Skip to main content
Log in

Time-domain simulation of the acoustic nonlinear response of acoustic liners at high sound pressure level

  • Original Paper
  • Published:
Nonlinear Dynamics Aims and scope Submit manuscript

Abstract

In the aeronautical field, numerical modeling of the acoustic response of liners at high sound pressure levels is done using impedance boundary conditions. The numerical studies are restricted to classical locally reacting liners with a known analytical expression of the impedance. In order to model numerically the acoustic response of a wider range of absorbent materials for which no impedance expression is known, this paper focuses on the numerical modeling of the acoustic response of perforated plate liners at high sound pressure levels in the time domain. To do so, a porous-based description of the perforated plate is used to represent the visco-thermal processes occurring inside the perforated plate. This is achieved through the use of the equivalent fluid model (EFM), which contains two irrational transfer functions described herein by a generic model that covers the Johnson–Champoux–Allard–Pride–Lafarge (JCAPL), the JCAL and JCA models. Nonlinear phenomena occurring at high sound pressure levels are taken into account by using Forchheimer’s correction in the time-domain EFM, which introduces a quadratic nonlinearity in the equations. The formulation of the nonlinear EFM equations in the time domain leads to an augmented system for which a proof of stability is given thanks to a Lyapunov functional. An approximate model is built for numerical simulations from the nonlinear EFM using a multipole approximation of the transfer functions. Stability conditions sufficient for the nonlinear multipole-based approximate EFM are provided. A numerical scheme using a discontinuous Galerkin method is developed to validate the model against experiments with perforated plate liners.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Notes

  1. The EFM is also used for inhomogeneous media with physical parameters depending on the space variable [24].

  2. As opposed to the Biot model [25, 26] used, for example, in biomechanics and geophysics [27] for poroelastic materials where the solid phase is in motion.

References

  1. Jones, M.G., Simon, F., Roncen, R.: Broadband and low-frequency acoustic liner investigations at NASA and ONERA. AIAA J. 60(4), 2481–2500 (2022). https://doi.org/10.2514/1.J060862

    Article  ADS  Google Scholar 

  2. Nayfeh, A.H., Tsai, M.-S.: Nonlinear acoustic propagation in two-dimensional ducts. J. Acoust. Soc. Am. 55(6), 1166–1172 (1974)

    Article  ADS  Google Scholar 

  3. Fernando, R., Druon, Y., Coulouvrat, F., Marchiano, R.: Nonlinear waves and shocks in a rigid acoustical guide. J. Acoust. Soc. Am. (2011). https://doi.org/10.1121/1.3531799

    Article  PubMed  Google Scholar 

  4. Peng, F.: Sound absorption of a porous material with a perforated facing at high sound pressure levels. J. Sound Vib. 425, 1–20 (2018). (10.1016/j.jsv.2018.03.028)

    Article  ADS  Google Scholar 

  5. Ingård, U., Labate, S.: Acoustic circulation effects and the nonlinear impedance of orifices. J. Acoust. Soc. Am. 22(2), 211–218 (1950). https://doi.org/10.1121/1.1906591

    Article  ADS  Google Scholar 

  6. Melling, T.H.: The acoustic impendance of perforates at medium and high sound pressure levels. J. Sound Vib. 29(1), 1–65 (1973). https://doi.org/10.1016/S0022-460X(73)80125-7

    Article  ADS  Google Scholar 

  7. Girvin, R.: Aircraft noise-abatement and mitigation strategies. J. Air Transp. Manag. 15(1), 14–22 (2009). https://doi.org/10.1016/j.jairtraman.2008.09.012

    Article  Google Scholar 

  8. Gautam, A., Celik, A., Azarpeyvand, M.: On the acoustic performance of double degree of freedom Helmholtz resonator based acoustic liners. Appl. Acoust. 191, 108661 (2022). https://doi.org/10.1016/j.apacoust.2022.108661

    Article  Google Scholar 

  9. Tayong, R., Dupont, T., Leclaire, P.: Sound absorption of a micro-perforated plate backed by a porous material under high sound excitation: measurement and prediction. Int. J. Eng. Technol. 2(4), 281–292 (2013). https://doi.org/10.14419/ijet.v2i4.1421

    Article  Google Scholar 

  10. Cao, L., Fu, Q., Si, Y., Ding, B., Yu, J.: Porous materials for sound absorption. Compos. Commun. 10, 25–35 (2018). https://doi.org/10.1016/j.coco.2018.05.001

    Article  ADS  Google Scholar 

  11. Atalla, N., Sgard, F.: Modeling of perforated plates and screens using rigid frame porous models. J. Sound Vib. 303(1), 195–208 (2007). https://doi.org/10.1016/j.jsv.2007.01.012

    Article  ADS  Google Scholar 

  12. Tam, C.K.W., Auriault, L.: Time-domain impedance boundary conditions for computational aeroacoustics. AIAA J. 34(5), 917–923 (1996). https://doi.org/10.2514/3.13168

    Article  ADS  Google Scholar 

  13. Fung, K.-Y., Ju, H.: Broadband time-domain impedance models. AIAA J. 39(8), 1449–1454 (2001). https://doi.org/10.2514/2.1495

    Article  ADS  Google Scholar 

  14. Reymen, Y., Baelmans, M., Desmet, W.: Time-domain impedance formulation based on recursive convolution. In: 12th AIAA/CEAS Aeroacoustics Conference (27th AIAA Aeroacoustics Conference). American Institute of Aeronautics and Astronautics, Southampon, UK (2006). https://doi.org/10.2514/6.2006-2685

  15. Li, X.Y., Li, X.D., Tam, C.K.W.: Improved multipole broadband time-domain impedance boundary condition. AIAA J. 50(4), 980–984 (2012). https://doi.org/10.2514/1.J051361

