Skip to main content
Log in

Finite-Difference Modeling of Acoustic and Gravity Wave Propagation in Mars Atmosphere: Application to Infrasounds Emitted by Meteor Impacts

  • Published:
Space Science Reviews Aims and scope Submit manuscript

Abstract

The propagation of acoustic and gravity waves in planetary atmospheres is strongly dependent on both wind conditions and attenuation properties. This study presents a finite-difference modeling tool tailored for acoustic-gravity wave applications that takes into account the effect of background winds, attenuation phenomena (including relaxation effects specific to carbon dioxide atmospheres) and wave amplification by exponential density decrease with height. The simulation tool is implemented in 2D Cartesian coordinates and first validated by comparison with analytical solutions for benchmark problems. It is then applied to surface explosions simulating meteor impacts on Mars in various Martian atmospheric conditions inferred from global climate models. The acoustic wave travel times are validated by comparison with 2D ray tracing in a windy atmosphere. Our simulations predict that acoustic waves generated by impacts can refract back to the surface on wind ducts at high altitude. In addition, due to the strong nighttime near-surface temperature gradient on Mars, the acoustic waves are trapped in a waveguide close to the surface, which allows a night-side detection of impacts at large distances in Mars plains. Such theoretical predictions are directly applicable to future measurements by the INSIGHT NASA Discovery mission.

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

Similar content being viewed by others

References

  • M. Abramowitz, I.A. Stegun, Handbook of Mathematical Functions: with Formulas, Graphs, and Mathematical Tables, vol. 55 (Courier Corporation, Hartford, 1964)

    MATH  Google Scholar 

  • J. Artru, T. Farges, P. Lognonné, Acoustic waves generated from seismic surface waves: propagation properties determined from doppler sounding observation and normal-modes modeling. Geophys. J. Int. 158, 1067–1077 (2004). doi:10.1111/j.1365-246X.2004.02377.x

    Article  ADS  Google Scholar 

  • J. Artru, V. Ducic, H. Kanamori, P. Lognonné, M. Murakami, Ionospheric detection of gravity waves induced by tsunamis. Geophys. J. Int. 160, 840–848 (2005). doi:10.1111/j.1365-246X.2005.02552.x

    Article  ADS  Google Scholar 

  • W.B. Banerdt, S. Smrekar, P. Lognonné, T. Spohn, S.W. Asmar, D. Banfield, L. Boschi, U. Christensen, V. Dehant, W. Folkner, D. Giardini, W. Goetze, M. Golombek, M. Grott, T. Hudson, C. Johnson, G. Kargl, N. Kobayashi, J. Maki, D. Mimoun, A. Mocquet, P. Morgan, M. Panning, W.T. Pike, J. Tromp, T. van Zoest, R. Weber, M.A. Wieczorek, R. Garcia, K. Hurst, InSight: a discovery mission to explore the interior of Mars, in Lunar and Planetary Science Conference, vol. 44, 2013, p. 1915

    Google Scholar 

  • D. Banfield, InSight Science Team, Atmospheric observations from the Mars insight mission, in Mars Atmosphere: Modelling and Observation, 5th International Workshop, ed. by F. Forget, M. Millour, 2014, p. 4304

    Google Scholar 

  • H.E. Bass, J.P. Chambers, Absorption of sound in the Martian atmosphere. J. Acoust. Soc. Am. 109, 3069–3071 (2001)

    Article  ADS  Google Scholar 

  • Q. Brissaud, R. Martin, R. Garcia, D. Komatitsch, Finite-difference numerical modelling of gravito-acoustic wave propagation in a windy and attenuating atmosphere. Geophys. J. Int. (2016). doi:10.1093/gji/ggw121

    Google Scholar 

  • S.L. Bruinsma, J.M. Forbes, Medium- to large-scale density variability as observed by CHAMP. Space Weather 6, 08002 (2008). doi:10.1029/2008SW000411

    Article  ADS  Google Scholar 

  • J.M. Carcione, Wave Fields in Real Media: Wave Propagation in Anisotropic, Anelastic, Porous and Electromagnetic Media, 2nd edn. (Elsevier, Oxford, 2007)

    Google Scholar 

  • J.M. Carcione, D. Kosloff, R. Kosloff, Wave propagation simulation in a linear viscoelastic medium. Geophys. J. Int. 95(3), 597–611 (1988)

