of a cantilever under tangential follower force according to Timoshenko beam theory

Frequency equation for the in-plane vibration of a clamped circular plate

Existence of super-harmonics in quarter-vehicle system responses with nonlinear inertia hydraulic track mount given sinusoidal force excitation

Acoustical behavior of purely reacting liners

Solvability condition in multi-scale analysis of gyroscopic continua

Newton-homotopy analysis method for nonlinear equations

Sound Propagation in Ducts

This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and

ACOUSTIC CONTROL USING IMPEDANCE CHARACTERISTICS

Temporal and Spectral Quantification of the Crackle Component in Supersonic Jet Noise

arxiv: v1 [physics.class-ph] 22 Sep 2008

FEDSM COMPUTATIONAL AEROACOUSTIC ANALYSIS OF OVEREXPANDED SUPERSONIC JET IMPINGEMENT ON A FLAT PLATE WITH/WITHOUT HOLE

Aeroacoustic and Aerodynamics of Swirling Flows*

Characteristics of forced convection heat transfer in vertical internally finned tube B

Note that W is the skin surface weight density in units of psf. An equivalent graph in terms of metric units is given in Appendix A.

(A Primitive Approach to) Vortex Sound Theory and Application to Vortex Leapfrogging

Approximate step response of a nonlinear hydraulic mount using a simplified linear model

Synchronization of an uncertain unified chaotic system via adaptive control

1 The formation and analysis of optical waveguides

Computers and Electrical Engineering

NUMERICAL VALIDATION OF AN ACOUSTIC IMAGING METHOD FOR DUCT SPINNING MODE WITH IN-DUCT CIRCULAR MICROPHONE ARRAY

Two-Port Indirect Acoustic Impedance eduction in presence of grazing flows

A Transfer Matrix Method for Estimating the Dispersion and Attenuation of Plane Waves in a Standing Wave Tube

Available online at ScienceDirect. Procedia Engineering 90 (2014 )

A Novel FEM Method for Predicting Thermoacoustic Combustion Instability

Initial inverse problem in heat equation with Bessel operator

This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and

Formulas of Acoustics

Generalized theory of annularly bounded helicon waves

Improvements of a parametric model for fan broadband and tonal noise

University of Southampton Research Repository eprints Soton

Eigenmode analysis. for turbomachinery applications

The Corrected Expressions for the Four-Pole Transmission Matrix for a Duct with a Linear Temperature Gradient and an Exponential Temperature Profile

Natural frequency analysis of fluid-conveying pipes in the ADINA system

Sound transmission in a slowly varying lined flow duct

IMPLEMENTING THE FFOWCS WILLIAMS AND HAWKINGS ACOUSTIC ANALOGY IN ANTARES

Semi-analytical computation of acoustic scattering by spheroids and disks

Vibrations of string. Henna Tahvanainen. November 8, ELEC-E5610 Acoustics and the Physics of Sound, Lecture 4

Experimental investigation of perforations interactions effects under high sound pressure levels

Free Vibration Analysis of Uniform Beams with Arbitrary Number of Cracks by using Adomian Decomposition Method

American International Journal of Research in Science, Technology, Engineering & Mathematics

ACCURATE FREE VIBRATION ANALYSIS OF POINT SUPPORTED MINDLIN PLATES BY THE SUPERPOSITION METHOD

Theoretical and numerical investigation of optimal impedance in lined ducts with flow

Distributed piezoelectric modal sensors for circular plates

Sound radiation from the open end of pipes and ducts in the presence of mean flow

DECEMBER 15-18, 1997 ADELAIDE, SOUTH AUSTRALIA

Analysis of Forced Convection Heat Transfer in Microencapsulated Phase Change Material Suspensions

Application of Reconstruction of Variational Iteration Method on the Laminar Flow in a Porous Cylinder with Regressing Walls

A broadband method for liner impedance eduction in the presence of a mean flow

Heriot-Watt University. A pressure-based estimate of synthetic jet velocity Persoons, Tim; O'Donovan, Tadhg. Heriot-Watt University.