    Article  ADS  Google Scholar 

  16. Alexander, W., Williams, C.: Fundamental DSP Concepts. In: Alexander, W., Williams, C. (eds.) Digital Signal Processing, pp. 19–157. Academic Press, Boston (2017). https://doi.org/10.1016/B978-0-12-804547-3.00002-4

    Chapter  Google Scholar 

  17. Rienstra, S.: Impedance models in time domain, including the extended Helmholtz resonator model. In: 12th AIAA/CEAS Aeroacoustics Conference (27th AIAA Aeroacoustics Conference), p. 2686. American Institute of Aeronautics and Astronautics, Cambridge, MA, USA (2006). https://doi.org/10.2514/6.2006-2686

  18. Dragna, D., Blanc-Benon, P.: Physically admissible impedance models for time-domain computations of outdoor sound propagation. Acta Acust. Acust. 100(3), 401–410 (2014). https://doi.org/10.3813/AAA.918719

    Article  Google Scholar 

  19. Luebbers, R., Hunsberger, F.P., Kunz, K.S., Standler, R.B., Schneider, M.: A frequency-dependent finite-difference time-domain formulation for dispersive materials. IEEE Trans. Electromagn. Compat. 32(3), 222–227 (1990). https://doi.org/10.1109/15.57116

    Article  Google Scholar 

  20. Cockburn, B.: Numerical resolution of Maxwell’s equations in Polarisable media at radio and lower frequencies. SIAM J. Sci. Stat. Comput. 6(4), 843–852 (1985). https://doi.org/10.1137/0906057

    Article  MathSciNet  Google Scholar 

  21. Carcione, J.M., Kosloff, D., Kosloff, R.: Viscoacoustic wave propagation simulation in the earth. Geophysics 53(6), 769–777 (1988). https://doi.org/10.1190/1.1442512

    Article  ADS  Google Scholar 

  22. Dragna, D., Pineau, P., Blanc-Benon, P.: A generalized recursive convolution method for time-domain propagation in porous media. J. Acoust. Soc. Am. 138(2), 1030–1042 (2015). https://doi.org/10.1121/1.4927553

    Article  ADS  PubMed  Google Scholar 

  23. Lafarge, D.: The Equivalent Fluid Model. In: Materials and Acoustics Handbook, pp. 153–204. Wiley, Hoboken (2009). Chap. 6. 10.1002/9780470611609.ch6

  24. Depollier, C., Fellah, Z.E.A., Fellah, M.: Propagation of transient acoustic waves in layered porous media: fractional equations for the scattering operators. Nonlinear Dyn. 38(1), 181–190 (2004). https://doi.org/10.1007/s11071-004-3754-8

    Article  MathSciNet  Google Scholar 

  25. Biot, M.A.: Theory of propagation of elastic waves in a fluid-saturated porous solid. II. Higher frequency range. J. Acoust. Soc. Am. 28(2), 179–191 (1956). https://doi.org/10.1121/1.1908241

    Article  ADS  MathSciNet  Google Scholar 

  26. Biot, M.A.: Theory of propagation of elastic waves in a fluid-saturated porous solid. I. Low-frequency range. J. Acoust. Soc. Am. 28(2), 168–178 (1956). https://doi.org/10.1121/1.1908239

    Article  ADS  MathSciNet  Google Scholar 

  27. Edelman, I.Y.: Asymptotic research of nonlinear wave processes in saturated porous media. Nonlinear Dyn. 13(1), 83–98 (1997). https://doi.org/10.1023/A:1008250024742

    Article  MathSciNet  Google Scholar 

  28. Attenborough, K., Bashir, I., Taherzadeh, S.: Outdoor ground impedance models. J. Acoust. Soc. Am. 129(5), 2806–2819 (2011). https://doi.org/10.1121/1.3569740

    Article  ADS  PubMed  Google Scholar 

  29. Horoshenkov, K.V., Hurrell, A., Groby, J.-P.: A three-parameter analytical model for the acoustical properties of porous media. J. Acoust. Soc. Am. 145(4), 2512–2517 (2019). https://doi.org/10.1121/1.5098778

    Article  ADS  PubMed  Google Scholar 

  30. Horoshenkov, K.V., Hurrell, A., Groby, J.-P.: Erratum: a three-parameter analytical model for the acoustical properties of porous media. [J. Acoust. Soc. Am. 145(4), 2512-2517. J. Acoust. Soc. Am. 147(1), 146–146 (2020). https://doi.org/10.1121/10.0000560

    Article  ADS  PubMed  Google Scholar 

  31. Wilson, D.K.: Relaxation-matched modeling of propagation through porous media, including fractal pore structure. J. Acoust. Soc. Am. 94(2), 1136–1145 (1993). https://doi.org/10.1121/1.406961

    Article  ADS  MathSciNet  Google Scholar 

  32. Wilson, D.K.: Simple, relaxational models for the acoustical properties of porous media. Appl. Acoust. 50(3), 171–188 (1997). https://doi.org/10.1016/S0003-682X(96)00048-5

    Article  Google Scholar 

  33. Champoux, Y., Allard, J.-F.: Dynamic tortuosity and bulk modulus in air-saturated porous media. J. Appl. Phys. 70(4), 1975–1979 (1991). https://doi.org/10.1063/1.349482

    Article  ADS  Google Scholar 

  34. Johnson, D.L., Koplik, J., Dashen, R.: Theory of dynamic permeability and tortuosity in fluid-saturated porous media. J. Fluid Mech. 176, 379–402 (1987). https://doi.org/10.1017/S0022112087000727

