Log in

Baroclinic Model of Jupiter’s Great Red Spot

  • Published:
Izvestiya, Atmospheric and Oceanic Physics Aims and scope Submit manuscript

Abstract

This paper proposes a quasi-geostrophic baroclinic model of Jupiter’s Great Red Spot (GRS) as a localized eddy formation in a continuously stratified rotating atmosphere under the action of a horizontal shear flow in the f-plane approximation. On the basis of the theory of ellipsoidal vortices, an analytical relationship is obtained between the geometric dimensions of the vortex, the potential vorticity of the vortex core, and the characteristics of the background flow. Measurements of a number of characteristics of both the vortex and the background flow in the Voyager 1 (1979), Galileo (1996), and Cassini (2000) missions were used. Based on the theory, the vertical size of the GRS was calculated, which turned out to be close to the same characteristic measured in the Voyager 1 (1979) mission.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (France)

Instant access to the full article PDF.

Fig. 1.
Fig. 2.

REFERENCES

  1. Antipov, S.V., Nezlin, M.V., Snezhkin, E.N., and Trubnikov, A.S., Rossby autosoliton and a laboratory model of the Jupiter Great Red Spot, Zh. Eksp. Teor. Fiz., 1985, vol. 89, no. 6, pp. 1905–1920.

    Google Scholar 

  2. Beaumont, D.N., Solitary waves on an unsymmetrical shear flow with applications to Jupiter’s Great Red Spot, Icarus, 1980, vol. 41, 400–409.

    Article  Google Scholar 

  3. Bolton, S.J., Levin, S.M., Guillot, T., Li, C., Kaspi, Y., Orton, G., Wong, M.H., Oyafuso, F., Allison, M., Arballo, J., and Atreya, S., Microwave observations reveal the deep extent and structure of Jupiter’s atmospheric vortices, Science, 2021, vol. 374, no. 6570, pp. 968–972.

    Article  Google Scholar 

  4. Bouchet, F. and Sommeria, J., Emergence of intense jets and Jupiter’s Great Red Spot as maximum-entropy structures, J. Fluid Mech., 2002, vol. 464, pp. 165–207.

    Article  Google Scholar 

  5. Busse, F.H., A simple model of convection in the Jovian atmosphere, Icarus, 1976, vol. 29, no. 2, pp. 255–260.

    Article  Google Scholar 

  6. Busse, F.H., Convection driven zonal flows and vortices in the major planets, Chaos, 1994, vol. 4, no. 2, pp. 123–134.

    Article  Google Scholar 

  7. Chaplygin, S.A., Sobranie sochinenii (Collection of Works), vol. 2, Moscow: Gostekhizdat, 1948.

  8. Chavanis, P.H., Statistical mechanics of geophysical turbulence: Application to Jovian flows and Jupiter’s Great Red Spot, Phys. D (Amsterdam), 2005, vol. 200, nos. 3–4, pp. 257–272.

    Article  Google Scholar 

  9. Choi, D.S. and Banfield, D., Velocity and vorticity measurements of Jupiter’s Great Red Spot using automated cloud feature tracking, Icarus, 2007, vol. 188, no. 1, pp. 35–46.

    Article  Google Scholar 

  10. Dowling, T.E. and Ingersoll, A.P., Jupiter’s Great Red Spot as a shallow water system, J. Atmos. Sci., 1989, vol. 46, no. 21, pp. 3256–3278.

    Article  Google Scholar 

  11. Flierl, G.R., Baroclinic solitary waves with radial symmetry, Dyn. Atmos. Oceans, 1979, vol. 3, no. 1, pp. 15–38.

    Article  Google Scholar 

  12. Golitsyn, G.S., A similarity approach to the general circulation of planetary atmospheres, Icarus, 1970, vol. 13, no. 1, pp. 1–24.

  13. Heimpel, M.H., Yadav, R.K., Featherstone, N.A., and Aurnou, J.M., Polar and mid-latitude vortices and zonal flows on Jupiter and Saturn, Icarus, 2022, vol. 379, no. 6, p. 114942.

    Article  Google Scholar 

  14. Hide, R., Origin of Jupiter’s Great Red Spot, Nature, 1961, vol. 190, no. 4779, pp. 895–896.

    Article  Google Scholar 

  15. Ingersoll, A.P., Inertial Taylor columns and Jupiter’s Great Red Spot, J. Atmos. Sci., 1969, vol. 26, no. 4, pp. 744–752.

    Article  Google Scholar 

  16. Jalilian, P. and Liu, T., Analytical solution for large-scale rotating fluid layer with thermal convection, Fluid Dyn., 2019, vol. 54, no. 6, pp. 741–748.