An Observer for Phased Microphone Array Signal Processing with Nonlinear Output

Introduction to Acoustics Exercises

Signal Loss. A1 A L[Neper] = ln or L[dB] = 20log 1. Proportional loss of signal amplitude with increasing propagation distance: = α d

THERMOELASTIC ANALYSIS OF THICK-WALLED FINITE-LENGTH CYLINDERS OF FUNCTIONALLY GRADED MATERIALS

Research Article A Note on the Solutions of the Van der Pol and Duffing Equations Using a Linearisation Method

On the interaction of higher duct modes with a perforated liner system with bias flow

EFFECT OF ORIFICE GEOMETRY ON THE NON-LINEAR ACOUSTIC RESISTANCE OF PERFORATED PLATES IN THE TRANSITION REGIME

Maximizing the natural frequency of a beam with an intermediate elastic support

Effects of rotation on the dynamics of a circular cylindrical shell with application to tire vibration $

Semeniuk, B., Göransson, P. (2018) Modelling the Dynamic Viscous and Thermal Dissipation Mechanisms in a Fibrous Porous Material In:

W-8 Inlet In-duct Array Evaluation

A numerical analysis of a model of growth tumor

Higher Orders Instability of a Hollow Jet Endowed with Surface Tension

AEROELASTIC ANALYSIS OF COMBINED CONICAL - CYLINDRICAL SHELLS

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

Unsteady Flow of a Newtonian Fluid in a Contracting and Expanding Pipe

V (r,t) = i ˆ u( x, y,z,t) + ˆ j v( x, y,z,t) + k ˆ w( x, y, z,t)

A technique based on the equivalent source method for measuring the surface impedance and reflection coefficient of a locally reacting material

NATURAL FREQUENCIES OF A HONEYCOMB SANDWICH PLATE Revision F. A diagram of a honeycomb plate cross-section is shown in Figure 1.

Optimization for heat and sound insulation of honeycomb sandwich panel in thermal environments

NONLINEAR ELECTROMAGNETICS MODEL OF AN ASYMMETRICALLY DRIVEN CAPACITIVE DISCHARGE

Using windscreens to improve the efficiency of noise barriers in wind: finite-difference time-domain simulations

PART VIII: ABSORPTIVE SILENCER DESIGN

Nonlinear parabolic equation model for finite-amplitude sound propagation in an inhomogeneous medium over a non-flat, finite-impedance ground surface

Microphone reciprocity calibration: acoustic field in the coupler

Guided convected acoustic wave coupled with a membrane wall used as noise reduction device

Dynamics of four coupled phase-only oscillators

Transactions on Modelling and Simulation vol 10, 1995 WIT Press, ISSN X

Sound Propagation in Porous Media

On the variations of acoustic absorption peak with flow velocity in Micro Perforated Panels at high level of excitation

Sound Propagation in the Nocturnal Boundary Layer. Roger Waxler Carrick Talmadge Xiao Di Kenneth Gilbert

Journal of Computational Physics

Frequency lock-in is caused by coupled-mode flutter

Flow control. Flow Instability (and control) Vortex Instabilities

An energy-balance approach for oscillator parameter identification

Transcription:

Journal of Sound and Vibration 303 (2007) 277 286 JOURNAL OF SOUND AND VIBRATION www.elsevier.com/locate/jsvi A homotopy method for determining the eigenvalues of locally or non-locally reacting acoustic liners in flow ducts Xiaofeng Sun a,, Lin Du a, Vigor Yang b a School of Jet Propulsion, Beijing University of Aeronautics and Astronautics, Beijing, China b Department of Mechanical Engineering, The Pennsylvania State University, PA 16802, USA Received 27 February 2006; received in revised form 8 May 2006; accepted 19 January 2007 Available online 13 March 2007 Abstract This paper presents a unified algorithm for studying the eigenvalue problem of a lined duct using the homotopy method. Results from various validations show that the method developed in this work can provide accurate and reliable numerical solutions for sound-propagation computation. The investigation also indicates that homotopy methods not only overcome the computational difficulties of the existing methods for locally reacting liners, but also give a completely different way to calculate the eigenvalues of non-locally reacting liners, which have recently received considerable attention due to their potential application for future advanced liners. Finally, a model multi-segmented, non-locally reacting liner is employed to study the possibility of controlling sound attenuation through a bias flow. The simulation shows that by adjusting the bias flow of each segment, optimal sound attenuation can theoretically be achieved. r 2007 Elsevier Ltd. All rights reserved. 1. Introduction Acoustic liners are used extensively to suppress sound propagation in ducts. The design of these liners to achieve the most efficient attenuation has been a major subject of expert concern. For this reason, various methods for predicting the attenuation of acoustic waves in a lined duct have been developed over the last several decades. One of these methods is based on solving the eigenvalue equation in order to compute the radial wavenumbers for each duct propagation mode. The solution of eigenvalues is hence critical for this method to work robustly and efficiently. The primary objective of the present work is to make use of homotopy methods to form a unified algorithm, so as to study the eigenvalue problem appearing in both locally and non-locally reacting liners and also achieve insight into the mechanism of the interaction of sound propagation with various sound-absorption structures. Locally reacting liners that permit propagation only in the direction normal to the duct wall usually consist of a perforated sheet or thin layer of porous material followed by a honeycomb array and backed by the impervious wall of the duct. In general, we can describe the wall-boundary condition with the help of the concept of acoustic impedance without directly considering the sound propagation inside the liners. Corresponding author: Tel./fax: +86 10 82317408. E-mail address: sunxf@buaa.edu.cn (X. Sun). 0022-460X/$ - see front matter r 2007 Elsevier Ltd. All rights reserved. doi:10.1016/j.jsv.2007.01.020