    Article  ADS  CAS  Google Scholar 

  35. Lafarge, D.: Propagation du son dans les matériaux poreux à structure rigide saturés par un fluide viscothermique: Définition de paramètres géométriques, analogie électromagnétique, temps de relaxation. PhD thesis, Le Mans (1993). Thèse de doctorat dirigée par Jean-François Allard, Physique, Le Mans

  36. Lafarge, D., Lemarinier, P., Allard, J.F., Tarnow, V.: Dynamic compressibility of air in porous structures at audible frequencies. J. Acoust. Soc. Am. 102(4), 1995–2006 (1997). https://doi.org/10.1121/1.419690

    Article  ADS  Google Scholar 

  37. Pride, S.R., Morgan, F.D., Gangi, A.F.: Drag forces of porous-medium acoustics. Phys. Rev. B 47(9), 4964–4978 (1993). https://doi.org/10.1103/PhysRevB.47.4964

    Article  ADS  CAS  Google Scholar 

  38. Zhao, J., Bao, M., Wang, X., Lee, H., Sakamoto, S.: An equivalent fluid model based finite-difference time-domain algorithm for sound propagation in porous material with rigid frame. J. Acoust. Soc. Am. 143(1), 130–138 (2018). https://doi.org/10.1121/1.5020268

    Article  ADS  PubMed  Google Scholar 

  39. Alomar, A., Dragna, D., Galland, M.-A.: Time-domain simulations of sound propagation in a flow duct with extended-reacting liners. J. Sound Vib. 507, 116137 (2021). https://doi.org/10.1016/j.jsv.2021.116137

    Article  Google Scholar 

  40. Moufid, I., Matignon, D., Roncen, R., Piot, E.: Energy analysis and discretization of the time-domain equivalent fluid model for wave propagation in rigid porous media. J. Comput. Phys. 451, 110888 (2022). https://doi.org/10.1016/j.jcp.2021.110888

    Article  MathSciNet  Google Scholar 

  41. Craster, R.V., Guenneau, S.: Acoustic metamaterials: negative refraction, imaging, lensing and cloaking. Acoust. Metamater. Springer Science & Business Media, Dordrecht ISBN978-94-007-4812-5 Chapter 1 (2013)

  42. Bellis, C., Lombard, B.: Simulating transient wave phenomena in acoustic metamaterials using auxiliary fields. Wave Motion 86, 175–194 (2019). https://doi.org/10.1016/j.wavemoti.2019.01.010

    Article  ADS  MathSciNet  Google Scholar 

  43. Cummings, A.: Transient and multiple frequency sound transmission through perforated plates at high amplitude. J. Acoust. Soc. Am. 79(4), 942–951 (1986). https://doi.org/10.1121/1.393691

    Article  ADS  Google Scholar 

  44. Darcy, H.P.G.: Les Fontaines Publiques de la Ville de Dijon. Exposition et application des principes À Suivre et des Formules À Employer dans les Questions de Distribution D’eau, etc. V. Dalamont, Paris (1856)

  45. McIntosh, J.D., Lambert, R.F.: Nonlinear wave propagation through rigid porous materials. I: Nonlinear parametrization and numerical solutions. J. Acoust. Soc. Am. 88(4), 1939–1949 (1990). https://doi.org/10.1121/1.400217

    Article  ADS  Google Scholar 

  46. Aurégan, Y., Pachebat, M.: Measurement of the nonlinear behavior of acoustical rigid porous materials. Phys. Fluids 11(6), 1342–1345 (1999). https://doi.org/10.1063/1.869999

    Article  ADS  Google Scholar 

  47. Forchheimer, P.: Wasserbewegung durch Boden. Z. Ver. Dtsch. Ing. 45, 1782–1788 (1901)

    Google Scholar 

  48. Umnova, O., Attenborough, K., Standley, E., Cummings, A.: Behavior of rigid-porous layers at high levels of continuous acoustic excitation: theory and experiment. J. Acoust. Soc. Am. 114(3), 1346–1356 (2003). https://doi.org/10.1121/1.1603236

    Article  ADS  CAS  PubMed  Google Scholar 

  49. Wang, X., Peng, F., Chang, B.: Sound absorption of porous metals at high sound pressure levels. J. Acoust. Soc. Am. 126(2), 55–61 (2009). https://doi.org/10.1121/1.3162828

    Article  ADS  Google Scholar 

  50. Laly, Z., Atalla, N., Meslioui, S.-A.: Acoustical modeling of micro-perforated panel at high sound pressure levels using equivalent fluid approach. J. Sound Vib. 427, 134–158 (2018). https://doi.org/10.1016/j.jsv.2017.09.011

    Article  ADS  Google Scholar 

  51. Diab, D., Dragna, D., Salze, E., Galland, M.-A.: Nonlinear broadband time-domain admittance boundary condition for duct acoustics. Application to perforated plate liners. J. Sound Vib. 528, 116892 (2022). https://doi.org/10.1016/j.jsv.2022.116892

    Article  Google Scholar 

  52. Hélie, T., Matignon, D.: Diffusive representations for the analysis and simulation of flared acoustic pipes with visco-thermal losses. Math. Models Methods Appl. Sci. 16(4), 503–536 (2006). https://doi.org/10.1142/S0218202506001248

    Article  MathSciNet  Google Scholar 

  53. Monteghetti, F., Matignon, D., Piot, E., Pascal, L.: Design of broadband time-domain impedance boundary conditions using the oscillatory-diffusive representation of acoustical models. J. Acoust. Soc. Am. 140(3), 1663–1674 (2016). https://doi.org/10.1121/1.4962277

    Article  ADS  PubMed  Google Scholar 

  54. Blanc, E., Chiavassa, G., Lombard, B.: Wave simulation in 2D heterogeneous transversely isotropic porous media with fractional attenuation: a Cartesian grid approach. J. Comput. Phys. 275, 118–142 (2014). https://doi.org/10.1016/j.jcp.2014.07.002

