Log in

On the Nature of Potential Energy in Atmospheric Gravity Waves, or Why the Atmosphere Cannot Sustain a Tsunami

  • Published:
Pure and Applied Geophysics Aims and scope Submit manuscript

Abstract

Motivated by the generation of exceptionally large gravito-elastic waves during the Hunga Tonga–Hunga Ha’apai explosion of 15 January 2022, we examine theoretically the nature of the main air wave branch \(GR_0,\) whose undispersed celerity, \(\sim 313\) m/s, suggests that it may represent a “tsunami” of the atmospheric column for an effective thickness \(H_{eff}\approx 10\) km. However, we find that its potential energy is about 90% elastic across a wide frequency band, thus negating the widely held perception that it constitutes an oscillation between kinetic and gravitational energy. Based on the systematic study of the effect of finite compressibility on the dispersion and potential energy of a classic oceanic tsunami, we confirm that this feature of the branch \(GR_0\) stems from the similarity between its celerity and the average speed of sound in the atmosphere. We then show that this similarity is not fortuitous, but rather expected for a perfect gas, which, we conclude, cannot sustain “tsunamis”, i.e., oscillations between kinetic and gravitational energy.

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

Access this article

Price includes VAT (Canada)

Instant access to the full article PDF.

Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10
Figure 11
Figure 12
Figure 13
Figure 14
Figure 15

Similar content being viewed by others

Data availability

The paper does not use any data.

Notes

  1. The interpretation of these early papers is occasionally made difficult by the then customary use of the word “tidal waves” to describe tsunamis, i.e., gravitational oscillations of an oceanic mass which we now understand have nothing to do with tides.

  2. Throughout this paper, we use \(\Theta \) for absolute temperature, to distinguish it from periods T.

  3. A physical explanation of this situation is that the “free air” component of the restoring force (due to displacement in a gravity field including buoyancy) vanishes, but the “Bouguer” one (due to a change in that field upon a change of material properties) does not.

References

  • Abdolali, A., & Kirby, J. T. (2017). Role of compressibility on tsunami propagation. Journal of Geophysical Research Oceans, 122, 9780–9794.

    ADS  Google Scholar 

  • Abramowitz, M., & Stegun, I. (1965). Handbook of mathematical functions (8th ed.). Dover.

    Google Scholar 

  • Ben-Menahem, A. (1975). Source parameters of the Siberian explosion of June 30, 1908 from analysis and synthesis of seismic signals at four stations. Physics of the Earth and Planetary Interiors, 11, 1–35.

    ADS  Google Scholar 

  • Boyle, R. (1662). A defence of the doctrine touching the spring and weight of the air. Robinson.

    Google Scholar 

  • Brunt, D. (1927). The period of simple vertical oscillations in the atmosphere. Quarterly Journal of the Royal Meteorological Society, 53, 30–32.

    ADS  Google Scholar 

  • Carvajal, M., Sepúlveda, I., Gubler, A., & Garreaud, R. (2022). Worldwide signature of the 2022 Tonga volcanic tsunami. Geophysical Research Letters, 49, e2022GL098153.

    ADS  Google Scholar 

  • Dahlen, F. A., & Tromp, J. (1998). Theoretical global seismology. Princeton University Press.

    Google Scholar 

  • Dean, R. G., & Dalrymple, R. A. (2000). Water wave mechanics for engineers and scientists. World Scientific.

    Google Scholar 

  • Donn, W. L., & Ewing, W. M. (1962). Atmospheric waves from nuclear explosions—Part II: The Soviet test of 30 October 1961. Journal of Atmospheric Sciences, 19, 264–273.

    ADS  Google Scholar 

  • Dziewonski, A. M., & Anderson, D. L. (1981). Preliminary reference Earth model. Physics of the Earth and Planetary Interiors, 25, 297–356.

    ADS  Google Scholar 

  • Ewing, W. M., & Press, F. (1955). Tide-gage disturbances from the great eruption of Krakatoa. Transactions of the American Geophysical Union, 36, 53–60.

    Google Scholar 

  • Gill, A. E. (1982). Atmosphere-ocean dynamics. Academic Press.

    Google Scholar 

  • Gusman, A. R., Roger, R., Noble, C., Wang, X., & Power, W. (2022). The 2022 Hunga Tonga-Hunga Ha’apai Volcano air-wave generated tsunami. Pure and Applied Geophysics, 179, 3511–3525.

    ADS  Google Scholar 

  • Harkrider, D. G. (1964). Theoretical and observed acoustic-gravity waves from explosive sources in the atmosphere. Journal of Geophysical Research, 69, 2595–5321.