    Article  Google Scholar 

  17. Jones, C.A. and Kuzanyan, K.M., Compressible convection in the deep atmospheres of giant planets, Icarus, 2009, vol. 204, no. 1, pp. 227–238.

    Article  Google Scholar 

  18. Kaspi, Y., Inferring the depth of the zonal jets on Jupiter and Saturn from odd gravity harmonics, Geophys. Res. Lett., 2013, vol. 40, no. 4, pp. 676–680.

    Article  Google Scholar 

  19. Kaspi, Y., Galanti, E., Hubbard, W.B., Stevenson, D.J., Bolton, S.J., Iess, L., Guillot, T., Bloxham, J., Connerney, J.E.P., Cao, H., and Durante, D., Jupiter’s atmospheric jet streams extend thousands of kilometres deep, Nature, 2018, vol. 555, no. 7695, pp. 223–226.

    Article  Google Scholar 

  20. Keller, V.S. and Yavorskaya, I.M., Modeling hydrodynamic processes in the atmospheres of large planets, in Aeromekhanika i gazovaya dinamika (Aeromechanics and Gas Dynamics), Moscow: Nauka, 1976, pp. 256–279.

  21. Kida, S., Motion of an elliptic vortex in a uniform shear flow, J. Phys. Soc. Japan, 1981, vol. 50, no. 10, pp. 3517–3520.

  22. Kirchhoff, G., Vorlesungen über mathematische Physik, Leipzig: Teubner, 1874; Moscow: AN SSSR, 1962.

  23. Larichev, V.D. and Reznik, G.M., On two-dimensional Rossby waves, Dokl. Akad. Nauk SSSR, 1976, vol. 231, no. 5, pp. 1077–1079.

    Google Scholar 

  24. Marcus, P.S., Numerical simulation of Jupiter’s Great Red Spot, Nature, 1988, vol. 331, no. 6158, pp. 693–696.

    Article  Google Scholar 

  25. Marcus, P.S. and Lee, C., Jupiter’s Great Red Spot and zonal winds as a self-consistent, one-layer, quasigeostrophic flow, Chaos, 1994, vol. 4, no. 2, pp. 269–286.

    Article  Google Scholar 

  26. Maxworthy, T. and Redekopp, L.G., New theory of the great red spot from solitary waves in the Jovian atmosphere, Nature, 1976, vol. 260, no. 5551, pp. 509–511.

    Article  Google Scholar 

  27. Meacham, S.P., Quasigeostrophical ellipsoidal vortices in stratified fluid, Dyn. Atmos. Oceans, 1992, vol. 16, nos. 3–4, pp. 189–223.

    Article  Google Scholar 

  28. Michel, J. and Robert, R., Statistical mechanical theory of the Great Red Spot of Jupiter, J. Stat. Phys., 1994, vol. 77, no. 3, pp. 645–666.

    Article  Google Scholar 

  29. Miller, J., Weichman, P.B., and Cross, M.C., Statistical mechanics, Euler’s equation, and Jupiter’s red spot, Phys. Rev. A, 1992, vol. 45, no. 4, pp. 2328–2359.

    Article  Google Scholar 

  30. Nezlin, M.V., Baroclinic modification of the barotropic model of the Great Red Spot of Jupiter, Pis’ma Zh. Eks-p. Teor. Fiz., 1981, vol. 34, no. 2, pp. 83–86.

    Google Scholar 

  31. Pankratov, K.K. and Zhmur, V.V., A dynamics of desingularized quasigeostrophic vortices, Phys. Fluids A, 1991, vol. 3, no. 5, p. 1464.

    Article  Google Scholar 

  32. Parisi, M., Kaspi, Y., Galanti, E., Durante, D., Bolton, S.J., Levin, S.M., Buccino, D.R., Fletcher, L.N., Folkner, W.M., Guillot, T., and Helled, R., The depth of Jupiter’s Great Red Spot constrained by Juno gravity overflights, Science, 2021, vol. 374, no. 6570, pp. 964–968.

    Article  Google Scholar 

  33. Petviashvili, V.I., The Great Red Spot of Jupiter and a drift soliton in plasma, Pis’ma Zh. Eksp. Teor. Fiz., 1980, vol. 32, no. 11, pp. 632–635.

    Google Scholar 

  34. Read, P.L. and Gierasch, P.J., Map** potential-vorticity dynamics on Jupiter. II: The Great Red Spot from Voyager 1 and 2 data, Q. J. R. Meteorol. Soc., 2006, vol. 132, pp. 1605–1625.

    Article  Google Scholar 

  35. Romanova, H.H. and Tseitlin, V.Yu., Quasi-geostrophic motions in barotropic and baroclinic fluids, Izv. Akad. Nauk SSSR: Fiz. Atmos. Okeana, 1984, vol. 20, no. 2, pp. 115–124.