    Article  ADS  MathSciNet  Google Scholar 

  55. Alomar, A., Dragna, D., Galland, M.-A.: Extended-reacting liners in time-domain simulations for broadband attenuation with flow. J. Acoust. Soc. Am. 146(4), 2786–2786 (2019). https://doi.org/10.1121/1.5136651

    Article  ADS  Google Scholar 

  56. Xie, J., Ou, M.-J.Y., Xu, L.: A discontinuous Galerkin method for wave propagation in orthotropic poroelastic media with memory terms. J. Comput. Phys. 397, 108865 (2019). https://doi.org/10.1016/j.jcp.2019.108865

    Article  MathSciNet  CAS  Google Scholar 

  57. Haddar, H., Matignon, D.: Theoretical and numerical analysis of the Webster Lokshin model. [Research Report] RR-6558, INRIA (2008). https://hal.inria.fr/inria-00288254v2/document

  58. Lombard, B., Matignon, D.: Diffusive approximation of a time-fractional Burger’s equation in nonlinear acoustics. SIAM J. Appl. Math. 76(5), 1765–1791 (2016). https://doi.org/10.1137/16M1062491

    Article  MathSciNet  Google Scholar 

  59. Irmay, S.: On the theoretical derivation of Darcy and Forchheimer formulas. EOS Trans. Am. Geophys. Union 39(4), 702–707 (1958). https://doi.org/10.1029/TR039i004p00702

    Article  Google Scholar 

  60. Beavers, G.S., Sparrow, E.M.: Non-Darcy flow through fibrous porous media. J. Appl. Mech. 36(4), 711–714 (1969). https://doi.org/10.1115/1.3564760

    Article  ADS  Google Scholar 

  61. Joseph, D.D., Nield, D.A., Papanicolaou, G.: Nonlinear equation governing flow in a saturated porous medium. Water Resour. Res. 18(4), 1049–1052 (1982). https://doi.org/10.1029/WR018i004p01049

    Article  ADS  Google Scholar 

  62. Kuntz, H.L., Blackstock, D.T.: Attenuation of intense sinusoidal waves in air-saturated, bulk porous materials. J. Acoust. Soc. Am. 81(6), 1723–1731 (1987). https://doi.org/10.1121/1.394787

    Article  ADS  Google Scholar 

  63. Rasoloarijaona, M., Auriault, J.-L.: Nonlinear seepage flow through a rigid porous medium. Eur. J. Mech. B Fluids 13(2), 177–195 (1994)

    MathSciNet  Google Scholar 

  64. Wodie, J.-C., Lévy, T.: Correction non linéaire de la loi de Darcy. Comptes rendus de l’Académie des sciences. Série 2, Mécanique, Physique, Chimie, Sciences de l’univers, Sciences de la Terre 312(3), 157–161 (1991)

  65. Firdaouss, M., Guermond, J.-L., Le Quéré, P.: Nonlinear corrections to Darcy’s law at low Reynolds numbers. J. Fluid Mech. 343, 331–350 (1997). https://doi.org/10.1017/S0022112097005843

    Article  ADS  MathSciNet  Google Scholar 

  66. Avellaneda, M., Torquato, S.: Rigorous link between fluid permeability, electrical conductivity, and relaxation times for transport in porous media. Phys. Fluids A 3(11), 2529–2540 (1991). https://doi.org/10.1063/1.858194

    Article  ADS  MathSciNet  CAS  Google Scholar 

  67. Turo, D., Umnova, O.: Influence of Forchheimer’s nonlinearity and transient effects on pulse propagation in air saturated rigid granular materials. J. Acoust. Soc. Am. 134(6), 4763–4774 (2013). https://doi.org/10.1121/1.4824969

    Article  ADS  PubMed  Google Scholar 

  68. Matignon, D., Prieur, C.: Asymptotic stability of Webster-Lokshin equation. Math. Control Relat. Fields 4, 481–500 (2014). https://doi.org/10.3934/mcrf.2014.4.481

    Article  MathSciNet  Google Scholar 

  69. Gustavsen, B., Semlyen, A.: Rational approximation of frequency domain responses by vector fitting. IEEE Trans. Power Deliv. 14(3), 1052–1061 (1999). https://doi.org/10.1109/61.772353

    Article  Google Scholar 

  70. Gustavsen, B.: Improving the pole relocating properties of vector fitting. IEEE Trans. Power Deliv. 21(3), 1587–1592 (2006). https://doi.org/10.1109/TPWRD.2005.860281

    Article  Google Scholar 

  71. Deschrijver, D., Mrozowski, M., Dhaene, T., De Zutter, D.: Macromodeling of multiport systems using a fast implementation of the vector fitting method. IEEE Microw. Wirel. Compon. Lett. 18(6), 383–385 (2008). https://doi.org/10.1109/LMWC.2008.922585

    Article  Google Scholar 

  72. Monteghetti, F., Matignon, D., Piot, E.: Time-local discretization of fractional and related diffusive operators using Gaussian quadrature with applications. Applied Numerical Mathematics 155, 73–92 (2020) https://doi.org/10.1016/j.apnum.2018.12.003 . Structural Dynamical Systems: Computational Aspects held in Monopoli (Italy) on June 12-15, 2018

  73. Toulorge, T., Desmet, W.: Optimal Runge-Kutta schemes for discontinuous Galerkin space discretizations applied to wave propagation problems. J. Comput. Phys. 231(4), 2067–2091 (2012). https://doi.org/10.1016/j.jcp.2011.11.024

    Article  ADS  MathSciNet  Google Scholar 

  74. Cohen, G., Pernet, S.: Finite Element and Discontinuous Galerkin Methods for Transient Wave Equations. Springer, Dordrecht (2017). https://doi.org/10.1007/978-94-017-7761-2