    Article  MATH  ADS  Google Scholar 

  • J.E. Chappelow, V.L. Sharpton, Atmospheric variations and meteorite production on Mars. Icarus 184, 424–435 (2006). doi:10.1016/j.icarus.2006.05.013

    Article  ADS  Google Scholar 

  • A. Colaïtis, A. Spiga, F. Hourdin, C. Rio, F. Forget, E. Millour, A thermal plume model for the Martian convective boundary layer. J. Geophys. Res., Planets 118, 1468–1487 (2013). doi:10.1002/jgre.20104

    Article  ADS  Google Scholar 

  • C. de Groot-Hedlin, M.A.H. Hedlin, K. Walker, Finite difference synthesis of infrasound propagation through a windy, viscous atmosphere: application to a bolide explosion detected by seismic networks. Geophys. J. Int. 185(1), 305–320 (2011). doi:10.1111/j.1365-246X.2010.04925.x

    Article  ADS  Google Scholar 

  • J.-X. Dessa, J. Virieux, S. Lambotte, Infrasound modeling in a spherical heterogeneous atmosphere. Geophys. Res. Lett. 32, 12808 (2005). doi:10.1029/2005GL022867

    Article  ADS  Google Scholar 

  • F. Ding, W. Wan, H. Yuan, The influence of background winds and attenuation on the propagation of atmospheric gravity waves. J. Atmos. Sol.-Terr. Phys. 65, 857–869 (2003). doi:10.1016/S1364-6826(03)00090-7

    Article  ADS  Google Scholar 

  • W.N. Edwards, Meteor generated infrasound: theory and observation, in Infrasound Monitoring for Atmospheric Studies (Springer, Amsterdam, 2010), pp. 361–414

    Chapter  Google Scholar 

  • F. Forget, F. Hourdin, R. Fournier, C. Hourdin, O. Talagrand, M. Collins, S.R. Lewis, P.L. Read, J.-P. Huot, Improved general circulation models of the Martian atmosphere from the surface to above 80 km. J. Geophys. Res. 104, 24155–24176 (1999). doi:10.1029/1999JE001025

    Article  ADS  Google Scholar 

  • F. González-Galindo, F. Forget, M.A. López-Valverde, M. Angelats i Coll, E. Millour, A ground-to-exosphere Martian general circulation model, I: seasonal, diurnal, and solar cycle variation of thermospheric temperatures. J. Geophys. Res., Planets 114, 04001 (2009). doi:10.1029/2008JE003246

    Article  ADS  Google Scholar 

  • W. Gropp, E. Lusk, A. Skjellum, Using MPI, portable parallel programming with the Message-Passing Interface (MIT Press, Cambridge, USA, 1994)

    MATH  Google Scholar 

  • A.D. Hanford, L.N. Long, The direct simulation of acoustics on Earth, Mars, and Titan. J. Acoust. Soc. Am. 125, 640 (2009). doi:10.1121/1.3050279

    Article  ADS  Google Scholar 

  • C.P. Haynes, C. Millet, A sensitivity analysis of meteoric infrasound. J. Geophys. Res., Planets 118, 2073–2082 (2013). doi:10.1002/jgre.20116

    Article  ADS  Google Scholar 

  • C.O. Hines, Internal atmospheric gravity waves at ionospheric heights. Can. J. Phys. 38, 1441 (1960). doi:10.1139/p60-150

    Article  ADS  Google Scholar 

  • D. Komatitsch, R. Martin, An unsplit convolutional perfectly matched layer improved at grazing incidence for the seismic wave equation. Geophysics 72(5), 155–167 (2007)

    Article  ADS  Google Scholar 

  • S.R. Lewis, M. Collins, P.L. Read, F. Forget, F. Hourdin, R. Fournier, C. Hourdin, O. Talagrand, J.-P. Huot, A climate database for mars. J. Geophys. Res. 104, 24177–24194 (1999). doi:10.1029/1999JE001024

    Article  ADS  Google Scholar 

  • J.Y. Liu, Y.B. Tsai, S.W. Chen, C.P. Lee, Y.C. Chen, H.Y. Yen, W.Y. Chang, C. Liu, Giant ionospheric disturbances excited by the M9.3 Sumatra earthquake of 26 December 2004. Geophys. Res. Lett. 33, 2103 (2006). doi:10.1029/2005GL023963