278 X. Sun et al. / Journal of Sound and Vibration 303 (2007) 277 286 The eigenvalue equation derived from the impedance condition is usually solved by Eversman s integration method [1,2], in which a nonlinear ordinary differential equation (ODE) is derived by introducing a parameter perturbation to the transcendental algebraic equation. The solution of eigenvalues is thus changed into an initial-valued problem. With this method, one can obtain the eigenvalues by numerically solving the ODE with the hard wall eigenvalues as the initial condition of the soft wall. This method has had a great influence on the study of sound propagation in ducts with locally reacting liners, especially in the acoustic design of aircraft nacelles. There is singularity, however, at the lowest-order mode, the plane wave, in the ODE derived by Eversman s method for both rectangular and circular ducts [3]. Furthermore, some investigations have also shown that there are numerical instabilities or mode jump phenomena under some impedance boundary conditions. The problem has received little attention in recent years, although it remains insufficiently understood. It is noted, however, that the calculation of the eigenvalues still plays an important role in the preliminary stage of duct acoustic design in many fields related to practical engineering applications. Overcoming the singularity problem apparent in the existing method by developing a new approach is also a problem of academic interest. Non-locally reacting liners, also known as bulk-reacting liners, permit propagation in more than one direction. Their applicability to a relatively wide frequency range and ability to control sound attenuation has received considerable attention in both experimental and theoretical investigations [4 10]. Especially for the calculation of the eigenvalues of non-locally reacting liners, most of the work is based on the Newton Raphson scheme. Common problems are the selection of initial values and the convergence of the algorithm. Once again, a method for accurately and effectively solving the eigenvalues still deserves to be further investigated for both locally and non-locally reacting liners. The homotopy method may effectively resolve the problem of computing eigenvalues. The method, as first proposed by Scarf [11] and further developed by Eaves and Scarf [12], offers a powerful means of determining solutions to complex systems of equations [13]. In this paper, we will use the basic concept of the homotopy method to constitute various homotopy equations to calculate the eigenvalues for both locally and non-locally reacting liners. In this framework, the homotopy equation we suggest has no singularity for the lowest mode, i.e., the plane wave mode, and one can flexibly choose homotopy parameters to avoid numerical instability or mode jump problems. In particular, for a non-locally reacting liner, the effect of the liner cannot be described by effective impedance, since acoustic waves in the liner also propagate parallel to the wall. Because of this, it is invalid to directly extend Eversman s method to the non-locally reacting liners. However, with the application of the homotopy concept, it is possible to constitute the homotopy equation for the non-locally reacting liners. It can then be changed into a nonlinear ODE and then solved as an initial problem of the ODE. We present various numerical results for both porous liners and perforated liners, with an emphasis on comparison with existing results. In particular, we compare our numerical results with the recent experimental data by Eldredge and Dowling [10], and show very good agreement. Finally, this work shows that it is convenient to extend the present model to include the effect of multi-segmented non-locally reacting liners, and is suitable for optimizing computation by making use of the existing transfer matrix or mode-matching method presented by Zorumski [14]. Our numerical simulation for multi-segmented liners shows the capability to control the sound attenuation with bias flow. In the following sections, we will first make a brief introduction to the homotopy methods related to our research and then introduce various numerical results by constituting different homotopy equations for both locally and non-locally reacting liners. In the last part, we will suggest a model for studying multi-segmented non-locally reacting liners and then give the relevant numerical results and analysis. 2. Homotopy methods Homotopy, or continuation, methods are globally convergent numerical techniques for solving nonlinear algebraic equations. They have been used extensively for various practical applications [13], such as trajectory generation and DC operating point problems. As mentioned in the introduction, we extend this method to the eigenvalue problem in duct acoustics, which has not been done previously. For this specific objective, it is necessary to make a brief description of homotopy methods, with an emphasis on the relevant homotopy formulation in connection with our eigenvalue algorithm. Suppose that we wish to find a solution