    Book  Google Scholar 

  75. Howerton, B.M., Vold, H., Jones, M.G.: Application of swept sine excitation for acoustic impedance eduction. In: 25th AIAA/CEAS Aeroacoustics Conference (2019). https://doi.org/10.2514/6.2019-2487

  76. Billard, R.: Study of perforated liners for aeronautics. PhD thesis (2021). Thèse de doctorat dirigée par Gwénaël Gabard et Gilles Tissot, Aéro-acoustique, Le Mans. http://www.theses.fr/2021LEMA1012

  77. Allard, J.F., Atalla, N.: Propagation of Sound in Porous Media. Wiley, Chichester (2009). https://doi.org/10.1002/9780470747339

    Book  Google Scholar 

  78. Motsinger, R.E., Kraft, R.E.: Design and performance of duct acoustic treatment: aeroacoustics of flight vehicles; Chapter 14, Vol. 2: noise control. NASA RP 1258 (1991)

  79. Barree, R.D., Conway, M.W.: Beyond Beta Factors: A complete model for Darcy, Forchheimer, and Trans-Forchheimer flow in porous media. In: SPE Annual Technical Conference and Exhibition. Society of Petroleum Engineers, Houston, TX (2004). https://doi.org/10.2118/89325-ms

  80. Huang, H., Ayoub, J.: Applicability of the Forchheimer equation for Non-Darcy flow in porous media. SPE J. 13(01), 112–122 (2008). https://doi.org/10.2118/102715-PA

    Article  Google Scholar 

  81. Jones, M.G., Watson, W.R., Nark, D.M., Schiller, N.H.: Evaluation of spanwise variable impedance liners with three-dimensional aeroacoustics propagation codes (2017). https://doi.org/10.2514/6.2017-3021

  82. d’Andréa-Novel, B., Boustany, F., Conrad, F., Rao, B.P.: Feedback stabilization of a hybrid PDE-ODE system: application to an overhead crane. Math. Control Signals Syst. 7(1), 1–22 (1994). https://doi.org/10.1007/BF01211483

    Article  MathSciNet  Google Scholar 

  83. Hesthaven, J.S., Warburton, T.: Nodal Discontinuous Galerkin Methods: Algorithms, Analysis, and Applications. Springer, New York, NY (2007). https://doi.org/10.1007/978-0-387-72067-8

    Book  Google Scholar 

  84. Monteghetti, F., Matignon, D., Piot, E.: Energy analysis and discretization of nonlinear impedance boundary conditions for the time-domain linearized Euler equations. J. Comput. Phys. 375, 393–426 (2018). https://doi.org/10.1016/j.jcp.2018.08.037

    Article  ADS  MathSciNet  Google Scholar 

Download references

Funding

This research has been financially supported by ONERA and by ISAE-SUPAERO, through the EUR TSAE under grant ANR-17-EURE-0005.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed to the study conception and design. Material preparation, data collection and analysis were performed by IM and RR. The first draft of the manuscript was written by IM, and all authors commented on previous versions of the manuscript. All authors read and approved the final manuscript

Corresponding author

Correspondence to Rémi Roncen.

Ethics declarations

Conflict of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A Energy analysis

The energy balance of the EFM with the generic model defined in [40] is briefly introduced in this appendix. First, the classical mechanical energy is defined and composed of the kinetic energy and the potential energy:

$$\begin{aligned} E_{\text {m}}(t) := \dfrac{\rho _0\alpha _\infty }{2}\displaystyle \int _{\Omega }\left\Vert {\textbf{u}}\right\Vert ^2\text {d}{\textbf{x}} ~+~ \dfrac{\chi _0}{2} \displaystyle \int _{\Omega }p^2\text {d}{\textbf{x}}\,. \end{aligned}$$
(A1)

Then, an energy is defined for the diffusive variables:

$$\begin{aligned} \begin{aligned} E_{\text {diff}}(t):=&~ \rho _0\,\alpha _\infty \,N\,E_{\varvec{\varphi }}(t)\\&+~\chi _0\,(\gamma -1)\,N'\,E_\psi (t)\\&+~\chi _0\,(\gamma -1)\,E_{\psi _0}(t), \end{aligned} \end{aligned}$$
(A2)

where

$$\begin{aligned}&E_{\varvec{\varphi }}(t) := \dfrac{1}{2}\displaystyle \int _{\Omega }\int _{0}^\infty \mu _{\text {g}}(\xi )\,\xi \,\left\Vert \varvec{\varphi }(\xi ;t,{\textbf{x}})\right\Vert ^2\,\text {d}\xi \,\text {d}{\textbf{x}}, \end{aligned}$$
(A3a)
$$\begin{aligned}&E_{\psi }(t) := \dfrac{1}{2}\displaystyle \int _{\Omega }\int _{L'}^\infty \nu _{\text {g}}(\xi )\,\xi \,|\psi (\xi ;t,{\textbf{x}})|^2\,\text {d}\xi \,\text {d}{\textbf{x}}, \end{aligned}$$
(A3b)
$$\begin{aligned}&E_{\psi _0}(t) := \dfrac{1}{2}\displaystyle \int _{\Omega }r_0\,(-s_0)\,|\psi (-s_0;t,{\textbf{x}})|^2\,\text {d}{\textbf{x}}. \end{aligned}$$
(A3c)