    Article  ADS  Google Scholar 

  • R. Martin, D. Komatitsch, S.D. Gedney, E. Bruthiaux, A high-order time and space formulation of the unsplit perfectly matched layer for the seismic wave equation using auxiliary differential equations (ADE-PML). Comput. Model. Eng. Sci. 56(1), 17 (2010). doi:10.3970/cmes.2010.056.017

    MathSciNet  MATH  Google Scholar 

  • B. Mikhailenko, A. Mikhailov, Numerical modeling of seismic and acoustic-gravity waves propagation in an Earth-atmosphere model in the presence of wind in the air. Numer. Anal. Appl. 7(2), 124–135 (2014)

    Article  MATH  Google Scholar 

  • P.M. Morse, K.U. Ingard, Theoretical Acoustics (Princeton Univ. Press, New Jersey, 1968)

    Google Scholar 

  • C.J. Nappo, An Introduction to Atmospheric Gravity Waves (Academic Press, Amsterdam, 2002), p. 276

    Google Scholar 

  • V.E. Ostashev, D.K. Wilson, L. Liu, D.F. Aldridge, N.P. Symons, D. Marlin, Equations for finite-difference, time-domain simulation of sound propagation in moving inhomogeneous media and numerical implementation. J. Acoust. Soc. Am. 117(2), 503–517 (2005). doi:10.1121/1.1841531

    Article  ADS  Google Scholar 

  • A. Petculescu, R.M. Lueptow, Atmospheric acoustics of Titan, Mars, Venus, and Earth. Icarus 186, 413–419 (2007). doi:10.1016/j.icarus.2006.09.014

    Article  ADS  Google Scholar 

  • A.L. Pichon, K. Antier, Y. Cansi, B. Hernandez, E. Minaya, B. Burgoa, D. Drob, L.G. Evers, J. Vaubaillon, Evidence for a meteoritic origin of the September 15, 2007, Carancas crater. Meteorit. Planet. Sci. 43, 1797–1809 (2008). doi:10.1111/j.1945-5100.2008.tb00644.x

    Article  ADS  Google Scholar 

  • O. Popova, I. Nemtchinov, W.K. Hartmann, Bolides in the present and past martian atmosphere and effects on cratering processes. Meteorit. Planet. Sci. 38, 905–925 (2003). doi:10.1111/j.1945-5100.2003.tb00287.x

    Article  ADS  Google Scholar 

  • L.M. Rolland, G. Occhipinti, P. Lognonné, A. Loevenbruck, Ionospheric gravity waves detected offshore Hawaii after tsunamis. Geophys. Res. Lett. 37, 17101 (2010). doi:10.1029/2010GL044479

    Article  ADS  Google Scholar 

  • G. Tancredi, J. Ishitsuka, P.H. Schultz, R.S. Harris, P. Brown, D.O. Revelle, K. Antier, A. Le Pichon, D. Rosales, E. Vidal, M.E. Varela, L. Sánchez, S. Benavente, J. Bojorquez, D. Cabezas, A. Dalmau, A meteorite crater on Earth formed on September 15, 2007: the Carancas hypervelocity impact. Meteorit. Planet. Sci. 44, 1967–1984 (2009). doi:10.1111/j.1945-5100.2009.tb02006.x

    Article  ADS  Google Scholar 

  • J.D. Walker, Loading sources for seismological investigations of asteroids and comets. Int. J. Impact Eng. 29, 757–769 (2003). doi:10.1016/j.ijimpeng.2003.10.022

    Article  Google Scholar 

  • F.J.W. Whipple, The great Siberian meteor and the waves, seismic and aerial, which it produced. Q. J. R. Meteorol. Soc. 56, 287–304 (1930)

    Google Scholar 

  • J.-P. Williams, Acoustic environment of the Martian surface. J. Geophys. Res. 106, 5033–5042 (2001). doi:10.1029/1999JE001174

    Article  ADS  Google Scholar 

  • D.K. Wilson, L. Liu, Finite-difference, time-domain simulation of sound propagation in a dynamic atmosphere. Technical report, Cold Regions Research and Engineering Lab, Hanover Netherlands, 2004

Download references

Acknowledgements

We acknowledge Don Banfield and an anonymous reviewer for their detailed review of the manuscript. We thank the INSIGHT science team for fruitful discussions. We also thank the “Région Midi-Pyréenées” (France) and “Université fédérale de Toulouse” for funding the PhD grant of Quentin Brissaud. This study was also supported by CNES through space research scientific projects. Computer resources were provided by granted projects No. p1138 at CALMIP computing centre (Toulouse France), Nos. t2014046351 and t2015046351 at CEA centre (Bruyères, France). This is Insight Contribution Number 16.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Raphael F. Garcia.