    Google Scholar 

  • Harkrider, D. G., & Press, F. (1967). The Krakatoa air-sea waves: An example of pulse propagation in coupled systems. Geophysical Journal of the Royal astronomical Society, 13, 149–159.

    ADS  Google Scholar 

  • Haskell, N. A. (1953). The dispersion of surface waves in multilayered media. Bulletin of the Seismological Society of America, 43, 17–24.

    MathSciNet  Google Scholar 

  • Hébert, H., Reymond, D., Krien, Y., Vergoz, J., Schindelé, F., Roger, J., & Loevenbruck, A. (2009). The 15 August 2007 Peru earthquake and tsunami: Influence of the source characteristics on the tsunami heights. Pure and Applied Geophysics, 166, 211–232.

    ADS  Google Scholar 

  • Holton, J. R. (2004). Dynamic meteorology (4th ed.). Elsevier Academic Press.

    Google Scholar 

  • Kanamori, H., & Cipar, J. J. (1974). Focal process of the great Chilean earthquake, May 22, 1960. Physics of the Earth and Planetary Interiors, 9, 128–136.

    ADS  Google Scholar 

  • Kanamori, H., Mori, J., & Harkrider, D. G. (1994). Excitation of atmospheric oscillations by volcanic eruptions. Journal of Geophysical Research, 99, 21947–21961.

    ADS  Google Scholar 

  • Kittel, C., & Kroemer, H. (1980). Thermal physics. W.H. Freeman.

    Google Scholar 

  • Kramp, C. (1808). Eléments d’arithmétique universelle. Thiriart.

    Google Scholar 

  • Lamb, H. (1910). On atmospheric oscillations. Proceedings of the Royal Society (London), Series A, 84, 551–572.

    ADS  Google Scholar 

  • Lamb, H. (1932). Hydrodynamics (6th ed.). Cambridge University Press.

    Google Scholar 

  • Laplace, P. S. (1805). Traîté de mécanique céleste (Vol. 4). Duprat.

    Google Scholar 

  • Lognonné, P., Clévédé, E., & Kanamori, H. (1998). Computation of seismograms and atmospheric oscillations by normal-mode summation for a spherical Earth model with realistic atmosphere. Geophysical Journal International, 135, 388–406.

    ADS  Google Scholar 

  • Minzner, R. A., Champion, K. S. W., & Pond, H. L. (1959). The ARDC model atmosphere. Air Force Office of Scientific Research.

    Google Scholar 

  • Nakamura, K., & Watanabe, H. (1961). Tsunami forerunner observed in case of the Chile Tsunami of 1960. In Report on the Chilean Tsunami of May 24, 1960, as observed along the Coast of Japan. Committee for Field Investigation of the Chilean Tsunami of 1960.

  • Newton, I. (1687). Philosophiæ naturalis principa mathematica. Strester.

    Google Scholar 

  • Okal, E. A. (1982). Mode-wave equivalence and other asymptotic problems in tsunami theory. Physics of the Earth and Planetary Interiors, 30, 1–11.

    ADS  Google Scholar 

  • Okal, E. A. (1988). Seismic parameters controlling far-field tsunami amplitudes: A review. Natural Hazards, 1, 67–96.

    Google Scholar 

  • Okal, E. A. (2022). Air waves from the 2022 Tonga explosion: Theoretical studies and an oversight in the reporting of DART sensor data. Seismological Research Letters, 93, 1187–1188. [abstract].

    Google Scholar 

  • Okal, E. A., & Talandier, J. (1991). Single-station estimates of the seismic moment of the 1960 Chilean and 1964 Alaskan earthquakes, using the mantle magnitude \(M_m\). Pure and Applied Geophysics, 136, 103–126.

    ADS  Google Scholar 

  • Pekeris, C. L. (1937). Atmospheric oscillations. Proceedings of the Royal Society (London), Series A, 158, 650–671.

    ADS  Google Scholar 

  • Pekeris, C. L. (1939). The propagation of a pulse in the atmosphere. Proceedings of the Royal Society (London), Series A, 171, 434–449.

    ADS  Google Scholar 

  • Press, F., & Harkrider, D. G. (1962). Propagation of acoustic-gravity waves in the atmosphere. Journal of Geophysical Research, 67, 3889–3908.

    ADS  Google Scholar 

  • Rabinovich, A. B., Woodworth, P. L., & Titov, V. V. (2011). Deep-sea observations and modeling of the 2004 Sumatra tsunami in Drake Passage. Geophysical Research Letters, 38, L16604.