x 2 D R n to FðxÞ ¼0 (1) which we know to exist, where F : R n! R n is a smooth mapping, and D is compact. Without a good approximation of the zero point x, an iterative solution to Eq. (1) will often fail, because poor starting values are likely to be chosen. As a possible remedy, we might first solve an easier system, also having a solution in D GðxÞ ¼0, (2) where G : R n! R n. From these, we can construct the linear homotopy Hðx; tþ ¼tFðxÞþð1 tþgðxþ, (3) where H : R nþ1! R n. For t ¼ 0, Fðx; 0Þ ¼GðxÞ ¼0 is a vector system with known solutions, whereas for t ¼ 1, Hðx; 1Þ ¼FðxÞ ¼0 yields the solution of the original problems. The parameter t is called the homotopy parameter. If the functions t! xðtþ and G are differentiable, then differentiating Eq. (3) with respect to t gives qh qx qx qt þ qh qt ¼ 0; xð0þ ¼a; 0oto1, (4) which is a system of differential equations with the initial condition xð0þ. More importantly, this actually shows that the solution of the algebraic equation presented in Eq. (3) can be transformed into an initial value problem of ordinary equations in the form of qx qt ¼ qh=qt qh=qx. (5) It should be noted that, for different GðxÞ, there will be different integration paths in solving the differential equation (5). However, it can be verified that these paths will never have bifurcations and will never be infinite in length. Thus, in principle, they can be followed from one end to another, or, in the case of loops, from an arbitrary starting point back to the same point [13]. 3. Homotopy solution of the eigenvalue problem for a locally reacting liner The propagation of sound in a duct with uniform flow is governed by the convected wave equation r 2 p 1 D 2 0 p c 2 0 Dt 2 ¼ 0, (6a) where D 0 =Dt ¼ðq=qtÞþu 0 ðq=qzþ and u 0 is the mean flow velocity in the axial direction, r 2 is the Laplacian operator and p is the acoustic pressure. To determine the sound attenuation in the lined duct shown in Fig. 1, the separation of variables can be applied to solve the equation. Furthermore, the application of the displacement continuity condition and momentum equation [1], gives where g ¼ k r =k, and X. Sun et al. / Journal of Sound and Vibration 303 (2007) 277 286 279 FðgÞ ¼g tanðkb gþ ib 0 w 2 ¼ 0, (6b) w ¼ 1 Ma½1 ð1 Ma2 Þg 2 Š 1=2 1 Ma 2, (6c) Fig. 1. The geometry of an acoustically locally reacting liner two-dimensional duct.