Appendices

Appendix 1: Analytical Solution for Acoustic Wave Propagation in a Windy Homogeneous Atmosphere with Vibrational Relaxation

In the framework of acoustic wave propagation in a windy homogeneous medium with vibrational relaxation, the state and momentum equations read

$$ \begin{aligned} (\partial_{t} + \mathbf{w}.\nabla ) p &= -\nabla .\mathbf{v}\ast \varPhi -\rho c^{2} S, \\ \rho (\partial_{t} + \mathbf{w}.\nabla )\mathbf{v} &= -\nabla p, \end{aligned} $$
(22)

where, \(\forall t \in \mathbb{R}^{+}\)

$$ \varPhi (t) = H(t)\ast \psi (t), $$
(23)

with \(H\) the Heaviside function and \(\psi \) the relaxation function of the viscoelastic medium considered.

Taking the divergence of (22)-2 yields

$$ (\partial_{t} + \mathbf{w}.\nabla )\nabla .\mathbf{v} = - \frac{1}{ \rho }\Delta p, $$
(24)

where \(\Delta \) is the Laplacian operator. Thus, taking the total derivative \(D_{t}\) (where \(D_{t} = \partial_{t} + \mathbf{v}\nabla \)) of (22)-1 yields

$$ \begin{aligned}[b] (\partial_{t} + \mathbf{w}.\nabla )^{2} p &= -\nabla .(\partial_{t} + \mathbf{w}.\nabla )\mathbf{v}\ast \varPhi - (\partial_{t} + \mathbf{w}. \nabla )S \\ &= \frac{1}{\rho }\Delta p\ast \varPhi -\rho c^{2} (\partial_{t} + \mathbf{w}.\nabla )S. \end{aligned} $$
(25)

By considering, in (25), plane wave solutions of the form \(p = \tilde{p}e^{-i\omega t}\) and \(\varPsi \), defined in (16), the relaxation function for a Zener mechanism (Carcione et al. 1988), we can perform the complex integration of the convolution term

$$ \begin{aligned}[b] \Delta \tilde{p} \int_{-\infty }^{\infty} e^{-i\omega t}\tilde{\varPhi }H dt &= \Delta \tilde{p} \int_{-\infty }^{\infty} e^{-i\omega t}\frac{1}{ \tau_{\epsilon }}\biggl(1 - \frac{\tau_{\epsilon }}{\tau_{\sigma }}\biggr)e^{-\frac{t}{ \tau_{\sigma }}}H dt \\ &= \Delta \tilde{p}\frac{1}{\tau_{\epsilon }}\biggl(1 - \frac{\tau_{\epsilon }}{\tau_{\sigma }}\biggr) \int_{0}^{\infty} e^{-(i\omega + \frac{1}{ \tau_{\sigma }})t}dt \\ &=\Delta \tilde{p}\frac{1}{\tau_{\epsilon }}\biggl(1 - \frac{\tau_{\epsilon }}{\tau_{\sigma }}\biggr) \frac{1 + i\omega \tau_{\epsilon }}{1 + i\omega \tau_{\sigma }} \end{aligned} $$
(26)

then, from the plane wave form in the Fourier space \((\partial_{t} = -i \omega )\), and one can recast (25) as

$$ (-i\omega + \mathbf{w}.\nabla )^{2} \tilde{p} = \frac{\hat{K_{c}}}{ \rho }\Delta \tilde{p} - \rho c^{2}(-i\omega + \mathbf{w}.\nabla ) \tilde{S}. $$
(27)

The tilde denotes the time Fourier transform and \(\hat{K_{c}}\) is the complex bulk modulus, in the case of a single relaxation mechanism, which reads (Carcione et al. 1988)

$$ \hat{K_{c}} = \rho c^{2} \frac{1 + i\omega \tau_{\epsilon }}{1 + i \omega \tau_{\sigma }} = \rho \hat{c}^{2}, $$
(28)

where the complex velocity \(\hat{c}\) reads

$$ \hat{c} = c \sqrt{\frac{1 + i\omega \tau_{\epsilon }}{1 + i\omega \tau_{\sigma }} } $$
(29)