    ADS  Google Scholar 

  • Saito, M. (1967). Excitation of free oscillations and surface waves by a point source in a vertically heterogeneous Earth. Journal of Geophysical Research, 72, 3689–3699.

    ADS  Google Scholar 

  • Taylor, G. I. (1937). The oscillations of the atmosphere. Proceedings of the Royal Society (London), Series A, 158, 318–326.

    Google Scholar 

  • Titov, V. V., Kânoğlu, U., & Synolakis, C. E. (2016). Development of MOST for real-time tsunami forecasting. Journal of Waterways, Port and Coastal Engineering, 142, 03116004.

    Google Scholar 

  • Tsai, V. C., Ampuero, J.-P., Kanamori, H., & Stevenson, D. J. (2013). Estimating the effect of Earth elasticity and variable water density on tsunami speeds. Geophysical Research Letters, 40, 492–496.

    ADS  Google Scholar 

  • UNESCO. (1981). Background papers and supporting data on the International Equation of State of seawater. UNESCO Technical Papers in Marine Science, 38, 192 pp., Paris.

  • Väisälä, V. (1925). Über die Wirkung der Windschwankungen auf die Pilotbeobachtungen. Societas Scientarium Fennica Commentationes, 2, 1–46.

    Google Scholar 

  • Ward, S. N. (1980). Relationships of tsunami generation and an earthquake source. Journal of Physics of the Earth, 28, 441–474.

    Google Scholar 

  • Wares, G. W., Champion, K. W., Pond, H. L., & Cole, A. E. (1960). Model atmospheres. In Handbook of geophysics (pp. (1-1)–(1-43)). The Macmillan Co.

  • Watada, S. (2013). Tsunami speed variations in density-stratified compressible global oceans. Geophysical Research Letters, 40, 4001–4006.

    ADS  Google Scholar 

  • Watada, S., Kusumoto, S., & Satake, K. (2014). Travel-time delay and initial phase reversal of distant tsunamis coupled with the self-gravitating elastic Earth. Journal of Geophysical Research, Solid Earth, 119, 4287–4310.

    ADS  Google Scholar 

  • Watanabe, S., Hamilton, K., Sakazaki, T., & Nakano, M. (2022). First detection of the Pekeris internal global atmospheric resonance: Evidence for the 2022 Tonga eruption and from global reanalysis data. Journal of Atmospheric Sciences, 79, 3027–3043.

    ADS  Google Scholar 

  • Wessel, P., & Smith, W. H. F. (1991). Free software helps map and display data. Eos, Transactions of the American Geophysical Union, 72, 441 and 445–446.

  • Wexler, H., & Hass, W. A. (1962). Global atmospheric pressure effects of the October 30, 1961, explosion. Journal of Geophysical Research, 67, 3875–3887.

    ADS  Google Scholar 

  • Whipple, F. J. W. (1930). The great Siberian meteor and the waves, seismic and aerial, which it produced. Quarterly Journal of the Royal Meteorological Society (London), 56, 287–304.

    Google Scholar 

  • Wiggins, R. A. (1976). A fast, new computational algorithm for free oscillations and surface waves. Geophysical Journal of the Royal astronomical Society, 47, 135–150.

    ADS  Google Scholar 

  • Yamamoto, R. (1957). Microbarographic oscillation produced by the Soviet explosion of a hydrogen bomb on 22 November, 1955. Bulletin of the American Meteorological Society, 38, 536–539.

    ADS  Google Scholar 

Download references

Acknowledgements

I thank Alexander Rabinovich for discussion on a previous version of the paper, and in particular for pointing out important references. Some figures were produced using the GMT package (Wessel & Smith, 1991). Bessel functions were computed using the Ke!san online calculator (http://www.keisan.casio.com).

Funding

This study used no external funding.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Emile A. Okal.

Ethics declarations

Conflict of interest

The author disclosed no competing interest.

Additional information

Publisher's Note

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

Appendices

Appendix 1

We recall here the definition of the 6 components of the eigenvector of a spheroidal mode in the formalism of Saito (1967, p. 3690), later used by Kanamori and Cipar (1974).