280 X. Sun et al. / Journal of Sound and Vibration 303 (2007) 277 286 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi k z =k ¼ð Ma 1 ð1 Ma 2 Þg 2 Þ=ð1 Ma 2 Þ, (6d) and b 0 is the specific admittance ratio, k is wavenumber, Ma is the Mach number of the mean flow in the duct. Besides, z is the duct axial coordinate and r is the transverse coordinate measured from the hard wall. k rn is the transverse wavenumber for the nth mode of propagation and k zn is the corresponding axial wavenumber. In terms of the basic form of linear homotopy shown in Eq. (3), the two different forms of the homotopy equation can be constructed into Hðg; tþ ¼ð1 tþ tanðkb gþþtfðgþ ¼0, Hðg; tþ ¼ð1 tþðg g 0 ÞþtFðgÞ ¼0, (7a) (7b) where g 0 is the eigenvalue of hard wall. In Eq. (7a), GðxÞ ¼tanðkb g 0 Þ, while in Eq. (7b), GðxÞ ¼ðg g 0 Þ, g 0 is the starting value at t ¼ 0. As introduced in Eq. (5), the relevant differential equations can be arrived at dg dt ¼ tanðkb gþþfðgþ ð1 tþkb sec 2 ðkb gþþtðqf=qgþ, dg dt ¼ ðg g 0ÞþFðgÞ ð1 tþþtðqf=qgþ. (8a) (8b) As we mentioned in Section 2, different GðxÞ can be used to form different homotopy equations. Obviously, Eqs. (8a) and (8b) are both nonlinear differential equations with the same initial value. Thus, the solution of the algebraic equation shown in Eqs. (7a) and (7b) has been transformed into an initial value problem of ODE. Naturally, various tools for the solution of ODE can be used to calculate the eigenvalue. On the other hand, it is noted that Eversman [1,2] derived the following ordinary equation by assuming that the eigenvalues are a function of some parameter, t and letting the admittance b 0 ðtþ ¼tb 0f vary from 0ptp1 dg dt ¼ iw 2 b 0f tanðkb gþþkb g sec 2 ðkb gþ2ib 0 ðtþwmag=v, 1=2 (8c) where v ¼ 1 ð1 Ma 2 Þðk z =kþ 2. It is noted that for the case of plane wave, Eq. (8c) will become singular, since the starting value g ¼ 0. Eversman proposed to use small b 0 and g for the first step of integration to avoid this problem. In addition, if we use Eqs. (8a) and (8b), there is no singularity for any starting values, and naturally it is more convenient to obtain more accurate results by numerically integrating the related ordinary equations. Table 1 shows an excellent comparison of the results from Eqs. (8a) and (8b) and those from Ref. [2]. As we have mentioned previously, using different homotopy equations will result in different integration paths, but the homotopy method guarantees that these paths will finally return to the same point. Fig. 2 shows such results using different homotopy equations. Table 1 Comparison of results for ðk r =kþ in a lined two-dimensional duct with uniform flow, Ma ¼ 0:5; kb ¼ 1:0; b 0 ¼ 0:72 i0:42 Mode Eq. (8a) Eq. (8b) Eversman Starting value 1þ 1:141 þ i0:217 1:141 þ i0:217 1:142 þ i0:217 0:0 þ i0:0 2þ 4:157 i0:782 4:157 i0:782 4:158 i0:781 p þ i0:0 3þ 4:313 i1:988 4:313 i1:988 4:306 i1:995 2p þ i0:0 4þ 7:857 i0:479 7:916 i0:619 7:857 i0:478 3p þ i0:0 5þ 11:046 i0:327 11:077 i0:406 11:046 i0:326 4p þ i0:0 1 0:523 þ i0:353 0:523 þ i0:353 0:523 þ i0:353 0:0 þ i0:0 2 2:389 þ i0:312 2:389 þ i0:312 2:388 þ i0:312 p þ i0:0 3 5:244 i0:081 5:244 i0:081 5:243 i0:081 2p þ i0:0 4 8:214 i0:165 8:214 i0:165 8:214 i0:165 3p þ i0:0 5 11:249 i0:170 11:249 i0:170 11:248 i0:170 4p þ i0:0

X. Sun et al. / Journal of Sound and Vibration 303 (2007) 277 286 281 0.5 1 + Im (kr) 0.0-0.5-1.0-1.5 eq. (8a) eq. (8b) Eversman 2 + 4 + 5 + -2.0 3 + 0 2 4 6 8 10 12 Re (kr) Fig. 2. The integration paths formed by using the different homotopy equations, Ma ¼ 0:5; kb ¼ 1:0; b 0 ¼ 0:72 i0:42. 4. Homotopy solution of the eigenvalue problem for a non-locally reacting liner 4.1. Eigenvalue equation for a single-layer non-locally reacting liner Consider in the polar cylindrical coordinate system ðx; r; yþ a duct with rigid walls lined with an isotropic porous medium material. Also, we assume that the effect of the mean flow velocity in the porous medium is negligible, provided the resistivity of the material is not small. With these assumptions [5], the acoustic field in the porous medium is governed by r e qu p qt þrp p þ su p ¼ 0, (9) O r 0 p qt þru p ¼ 0, (10) p p ¼ c 2 e r p, (11) where c is the speed of sound, c e is the effective speed of sound in the porous material, r 0 is density, r e ¼ sr 0 =O is the effective density of the gas in the porous material, O is the porosity of the reacting liner material, s is the resistivity of porous material, and s is a structure factor which is mainly dependent on the structural properties of the material and the internal friction of the gas. Obviously, propagation in the duct with a non-locally reacting liner can be expressed as superposition of acoustic modes of the form p n ¼ A n J m ðarþe iðot k zn zþ ; ror 1, (12a) p n ¼½B n J m ðbrþþc n Y m ðbrþše iðot k zn zþ ; r 1 oror 2, (12b) where a is the eigenvalue of the non-locally reacting liner in the radial direction ðrpr 1 Þ, and b is the eigenvalue in the radial direction ðr 1 prpr 2 Þ. By the application of the displacement continuity condition and momentum equation, the relevant relations can be written as ¼ 0, (13a) r¼r2