Note that the energy \(E_{\psi _0}\) is similar to \(E_{\psi }\) but for a fixed pole \(\xi =-s_0\) with a positive residue \(r_0\), and all the model parameters (M, N, L, M, \(N'\), \(L'\)) are positive. The energy balance:

$$\begin{aligned}{\mathcal {E}}_{\text {lin}}(t):= \,E_{\text {m}}(t)\, +\, E_{\text {diff}}(t), \end{aligned}$$

is computed based on the definitions (A1) and (A2), and its derivative is found to be

$$\begin{aligned} \dfrac{\text {d}}{\text {d}t}{\mathcal {E}}(t)&= -~\rho _0\,\alpha _\infty \, M\displaystyle \int _{\Omega }\left\Vert {\textbf{u}}\right\Vert ^2\,\text {d}{\textbf{x}}\\&\quad -\rho _0\,\alpha _\infty \, N\displaystyle \int _{\Omega }\int _{L}^\infty \mu _{\text {g}}(\xi )\,\left\Vert \partial _t\varvec{\varphi }(\xi )\right\Vert ^2\,\text {d}\xi \,\text {d}{\textbf{x}}\\&\quad -\chi _0\,(\gamma -1)\, N'\displaystyle \int _{\Omega }\int _{L'}^\infty \nu _{\text {g}}(\xi )\,\left( \partial _t\psi (\xi )\right) ^2\,\text {d}\xi \,\text {d}{\textbf{x}}\\&\quad -\chi _0\,(\gamma -1) \displaystyle \int _{\Omega }r_0\,\left( \partial _t\psi (-s_0)\right) ^2\,\text {d}\xi \,\text {d}{\textbf{x}},\\&\leqslant 0, \end{aligned}$$

when there is no contribution at the boundary \(\partial \Omega \) (i.e., \(p=0\), or \({\textbf{u}}\cdot {\textbf{n}}=0\) on \(\partial \Omega \)) and the functions are defined in the appropriate functional spaces [57]. Regarding the asymptotic stability of this model, it is insufficient to prove that the extended energy \({\mathcal {E}}(t)\) serves as a Lyapunov functional for the dynamical system, i.e., that it decreases along the trajectories in the solution space. The comprehensive proof hinges upon the spectral analysis of the generator of the underlying linear semigroup, as explored in [68]. Notably, this analysis can be directly applied to our linear model.

For the nonlinear EFM, we make use of the same energy functional \({\mathcal {E}}_{\text {nl}} = {\mathcal {E}}_{\text {lin}}\). The only difference appears in the computation of the derivatives, where an additional negative term can be found for \(\frac{\text {d}}{\text {d}t}{\mathcal {E}}_{\text {nl}}\) (see equation (16)). Thus, with a simple comparison principle, the asymptotic stability of the nonlinear EFM can be assessed.

Appendix B Nonlinear Semigroup

To examine the nonlinear aspect of the EFM, we will specifically consider the case where the diffusive parameters N, \(M'\), \(N'\), \(r_0\) are set to 0. For this purpose, we will closely follow the example detailed in [82]. The nonlinear system is as follows:

$$\begin{aligned} \partial _t{\textbf{u}}= & {} -\dfrac{1}{\rho _0\,\alpha _\infty }\varvec{\nabla }p - \Psi ({\textbf{u}} ), \end{aligned}$$
(B1)
$$\begin{aligned} \partial _t p= & {} -\dfrac{1}{\chi _0}\varvec{\nabla }\cdot {\textbf{u}}\, \end{aligned}$$
(B2)

where \(\Psi : {\mathbb {R}}^2 \rightarrow {\mathbb {R}}^2\), \({\textbf{u}} \mapsto \Psi ({\textbf{u}}):=\left\Vert {\textbf{u}}\right\Vert \,{\textbf{u}}\). The previous system can be seen as the abstract evolution equation \(\partial _t X + {{{\mathcal {A}}}}(X) = 0\), with a nonlinear operator \({{{\mathcal {A}}}}\) acting on an appropriate Hilbert space. This operator is maximal monotone.

Maximality means that the operator \(I + {{{\mathcal {A}}}}\) is onto on \({{\mathcal {H}}}\). The careful computation of this property amounts to show that the mapping \(I_2+ \Psi \) is onto on \({\mathbb {R}}^2\), which happens to be the case, since \((1+ \left\Vert {\textbf{u}}\right\Vert )\,{\textbf{u}} = {\textbf{v}}\) \(\Leftrightarrow \) \({\textbf{u}} = \bigg (\frac{1}{2} + \sqrt{\frac{1}{4}+ \left\Vert {\textbf{v}}\right\Vert }\bigg )^{-1}\,{\textbf{v}}\).

Monotonicity in the nonlinear case means that \(({{{\mathcal {A}}}}(X) - {{{\mathcal {A}}}}(Y), X-Y)_{{{\mathcal {H}}}} \ge 0\) for any X, Y in the domain of \({{{\mathcal {A}}}}\). A careful, though classical, computation of this scalar product shows that a sufficient condition is given on the nonlinear term by:

$$\begin{aligned} ({\textbf{u}} - {\textbf{v}}) \cdot (\Psi ({\textbf{u}}) - \Psi ({\textbf{v}}) )\ge 0\, \quad \forall \,{\textbf{u}}, {\textbf{v}} \in {\mathbb {R}}^2. \end{aligned}$$
(B3)

Let us compute \( ({\textbf{u}} - {\textbf{v}}) \cdot (\left\Vert {\textbf{u}}\right\Vert \,{\textbf{u}} - \left\Vert {\textbf{v}}\right\Vert \,{\textbf{v}} ) = \left\Vert {\textbf{u}}\right\Vert ^3 + \left\Vert {\textbf{v}}\right\Vert ^3 - (\left\Vert {\textbf{u}}\right\Vert + \left\Vert {\textbf{v}}\right\Vert )\,{\textbf{u}}\cdot {\textbf{v}} \ge \left\Vert {\textbf{u}}\right\Vert ^3 + \left\Vert {\textbf{v}}\right\Vert ^3 - (\left\Vert {\textbf{u}}\right\Vert + \left\Vert {\textbf{v}}\right\Vert )\,\left\Vert {\textbf{u}}\right\Vert \,\left\Vert {\textbf{v}}\right\Vert = (\left\Vert {\textbf{u}}\right\Vert - \left\Vert {\textbf{v}}\right\Vert )^2\,(\left\Vert {\textbf{u}}\right\Vert +\left\Vert {\textbf{v}}\right\Vert ) \ge 0 \).