\(\bullet \) \(y_1 (r)\) represents the vertical (radial) component of particle motion, given as

$$\begin{aligned} u_r (r, \theta , \phi ; \, t) = y_1 (r) \, Y_l^m (\theta , \,\phi ). \end{aligned}$$
(38)

\(\bullet \) \(y_2 (r)\) represents the radial component of the traction, given as

$$\begin{aligned} \sigma _{rr} (r, \theta , \phi ; \, t) = y_2 (r) \, Y_l^m (\theta , \,\phi ). \end{aligned}$$
(39)

\(\bullet \) \(y_3 (r)\) represents the orthoradial component of particle motion, given as

$$\begin{aligned} u_{\theta } (r, \theta , \phi ; \, t) = y_3 (r) ~ \frac{\partial \, Y_l^m (\theta , \,\phi )}{\partial \,\theta }. \end{aligned}$$
(40)

\(\bullet \) \(y_4 (r)\) represents the shear component of the traction, given as

$$\begin{aligned} \sigma _{r\,\theta } (r, \theta , \phi ; \, t) = y_4 (r) ~ \frac{\partial \, Y_l^m (\theta , \,\phi )}{\partial \,\theta }. \end{aligned}$$
(41)

\(\bullet \) \(y_5 (r)\) represents the change in gravity potential, given as

$$\begin{aligned} \psi (r, \theta , \phi ; \, t) = y_5 (r) \, Y_l^m (\theta , \,\phi ). \end{aligned}$$
(42)

\(\bullet \) Finally, \(y_6 (r)\) is simply defined as

$$\begin{aligned} y_6 (r) = \frac{d y_5 (r) }{dr} - 4 \pi G\, \rho \, y_1 (r) \end{aligned}$$
(43)

where the \(Y_l^m\) are the spherical harmonics of degree l and order m. In all above equations, the time dependence \(e^{\,i \omega t }\) has been omitted for simplicity. Obviously, \(y_1\) and \(y_3\) have dimensions of length, \(y_2\) and \(y_4\) of pressure, \(y_5\) of velocity squared, and \(y_6\) of acceleration.

In this fashion, all boundary conditions between spherical shells with different mechanical properties simply require the continuity of the six components of the vector y, with the exception of \(y_3\) when at least one of the layers is fluid.

Note that in all fluid layers (outer core, ocean, atmosphere), \(y_4\) is identically zero, and \(y_3\) becomes a spurious variable

$$\begin{aligned} y_3 = \frac{1}{r \, \omega ^2}\;\left[ g \, y_1 - \frac{y_2}{\rho} - y_5\right] \end{aligned}$$
(44)

so that the differential system becomes 4-dimensional. Also, in a fluid, \(y_2\) is simply the opposite of the overpressure during the oscillation.

Finally, for large l and \(\theta \) not close to 0 or \(\pi ,\) asymptotic expansions of the \(Y_l^m\) show that the partial derivative in (40) results in an orthoradial particle displacement \(u_{\theta }\) of order \((l\, y_3)\) while the vertical component remains of order \(y_1.\)

Appendix 2

We summarize here some of the steps in L10’s derivation of his Equations (63) and (66) p. 563. In particular, we emphasize the occasionally different notation used in his paper.


1. Note that L10 orients the vertical axis (which he calls y) downwards with the origin at the top of the atmosphere, which is infinite in the isothermal model and otherwise depends on the particular structure used. On the other hand, we call it z and orient it upwards, with the origin consistently at the bottom of the atmosphere.


2. For a layering of the form (13), L10 does not use the parameter \(\phi ,\) but rather defines the [absolute] temperature gradient

$$\begin{aligned} \beta= & {} - \,\frac{d \, \Theta }{dz} = \frac{\phi \,-1}{\phi }\,\cdot \,\frac{\Theta _0}{H} \\= & {} \frac{\phi \,-1}{\phi }\,\cdot \,\frac{g M}{R} \quad (\text {units:}\ \text {K}/\text {km}). \end{aligned}$$
(45)

Conversely,

$$\begin{aligned} \phi =\frac{1}{(1-\beta \,R/Mg)} = \frac{1}{(1-\beta \,H/\Theta _0 )} \end{aligned}$$
(46)

with H defined as

$$\begin{aligned} H= \frac{R\, \Theta _0}{ M \,g } \end{aligned}$$
(12)

(Note that L10’s notation is R for our R/M).


3. Define

$$\begin{aligned} m =\frac{1}{\phi -1} =\frac{\Theta _0}{\beta \,H} -1 =\frac{g M}{R \beta } - 1 \quad (\text {dimensionless}). \end{aligned}$$
(47)

Note that Lamb’s notation is n in L10; m in Lamb (1932).


4. Define h as the full height of the atmosphere

$$\begin{aligned} h =\frac{\Theta _0}{\beta } =\frac{\phi }{\phi -1}\,\cdot \,H =m\,\phi \,H \end{aligned}$$
(48)

where h is defined in Eq. (12). h is equivalent to our \(\zeta \) in (16).