The complex equation (27) can then be expanded as

$$ \begin{gathered} \bigl(-\omega^{2} + w_{x}^{2} \partial_{x}^{2} - 2i\omega w_{x} \partial_{x}\bigr) \tilde{p} = \frac{\hat{K_{c}}}{\rho }\bigl( \partial_{x}^{2} + \partial _{z}^{2}\bigr) \tilde{p} - \rho c^{2}(-i\omega + \mathbf{w}.\nabla ) \tilde{S}, \\ \biggl\{ \frac{\rho }{\hat{K_{c}}}\bigl(-\omega^{2} + w_{x}^{2} \partial_{x} ^{2} - 2i\omega w_{x} \partial_{x}\bigr) - \partial_{x}^{2} - \partial_{z} ^{2} \biggr\} \tilde{p} = -\frac{\rho^{2}}{\hat{K_{c}}} c^{2}(-i \omega + \mathbf{w}.\nabla )\tilde{S}, \\ \biggl\{ \frac{\rho }{\hat{K_{c}}}\bigl(\omega^{2} - w_{x}^{2} \partial_{x} ^{2} + 2i\omega w_{x} \partial_{x}\bigr) + \partial_{x}^{2} + \partial_{z} ^{2} \biggr\} \tilde{p} = \frac{\rho^{2}}{\hat{K_{c}}} c^{2}(-i \omega + \mathbf{w}.\nabla )\tilde{S}, \\ \biggl\{ \frac{1}{\hat{c}^{2}}\bigl(\omega^{2} - w_{x}^{2} \partial_{x}^{2} + 2i\omega w_{x} \partial_{x}\bigr) + \partial_{x}^{2} + \partial_{z}^{2} \biggr\} \tilde{p} = \frac{\rho c^{2}}{\hat{c}^{2}} (-i \omega + \mathbf{w}.\nabla )\tilde{S}, \\ \bigl(\partial_{x}^{2} + \partial_{z}^{2} + \hat{k}^{2} + 2i\hat{k}\hat{M} \partial_{x} - \hat{M}^{2}\partial_{x}^{2}\bigr) \tilde{p} = - \frac{ \rho c^{2}}{\hat{c}^{2}} (-i\omega + \mathbf{w}.\nabla )\tilde{S}, \end{gathered} $$
(30)

with \(\hat{M}\) the Mach number, that is,

$$ \hat{M} = \frac{w_{x}}{\hat{c}} = \frac{w_{x}}{c \frac{1 + i\omega \tau_{\epsilon }}{1 + i\omega \tau_{\sigma }}}, $$
(31)

where \(\hat{c}\) is the complex velocity and \(k\) is the wavenumber \(\hat{k} = \frac{\omega }{\hat{c}}\). The hat over \(\hat{M}\) and \(\hat{k}\) is introduced to differentiate from the usual Mach and wave numbers in an inviscid medium.

Finally, we will consider

$$ \tilde{S} = \frac{2i\tilde{A}\hat{K_{c}}}{\omega \rho c^{2} }e^{-i \omega t}\delta_{x} \delta_{z}, $$
(32)

where \(\delta_{x}, \delta_{z}\) are the Kronecker delta symbols, and \(\tilde{A}\) is the Fourier transform of the amplitude of the time-dependent source function, that is,

$$ A(t) = -2(\pi f_{0})^{2}(t-t_{0})e^{-\{\pi f_{0}(t - t_{0})\}^{2}}. $$
(33)

We use the same method as in Ostashev et al. (2005) to obtain the pressure response in this framework, that is

$$ \tilde{p} = \frac{iA}{2(1-\tilde{M})^{3/2}} \biggl( H_{0}^{1}(\xi ) - \frac{i \tilde{M}\cos\beta }{\sqrt{1 - \tilde{M}^{2}\sin {\beta }^{2}}}H^{1} _{1}(\xi ) \biggr) e^{\frac{-i\tilde{k}\tilde{M}R\cos {\beta }}{1- \tilde{M}^{2}}}, $$
(34)

where \((H^{1}_{i})_{i=1,2}\) are the Hankel functions of first kind, and \(\xi \) reads

$$ \xi = \frac{\tilde{k}R\sqrt{1 - \tilde{M}^{2}\sin {\beta }^{2}}}{1- \tilde{M}^{2}}. $$
(35)