282 And X. Sun et al. / Journal of Sound and Vibration 303 (2007) 277 286 1 Ma k 2 z dp k dr ½pŠ r¼r 1þ0 r¼r 1 0 ¼ r¼r1 0 ¼ z dp dr (13b) r¼r1 þ0, Z dp ior 0 dr (13c) r¼r1 þ0. k 2 z ð1 Ma2 Þþ2Makk z ða 2 k 2 Þ¼0, (14a) " k 2 z þ b2 ¼ m 2 ¼ k c # O 1=2 2, (14b) c e z where r z ¼ 0 ðr e iðs=oþþ ; m ¼ o O 1=2, (15) c e z and Z is the specific acoustic impedance ratio. Substituting Eqs. (12) into Eqs. (13) yields FðaÞ ¼ 1 Ma k 2 z J m ðar 1 Þ k aj 0 m ðar 1Þ þ Z 1 ior 0 zb J mðbr 1 ÞY 0 m ðbr 2Þ Y m ðbr 1 ÞJ 0 m ðbr 2Þ J 0 m ðbr 1ÞY 0 m ðbr 2Þ Y 0 m ðbr 1ÞJ 0 m ðbr ¼ 0. (16) 2Þ When there is no flow in the duct, Eq. (16) is the same as Rienstra s equation [6]. However, up to now, it has not been solved numerically. We solve this equation by the following method. First, according to the eigenvalue equation derived as above, the relevant homotopy equation can be written as Hðk r ; tþ ¼ð1 tþk r J 0 m ðk r0rþþtfðk r Þ¼0, (17a) Hðk r ; tþ ¼ð1 tþðk r k r0 ÞþtFðk r Þ¼0. (17b) Then Eqs. (17a) and (17b) can be transformed into two ODEs with the same initial values, which are finally solved by the integration scheme. We have found that it is difficult to obtain reasonable results for a porous material liner without perforated plates by simply using Eqs. (17a) and (17b). To overcome this difficulty, we can assume that the parameter s, O and s are also a function of t, satisfy sð0þ ¼1, Oð0Þ ¼1, sð0þ ¼0 for t ¼ 0 and sð1þ ¼s f, Oð1Þ ¼O f, sð1þ ¼s f for t ¼ 1. Physically, the starting value ðt ¼ 0Þ and the final value ðt ¼ 1Þ correspond to no porous material in the liner and the real boundary condition with porous material, respectively, whereas it is important to set up an appropriate expression of sðtþ, OðtÞ and sðtþ for different liner configurations in order to obtain robust numerical results. In general, these parameters can be considered as linear or exponential functions of the homotopy parameter. As the first example, we consider a single-layer, non-locally reacting liner without a cover plate. In this case, Fig. 3 shows the comparison between the present calculation and Nayfeh s result [4], which was solved using the Newton Raphson Scheme. Good agreement is obtained in terms of the trends of the results. The difference in the magnitude may be attributed to the effect of non-uniform flow considered in Nayfeh s model. The second example considers the liner without porous materials. The sound absorption mechanism originates from the vortex sound interactions [7,15]. With a bias flow Mach number of 0.03, Fig. 4 gives the frequency dependence of the radial eigenvalues and relevant sound attenuation with the frequency for the different modes. 4.2. Eigenvalue equation of double-layer non-locally reacting liners The geometry of a double-layer acoustically non-locally lined circular duct with uniform flow is shown in Fig. 5. a; b 1, and b 2 denote, respectively, the eigenvalues of the radial direction in the two layers. Therefore, the sound propagation in the duct can be described as p n ¼ A n J m ða n rþe iðot k zn zþ ; ror 1,

X. Sun et al. / Journal of Sound and Vibration 303 (2007) 277 286 283 1 Ω = 0.9, s = 1.5 abs (Im(kz)) 0.1 0.01 Ω = 0.8, s = 1.9 Ω = 0.8, s = 1.9 Ω = 0.9, s = 1.5 result of Nayfeh present result 1E-3 0.0 0.1 0.2 0.3 0.4 h Fig. 3. Variation of the attenuation rate with liner thickness, s ¼ 12:92 and o ¼ 4. a Im (kr * ) 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 n = 0 n = 1 n = 2 m = 0 k = 0 -- 18 n = 3 n = 4 0 2 4 6 8 10 12 14 Re (kr * ) b PWL (db) 60 50 40 30 20 10 (0.2) (0.1) (0.0) (0.3) (0.4) 0 0.0 0.5 1.0 1.5 2.0 2.5 f * = fr 1 /c Fig. 4. Effect of frequency variation on the eigenvalues (a) and insertion loss (b), r 1 ¼ 1:0m; h ¼ 0:2; Ma ¼ 0; L ¼ 1:0; a ¼ 3mm; f ¼ 5%; MaðUÞ ¼U=c 0 ¼ 0:03. p n ¼½B 1n J m ðb 1n rþþc 1n Y m ðb 1n rþše iðot k zn zþ ; r 1 oror 2, p n ¼½B 2n J m ðb 2n rþþc 2n Y m ðb 2n rþše iðot k zn zþ ; r 2 oror 3. (18) The boundary conditions of the double-layer non-locally reacting liner are ¼ 0, (19a) r¼r3 r¼r þ 2 ¼ r¼r 2, (19b) r¼r þ 2 ¼ðp 2 þ p 2 ÞZ 2 j r¼r2, (19c)