5. For the isentropic case \((\phi =\gamma =1.4)\)

$$\begin{aligned} \beta =\beta _S =\frac{2}{7}\,\cdot \,\frac{\Theta _0}{H} \quad m=m_S=\frac{5}{2}\quad h=h_S=\frac{7}{2}\,H. \end{aligned}$$
(49)

(L10’s notation: \(\beta _S=\beta _1\)).


6. For the isothermal case (\(\phi =1\))

$$\begin{aligned} \beta= & {} \beta _{\Theta }=0 \\ m= m_{\Theta }\;\;\rightarrow \;\;\infty \quad h=h_{\Theta }\;\;\rightarrow \;\;\infty . \end{aligned}$$
(50)

7. Note that L10 uses V for the celerity of the atmospheric “Lamb” wave (our C), and c for the speed of sound (our \(\alpha \)).


8. In the course of his derivation, L10 uses the potentially confusing notation \(\Pi (x)\) for the factorial: \(\Pi (x)=\Gamma (x+1)\) (x real) or x! (x integer), even though the latter had been introduced one century earlier by Kramp (1808).


9. Then, after considerable algebra, L10 derives the solution of the dispersion through the roots of his Equation (63 p. 563) reproduced here as (18), the celerity C of the Lamb wave being given by his Equation (66), reproduced as (19).

In (18) and (19), we prefer the notation \(\xi ,\) instead of L10’s \(\omega ,\) that dimensionless variable having no relation to an angular frequency.


10. In the limit of large m (\(\phi \rightarrow 1\)), and in the long-wavelength approximation, \(\xi \) is expected to itself be large, and one can use Abramowitz and Stegun’s (1965) Equation (9.3.1) p. 365:

$$\begin{aligned} J_m (z) \approx \frac{1}{\sqrt{2\pi \,m}}\,\cdot \, \left[ \frac{e \,z}{2\,m} \right] ^{\,m}. \end{aligned}$$
(51)

Hence

$$\begin{aligned}{} & {} \frac{z \, J_{m + 1} (z) }{2 \,J_m (z) } = \frac{e\,z^2}{4}\,\cdot \,\sqrt{m / (m+1)}\,\cdot \,\frac{m^m}{(m+1)^{m+1}} \\{} & {} \quad = \frac{e\,z^2}{4\,(m + 1)}\,\cdot \,[m/(m + 1)]^{ m + 1/2}. \end{aligned}$$
(52)

The solution to (17) is then

$$\begin{aligned} \xi ^2=(\beta _S / \beta -1)\,\cdot \,\frac{4 (m +1)}{e}\,\cdot \,[(m + 1)/m]^{m + 1/2} \end{aligned}$$
(53)

and substituting into L10’s Equation (66) p. 563,

$$\begin{aligned} C^2 = g H\,\cdot \frac{e}{(1 + 1/m)^m}\,\cdot \,(1 + 1/m)^{ -1/2}. \end{aligned}$$
(54)

In the limit \(m \rightarrow \infty ,\) the fraction in (54) goes to 1, and so does the last term in parentheses, so that \(C^2\approx gH,\) which justifies L10’s claim that the celerity of the “Lamb” air wave observed during the Krakatau explosion coincides with that of a would-be tsunami for a column of height H defined by (12). But as shown in the present study, that does not imply that the structure of the wave is that of a tsunami.

However, the approximation (51), on which this result is based, is valid only for large m,  i.e., when the layering is close to isothermal. If, on the opposite, \(\phi \) approaches \(\gamma \) (isentropic layering; \(m \rightarrow 5/2\)), then the parenthesis \((\beta _S / \beta - 1) \rightarrow 0\) but \(\xi \) will remain finite, in practice close to 7, the first non-zero root of \(J_{7/2} (\xi ).\) The celerity of the Lamb wave will also approach 0 as

$$\begin{aligned} C^2\approx (2/7) \,(\beta _S / \beta - 1)\,\cdot \, g H \end{aligned}$$
(55)

which is equivalent to L10’s first [un-numbered] equation on Page 564, except for a typographic error in the parenthesis which is identically zero as typeset in L10.

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

Okal, E.A. On the Nature of Potential Energy in Atmospheric Gravity Waves, or Why the Atmosphere Cannot Sustain a Tsunami. Pure Appl. Geophys. 181, 1–25 (2024). https://doi.org/10.1007/s00024-023-03402-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00024-023-03402-y

Keywords

Navigation