Thus, the nonlinear operator \({{{\mathcal {A}}}}\) is maximal monotone and generates a nonlinear semigroup on \({{\mathcal {H}}}\); existence and uniqueness of solutions to (B1)-(B2) follow.

Appendix C General resistivity correction

In Sect. 3, the dependence of the resistivity on the velocity amplitude is taken to be linear, in agreement with the Forchheimer’s correction. However, the resistivity dependence is quadratic at low Reynolds number, leading to

$$\begin{aligned} \sigma = \sigma _0\left( 1+C_f^\star \left\Vert {\textbf{u}}\right\Vert ^2\right) . \end{aligned}$$
(C1)

A theoretical study at low velocity magnitude, where this type of correction is necessary, leads to the same theoretical results given in Theorem 1. It can even be extended to a more general case, where the total resistivity is defined as

$$\begin{aligned} \sigma = \sigma _0\Bigl (1+\Phi \bigl (\left\Vert {\textbf{u}}\right\Vert \bigr )\Bigr ). \end{aligned}$$
(C2)

with \(\Phi :~{\mathbb {R}}^+\rightarrow {\mathbb {R}}^+\) of \(C^1\) (e.g., quadratic at low velocity amplitude and linear at high velocity amplitude as in Fig. 21). Starting from the nonlinear EFM corrected with (C2), the method used to obtain the energy balance variations (16) leads in this case to

$$\begin{aligned} \dfrac{\text {d}}{\text {d}t}{\mathcal {E}}_{\text {nl}}(t)&=~\dfrac{\text {d}}{\text {d}t}{\mathcal {E}}(t)-\rho _0\,\alpha _\infty \, M\displaystyle \int _{\Omega }\Phi \bigl (\left\Vert {\textbf{u}}\right\Vert \bigr )\,\left\Vert {\textbf{u}}\right\Vert ^2\,\text {d}{\textbf{x}}, \end{aligned}$$
(C3)

allowing to conclude on the stability of the more general nonlinear model, since \(\Phi >0\).

Fig. 21
figure 21

Example of a function \(\Phi :~{\mathbb {R}}^+\rightarrow {\mathbb {R}}^+\) of class \(C^1\), having a quadratic behavior for small x and linear for large x

In order to apply the theory of nonlinear semigroups of Appendix B to the more general case when \(\Psi ({\textbf{u}}):=\Phi (\left\Vert {\textbf{u}}\right\Vert )\,{\textbf{u}}\), the two following key ingredients must be checked:

  • Maximality of the nonlinear operator \({{\mathcal {A}}}\) is ensured, as soon as the scalar function \(x \mapsto x + x\,\Phi (x)\) is increasing on \({\mathbb {R}}^+\),

  • Monotonicity of the nonlinear operator \({{\mathcal {A}}}\) is ensured, as soon as the scalar function \(x \mapsto x\,\Phi (x)\) is increasing on \({\mathbb {R}}^+\).

The proof of the second item comes from the inequality \(({\textbf{u}} - {\textbf{v}}) \cdot (\Psi ({\textbf{u}}) - \Psi ({\textbf{v}}) ) \ge (\left\Vert {\textbf{u}}\right\Vert - \left\Vert {\textbf{v}}\right\Vert )\,(\left\Vert {\textbf{u}}\right\Vert \, \Phi (\left\Vert {\textbf{u}}\right\Vert )- \left\Vert {\textbf{v}}\right\Vert \,\Phi (\left\Vert {\textbf{v}}\right\Vert ) \ge 0 \). It is fulfilled by any monomial \(\Phi (x)= x^p\), where \(p>-1\) is not necessarily an integer.

This property can be extended in a straightforward way to positive linear combinations of such nonlinear operators, e.g., with \(\Phi (x) = \phi _0\,x + \phi _1\, x^2\) where \(\phi _0\ge 0\), \(\phi _1\ge 0\).

Appendix D Numerical scheme

The code used in this article is based on the one used in [40] which solves the linearized Euler equations and the linear EFM in 2D, using a discontinuous Galerkin method detailed in Sect. D.1 and a Runge–Kutta method detailed in Sect. D.2

1.1 Space discretization

Given a domain \(\Omega \subset {\mathbb {R}}^2\) for which there is a partition (\({\mathcal {T}}_h\)) where h denotes the maximum diameter of the elements in the partition. The approximation space is taken as \(V_h:=\{v,|\,\forall T\in {\mathcal {T}}_h,v_{|T}\in {\mathbb {P}}^k(T)\}\) where \({\mathbb {P}}^k(T)\) is the space of polynomials of degree at most k. A basis \(\{\lambda _j^i\in {\mathbb {P}}^k(T_i), j=1\}\), with \(d=(k+1)(k+2)/2\), is defined for each element \(T_i\in {\mathcal {T}}_h\). Thus, a scalar function q is approximated on a cell \(T_i\) by

$$\begin{aligned}&q_h(t,{\textbf{x}}) := \displaystyle \sum ^{d}_{j=1}q_h^{i,j}(t)\lambda _j^i({\textbf{x}}). \end{aligned}$$
(D1)