Using Hankel function asymptotic (Abramowitz and Stegun 1964) by considering the approximation \(|\tilde{k}R| >> 1\), yields

$$ \tilde{p} = \frac{A(\sqrt{1 - \tilde{M}^{2}\sin {\beta }^{2}}- \tilde{M}\cos {\beta })}{\sqrt{2\pi \tilde{k}R}(1-\tilde{M}^{2})(1- \tilde{M}^{2}\sin {\beta }^{2})^{3/4}}e^{\frac{i}{1-\tilde{M}^{2}}(\sqrt{1 - \tilde{M}^{2}\sin {\beta }^{2}}-\tilde{M}\cos {\beta })\tilde{k} R + \frac{i\pi }{4}}. $$
(36)

Appendix 2: Numerical Implementation of ADE-PML for a Medium Subject to Gravity

This boundary condition corresponds to mesh stretching in the boundary layer which requires to recast the derivatives in the new stretched set of complex-valued coordinates (Komatitsch and Martin 2007). We do this using ADE-PML memory variables. In this new reference frame (19) read

$$ \begin{aligned} \partial_{t} p & = -\mathbf{w}.\tilde{\nabla } p - \rho c^{2} \tilde{\nabla } . \mathbf{v} -\tilde{\nabla }p_{0} \mathbf{v} + e_{1}, \\ \partial_{t} \rho_{p} &= -\mathbf{w}.\tilde{\nabla } \rho_{p} - \tilde{\nabla }.(\rho \mathbf{v}), \\ \rho \partial_{t}\mathbf{v} &= - \rho \bigl\{ (\mathbf{v}.\tilde{ \nabla }) \mathbf{w} + (\mathbf{w}.\tilde{\nabla })\mathbf{v} \bigr\} + \tilde{ \nabla } . \boldsymbol{\Sigma } + \tilde{\mathbf{g}}\rho_{p}, \\ \partial_{t} e_{1} &= -\frac{1}{\tau_{\sigma }}\biggl[\biggl(1 - \frac{\tau_{ \sigma }}{\tau_{\epsilon }}\biggr) + e_{1}\biggr], \end{aligned} $$
(37)

with

$$ (\boldsymbol{\Sigma })_{ij} = -p\delta_{ij} + \mu \biggl( \tilde{\partial }_{j}( \mathbf{v}+\mathbf{w})_{i} + \tilde{ \partial }_{i}(\mathbf{v}+ \mathbf{w})_{j} - \frac{2}{3}\delta_{ij}\tilde{\nabla }.\mathbf{v}\biggr) + \eta_{V}\delta_{ij}\tilde{\nabla }.\mathbf{v}. $$

Note that we wrote the pressure equation (37)-1 in terms of the ambient pressure \(p_{0}\) rather than in terms of its value coming from hydrostatic equilibrium (i.e. in the interior domain \(\nabla p_{0} = \rho \mathbf{g}\)). Indeed, as we stretch coordinates the hydrostatic equilibrium equation is also modified as

$$ \tilde{\nabla } p_{0} = \rho \mathbf{g} . $$
(38)

In this new reference frame derivatives will read, for any unknown \(\phi \)

$$ \tilde{\partial }_{x} \phi = \frac{1}{\kappa_{x}}\partial_{x} \phi + Q _{x}^{\phi } ; \qquad \tilde{\partial }_{z} \phi = \frac{1}{\kappa_{z}}\partial_{z}\phi + Q_{z}^{\phi } , $$
(39)

where the memory variable is denoted \(Q_{x,z}^{\phi }\). Memory variables obey a first-order in time differential equation for \(x\)-derivatives and \(z\)-derivatives. We will only give details for the \(x-\)derivative since it is similar for the \(z\)-derivatives. We consider the four time sub-steps \((t^{n,i})_{i=1,4}\) as \(t^{n,i} = \Delta t (n+\nu_{i})\) for the four RK4 stages where \(\nu_{i} = 0, \nu_{2} = 0.5, \nu_{3} = 0.5\) and \(\nu_{4} = 1\). Then the memory variable \(Q_{x}^{\phi }\) obey the following differential equation

$$ Q_{x}^{\phi }\bigl(t^{n,i}\bigr) = b_{x,i} Q_{z}^{\phi }\bigl(t^{n,1}\bigr) + a_{x,i} \partial_{x}\phi \bigl(t^{n,i}\bigr) , $$
(40)