284 X. Sun et al. / Journal of Sound and Vibration 303 (2007) 277 286 non-locally reaction liner uniform flow Ma r z r1 r2 r3 Fig. 5. A circular duct with double-layer, acoustically non-locally reacting liner circular duct. r¼r þ 1 ¼ r¼r þ 1 1 Ma k z k 2 r¼r 1 ¼ðp 1 þ p 1 ÞZ 1 j r¼r1., (19d) By imposing the corresponding boundary conditions, we can arrive at the following eigenvalue equation for the double-layer non-locally reacting liners where FðaÞ ¼ 1 Z 1 þ 1 Ma k 2 z J m ðar 1 Þ k aj 0 m ðar 1Þ 1 b (19e) B 1 J m ðbr 1 ÞþC 1 Y m ðbr 1 Þ B 1 J 0 m ðbr 1ÞþC 1 Y 0 m ðbr 1Þ, (20) B 1 C 1 ¼ Z 2Y 0 m ðbr 3Þf 1 þ by 0 m ðbr 2Þf 2 Z 2 J 0 m ðbr 3Þf 1 þ bj 0 m ðbr 2Þf 2, (21) f 1 ¼ J m ðbr 2 ÞY 0 m ðbr 2Þ J 0 m ðbr 2ÞY m ðbr 2 Þ, (22a) f 2 ¼ J 0 m ðbr 2ÞY 0 m ðbr 2Þ J 0 m ðbr 2ÞY 0 m ðbr 2Þ, (22b) and Z is effective compliance [7]. Therefore, the homotopy equation can be expressed as Hða; tþ ¼ð1 tþj 0 m ðar 1ÞþtFðaÞ ¼0. (23) The same procedure as that used for a single-layer non-locally reacting liner can be applied to solve the above homotopy equation. In order to verify the computational results from Eq. (20), we have made a comparison with recent experimental data [10]. In this experiment, r 1 ¼ 12:7cm; r 2 ¼ 15:2cm; f 1 ¼ 0:04; f 2 ¼ 0:02; L ¼ 17:8cm; a 1 ¼ 3:3mm; a 2 ¼ 2:7mmandr 3 br 1. In fact, with the increase of r 3, the computation indicates that the absorption changes little. Fig. 6 shows that our model for the double non-locally reacting liners can give good agreement with the existing experimental results. 4.3. Controllability of non-locally reacting liner by the bias flow As we know, if the eigenvalue of the softwall has been calculated, the mode-matching method [14] can be used robustly to predict the attenuation of multi-segmented liners. To investigate the possibility of controlling the sound attenuation of non-locally reacting liner through the bias flow, we consider a three-segmented liner, as shown in Fig. 7. The parameters f and a represent the open area ratio and radius of the liner, respectively.

X. Sun et al. / Journal of Sound and Vibration 303 (2007) 277 286 285 a Absorption 1.0 Ma(U) = 0.023 and Ma = 0 0.8 0.6 0.4 0.2 0.0 0.0 0.5 1.0 1.5 2.0 kl result of Eldredge present result Experiment: 100dB, 115dB b Absorption 1.0 Ma(U) = 0.041 and Ma = 0 0.8 0.6 0.4 0.2 0.0 0.0 0.5 1.0 1.5 2.0 2.5 kl present result Experiment:100dB result of Eldredge Fig. 6. Comparison with other results [10]. M h1 M h2 M h3 uniform flow Ma B 1- B 2- B 3- A - C+ A+ B 1+ B 2+ B 3+ L 1 L 2 L 3 Fig. 7. The geometry of segmented non-locally reacting liners, L 1 ¼ 0:15 m; L 2 ¼ 0:25 m; L 3 ¼ 0:40 m; a ¼ 2mm; r 1 ¼ 1:0m; r 2 ¼ 1:2m; f ¼ 4%: For circumferential mode m ¼ 2, Fig. 8(a) gives the results for sound attenuation with a uniform bias flow for each segment of the liner. We can see that the variation of the reduced frequency kl will correspond to different sound attenuation. Furthermore, if a different combination of the bias flow for each section is used, as shown in Fig. 8(b), we can achieve better sound attenuation, by making use of the sound wave cancelation between the interfaces with different impedance walls. In future work, we may further explore the potential of multi-segmented liners, building on the present and previous work [10,16,17]. 5. Conclusions This paper presents a unified algorithm to study the eigenvalue problem for a lined duct using the homotopy method. The analysis has been carefully benchmarked and validated against the existing work. The homotopy method developed in present study can provides an accurate and reliable means for sound-propagation computation. The approach not only overcomes the computational difficulties of the existing methods for locally reacting liners, but also provide a completely different way to calculate the eigenvalues of non-locally reacting liners. The latter is of great interest due to its potential application for future advanced liners. Finally, a model multi-segmented, non-locally reacting liners is treated to explore the possibility of controlling sound