The solution vector \({\textbf{q}}_h(t,{\textbf{x}}) = (u_h~~v_h~~p_h)\) is then defined based on (D1). Applying the DG method [83], it is possible to write:

$$\begin{aligned} \begin{aligned} 0&= \dfrac{\text {d}}{\text {d}t}\int _{T_i}{\textbf{q}}_h(t,{\textbf{x}})\,\lambda _{j}^{i}\,\text {d}\Omega \\&\quad + \int _{T_i}{\textbf{F}}({\textbf{q}}_h(t, {\textbf{x}}))\,\cdot \nabla \lambda _j^i\,\text {d}\Omega \\&\quad - \int _{\partial T_i} \mathbf {F^*}({\textbf{q}}_h^e(t,{\textbf{x}}),{\textbf{q}}_h^i(t,{\textbf{x}}))\cdot \mathbf {n_i},\lambda _j^i\,\text {d}\sigma \\&\quad + \int _{T_i}{\textbf{b}}({\textbf{q}}_h(t,{\textbf{x}}))\,\lambda _j^i\,\text {d}\Omega , \end{aligned} \end{aligned}$$
(D2)

where \({\textbf{F}}(q_h)=(A_x{\textbf{q}}_h,A_y{\textbf{q}}_h)\), \({\textbf{b}}({\textbf{q}}_h)=B{\textbf{q}}_h\), \(\mathbf {n^i}=(n_x^i, n_y^i)\) is the outgoing unit normal vector with respect to the edge \(\partial _tT_i\) and \(\mathbf {F^*}\) is the numerical flux. The solution \({\textbf{q}}_h\) on the edge of a cell \(T_i\) is denoted \({\textbf{q}}_h^i\) or \({\textbf{q}}_h^e\) depending on whether it is, respectively, the internal or the external value of \(T_i\) that is taken into account. In the code, the numerical flux used is the flux vector splitting

$$\begin{aligned}&{\textbf{F}}^*({\textbf{q}}_h^i,{\textbf{q}}_h^e)\, = \, A^+{\textbf{q}}_h^i + A^-{\textbf{q}}_h^e, \end{aligned}$$
(D3)

where the incoming and outgoing flows are separated into two, respectively, associated with \(A^+\) and \(A^-\). The latter are matrices containing, respectively, the positive and negative eigenvalues of \(A=A_xn_x+A_yn_y\). Note that this flow is an exact solution of the one-dimensional Riemann problem with constant coefficients. At the boundary of the domain \(\partial \Omega \), the imposed flux is the centered flux

$$\begin{aligned}&{\textbf{F}}_{\text {BC}}^*({\textbf{q}}_h^i,{\textbf{q}}_h^e)\, := \dfrac{{\textbf{q}}_h^i + {\textbf{q}}_h^e}{2}, \end{aligned}$$
(D4)

where \({\textbf{q}}_h^e\) represents a ghost state defined with \({\textbf{q}}_h^i\). Finally, the semi-discrete equation reads:

$$\begin{aligned}&M\dfrac{\text {d}{\textbf{q}}_h}{\text {d}t}(t) := K\,{\textbf{q}}_h(t) + \widetilde{{\textbf{S}}}(t), \end{aligned}$$
(D5)

with \({\textbf{q}}_h\) the unknown and \(\widetilde{{\textbf{S}}}\) the source term.

1.2 Time discretization

The inversion of the mass matrix M in (D5) is direct thanks to its diagonal structure per block. Indeed, each block is of the order of a few tens of rows, thus small enough to calculate the inverse of each small block by a direct method. Thus, the spatial discretization (D5) can be rewritten

$$\begin{aligned}&\dfrac{\text {d}{\textbf{q}}_h}{\text {d}t}(t) = \mathbf {L_h}(t,{\textbf{q}}_h(t)) = E{\textbf{q}}_h(t) + {\textbf{G}}(t), \end{aligned}$$
(D6)

with \(\mathbf {\mathbf {L_h}}\) the semi-discrete operator, \(E = M^{-1}K\) and \({\textbf{G}}=M^{-1}\widetilde{{\textbf{S}}}\). The Runge–Kutta method used for the spatial discretization is the fourth-order RKF84 method given by [73] and which was notably used for the results given in [84]. This method is shown to be very efficient when used in combination with a DG method for wave propagation problems in the works of [73].

Let \(\{t^n\}^N_{n=0}\) be a partition of \([0,T]\subset {\mathbb {R}}^+\), \(\Delta t = t^{n+1}-t^{n}\) be the time step and \({\textbf{q}}_h^{n}\) be the approximate solution at time \(t^n\). The steps of the RKF84 algorithm are

$$\begin{aligned}&{{\textbf {q}}}^{(0)} = {\textbf{q}}_h^{n}, \end{aligned}$$
(D7a)
$$\begin{aligned}&{\textbf {dq}}^{(i)} = A_i{\textbf {dq}}^{(i-1)} + \mathbf {L_h}\left( t_n+c_i\Delta t,{\textbf{q}}^{(i-1)}\right) \end{aligned}$$
(D7b)
$$\begin{aligned}&{\textbf{q}}^{(i)} = {\textbf{q}}^{(i-1)} + B_i{\textbf {dq}}^{(i)}, \, \text {for } i = 1\dots 8, \end{aligned}$$
(D7c)
$$\begin{aligned}&{\textbf{q}}_h^{n+1} = {{\textbf {q}}}^{(8)}, \end{aligned}$$
(D7d)

where the coefficients \(A_i\), \(B_i\) and \(c_i\) are given by [73, Table A.9].

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Moufid, I., Roncen, R., Matignon, D. et al. Time-domain simulation of the acoustic nonlinear response of acoustic liners at high sound pressure level. Nonlinear Dyn 112, 3133–3162 (2024). https://doi.org/10.1007/s11071-023-09219-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-023-09219-7

Keywords

Navigation