where

$$ a_{x,i} = -d_{x}\frac{\Delta t\nu_{i}}{\kappa_{x}\kappa_{x}(1+\Delta t \theta \nu_{i}(\frac{d_{x}}{\kappa_{x}}+b_{x}))} ; \qquad b_{x} = \frac{ \kappa_{x}-\theta \Delta t\nu_{i}(d_{x} + \kappa_{x}\alpha_{x})}{ \kappa_{x}+\theta \Delta t\nu_{i}(d_{x} + \kappa_{x}\alpha_{x})} , $$
(41)

with

$$ \begin{gathered} \tau_{x} = \frac{1}{\frac{d_{x}}{\kappa_{x}}+\alpha_{x}}; \qquad \kappa _{x} = 1 + ( \kappa_{max}-1)^{m}; \\d_{x} = d_{0} \biggl(\frac{x}{L}\biggr)^{N}; \qquad \alpha_{x} = \alpha_{max}\biggl[1-\biggl(\frac{x}{L}\biggr)^{p} \biggr] , \end{gathered} $$
(42)

where the \(\theta \) parameter indicates if the computation is performed explicitly (\(\theta = 0\)), semi-implicitly (\(\theta = 0.5\)) or implicitly (\(\theta = 1\)). Here this parameter is taken as \(\theta = 0.5\). Since we are mainly interested in damping acoustic waves (most of the waves reaching the boundary will be acoustic waves) we take the same parameters as in Martin et al. (2010), meaning that

$$ \begin{gathered} N = 2; \qquad m = 2; \qquad p = 1; \qquad \alpha_{max} = \pi f_{0}; \\ \kappa_{max} = 7;\qquad d_{0} = - \frac{(N+1)c_{max,x}log(R_{c})}{2L} , \end{gathered} $$
(43)

where \(f_{0}\) is the dominant frequency of the source, \(c_{max,x}\) the maximum effective velocity along \(x\) (sum of background flow velocity and adiabatic sound speed) and \(R_{c}\) the theoretical reflection coefficient taken as \(R_{c} = 0.0001\). The same formulation holds for derivatives along \(z\).

Appendix 3: “Image” Boundary Conditions

This boundary condition consists in duplicating the physical domain and sources at ghost points located below the bottom boundary in order to obtain a wave interference ensuring zero vertical velocity at the boundary (Morse and Ingard 1968, Sect. 7.4). One enforces that the pressure is equal on each side of the boundary

$$ p(-\mathbf{x}, t) = p(\mathbf{x}, t), \quad \forall \mathbf{x} \in \varOmega \backslash \varGamma $$
(44)

and the sign of gravity and wind is changed in the ghost domain as

$$ \begin{aligned} g_{y}(-\mathbf{x}) &= - g_{y}(\mathbf{x}), \\ w_{x}(-\mathbf{x}) &= -w_{x}(-\mathbf{x}), \quad \forall \mathbf{x} \in \varOmega \backslash \varGamma . \end{aligned} $$
(45)

while all other unknowns are duplicated from the real domain to the ghost one

$$ \begin{aligned} \mu (-\mathbf{x}) &= \mu (\mathbf{x}), \eta_{V}(-\mathbf{x}) = \eta _{V}(\mathbf{x}), \\ \tau_{\epsilon }(-\mathbf{x}) &= \tau_{\epsilon }(\mathbf{x}), \tau_{\sigma }(-\mathbf{x}) = \tau_{\sigma }(\mathbf{x}), \quad \forall \mathbf{x} \in \varOmega \backslash \varGamma \end{aligned} $$
(46)

As a consequence, the pressure being equal on both sides of the boundary, the pressure gradient in the momentum equation is zero at the boundary. Also, setting opposite gravity directions above and below the boundary leads to zero gravity at the surface, hence the gravity term in the momentum equation vanishes. Therefore, at the boundary, the momentum equation is verified since the zero velocity condition is imposed. Finally, we impose the boundary to be motionless and we choose

$$ \begin{aligned} v_{x}(x, -z) &= -v_{x}(x, z), \\ v_{z}(x, -z) &= -v_{z}(x, z), \quad \forall z \in \varOmega. \end{aligned} $$
(47)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Garcia, R.F., Brissaud, Q., Rolland, L. et al. Finite-Difference Modeling of Acoustic and Gravity Wave Propagation in Mars Atmosphere: Application to Infrasounds Emitted by Meteor Impacts. Space Sci Rev 211, 547–570 (2017). https://doi.org/10.1007/s11214-016-0324-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11214-016-0324-6

Keywords

Navigation