286 X. Sun et al. / Journal of Sound and Vibration 303 (2007) 277 286 a db 60 50 40 30 20 10 m=2 M h =0.01 M h =0.02 M h =0.03 2.5 3.0 3.5 4.0 4.5 5.0 kl b db 60 50 40 30 20 10 m=2 M h =0.02 M h =0.03 M h1 =0.03 M h2 =0.03 M h2 =0.02 M h1 =0.01 M h2 =0.02 M h2 =0.01 3.0 3.5 4.0 4.5 5.0 kl attenuation through a bias flow. Simulation results indicated that optimal sound attenuation can be achieved by adjusting the bias flow of each segment. Acknowledgement Fig. 8. Effect of the variations of frequency and bias flow on the attenuation. This work was supported by NSFC under Grants 50136010. References [1] W. Eversman, Computation of axial and transverse wave numbers for uniform two-dimensional ducts with flow using a numerical integration scheme, Journal of Sound and Vibration 41 (2) (1975) 252 255. [2] W. Eversman, Initial values for the integration scheme to compute the eigenvalues for propagation in ducts, Journal of Sound and Vibration 50 (1) (1977) 159 162. [3] W. Eversman, Aeroacoustics of flight vehicles, in: H.H. Hubbard (Ed.), Theory and Practice, Vol. 2, Acoustical Society of America, 1994 (Chapter 13). [4] A.H. Nayfeh, J. Sun, D.P. Telonis, Effect of bulk liners on wave propagation in ducts, AIAA Journal 12 (6) (1974) 838 843. [5] A.H. Nayfeh, J.E. Kaiser, D.P. Telionis, Acoustics of aircraft engine-duct systems, AIAA Journal 13 (2) (1975) 130 153. [6] S.W. Rienstra, Contributions to the theory of sound propagation in ducts with bulk-reacting lining, Journal of the Acoustical Society of America 77 (5) (1985) 1681 1685. [7] I.J. Hughes, A.P. Dowling, The absorption of sound by perforated linings, Journal of Fluid Mechanics 218 (1990) 299 335. [8] F.P. Mechel, Modal solutions in circular and annular ducts with locally or bulk reacting lining, Acustica 84 (1998) 201 222. [9] P. Sijtsma, H. van der Wal, Modelling a spiralling type of non-locally reacting liner, AIAA Paper 2003-3308, 2003. [10] J.D. Eldredge, A.P. Dowling, The absorption of axial acoustic waves by a perforated liner with bias flow, Journal of Fluid Mechanics 485 (2003) 307 335. [11] H.E. Scarf, The approximation of fixed points of continuous mappings, SIAM Journal of Applied Mathematics 15 (1967) 1328 1343. [12] B.C. Eaves, H. Scarf, The solution of systems of piecewise linear equations, Mathematics of Operations Research 1 (1976) 1 27. [13] E.L. Allgower, K. Georg, Introduction to Numerical Continuation Methods (Classics in Applied Mathematics), SIAM, Philadelphia, PA, 2003. [14] W.E. Zorumski, Acoustic theory of axisymmetric multisection ducts, NASA TR-R-419, 1974. [15] M.S. Howe, On the theory of unsteady high Reynolds number flow through a circular aperture, Proceedings of the Royal Society of London A 366 (1979) 205 233. [16] X. Sun, X. Jing, H. Zhao, Control of blade flutter by smart casing treatment, Journal of Propulsion and Power 17 (2) (2001) 248 255. [17] H. Zhao, X. Sun, Active control of wall acoustic impedance, AIAA Journal 37 (7) (1999) 825 831.