    Google Scholar 

  36. Romanova, H.H. and Tseitlin, V.Yu., Solitary Rossby waves in a weakly stratified medium, Izv. Akad. Nauk SSSR: Fiz. Atmos. Okeana, 1985, vol. 21, no. 8, pp. 810–815.

    Google Scholar 

  37. Sagan, C., A truth table analysis of models of Jupiter’s great red spot, Comments Astrophys. Space Phys., 1971, vol. 3, pp. 65–72.

    Google Scholar 

  38. Sagdeev, R.Z., Shapiro, V.D., and Shevchenko, V.I., The Great Red Spot as a synoptic vortex in the Jovian atmosphere, Pis’ma Astron. Zh., 1981, vol. 7, no. 8, pp. 505–509.

    Google Scholar 

  39. Shetty, S. and Marcus, P.S., Changes in Jupiter’s Great Red Spot (1979–2006) and Oval BA (2000–2006), Icarus, 2010, vol. 210, no. 1, pp. 182–201.

    Article  Google Scholar 

  40. Shuleikin, V.V., Plane vortices with elliptical nuclei on Jupiter (Great Red Spot) and on the Earth, Astron. Zh., 1976, vol. 20, no. 4, pp. 850–859.

    Google Scholar 

  41. Simon, A.A., Tabataba-Vakili, F., Cosentino, R., Beebe, R.F., Wong, M.H., and Orton, G.S., Historical and contemporary trends in the size, drift, and color of Jupiter’s great red spot, Astron. J., 2018, vol. 155, no. 4, p. 151.

    Article  Google Scholar 

  42. Smoluchowski, R., Jupiter’s convection and its red spot, Science, 1970, vol. 168, no. 3937, pp. 1340–1342.

    Article  Google Scholar 

  43. Sommeria, J., Meyers, S.D., and Swinney, H.L., Laboratory simulation of Jupiter’s great red spot, Nature, 1988, vol. 331, no. 6158, pp. 689–693.

    Article  Google Scholar 

  44. Tikhomolov, E.M., Sustenance of vortex structures in a rotating fluid layer heated from below, Pis’ma Zh. Eksp. Teor. Phys., 1994, vol. 59, no. 3, pp. 155–158.

    Google Scholar 

  45. Volokitin, A.S. and Krasnosel’skikh, V.V., Rossby vortex as a model of the Jupiter planet Great Red spot, Dokl. Akad. Nauk SSSR, 1981, vol. 260, no. 3, pp. 588–591.

    Google Scholar 

  46. Williams, G.R., Ultra-long baroclinic waves and Jupiter’s Great Red Spot, J. Meteorol. Soc. Jpn., Ser. II, 1979, vol. 57, no. 2, pp. 196–198.

    Google Scholar 

  47. Williams, G.R., Jovian and comparative atmospheric modeling, Adv. Geophys., 1985, vol. 28, pp. 381–429.

    Article  Google Scholar 

  48. Williams, G.R., Jet sets, J. Meteorol. Soc. Jpn., Ser. II, 2003, vol. 81, no. 3, pp. 439–476.

    Google Scholar 

  49. Williams, G.R., Morrison, P.J., and Vilasur Swaminathan, R., Jovian vortices and jets, Fluids, 2019, vol. 4, no. 2, p. 104.

    Article  Google Scholar 

  50. Wong, M.H., Marcus, P.S., Simon, A.A., de Pater, I., Tollefson, J.W., and Asay-Davis, X., Evolution of the horizontal winds in Jupiter’s great red spot from one Jovian year of HST/WFC3 maps, Geophys. Res. Lett., 2021, vol. 48, no. 18, e2021GL093982.

  51. Yadav, R.K., Heimpel, M., and Bloxham, J., Deep convection-driven vortex formation on Jupiter and Saturn, Sci. Adv., 2020, vol. 6, no. 46, eabb9298.

  52. Yavorskaya, I.M. and Belyaev, Y.N., On a convective model of Jupiter, Acta Astronaut., 1982, vol. 9, nos. 6–7, pp. 481–486.

    Article  Google Scholar 

  53. Zhmur, V.V., Mezomasshtabnye vikhri okeana (Mesoscale Eddies of the Ocean) Moscow: GEOS, 2011.

  54. Zhmur, V.V. and Pankratov, K.K., Dynamics of an ellipsoidal near-surface eddy in an inhomogeneous flow, Okeanologiya, 1989, vol. 29, no. 2, pp. 205–211.

    Google Scholar 

Download references

Funding

This work was carried out as part of State Task no. 0128-2021-0002 and supported by the Russian Science Foundation, grant no. 22-17-00267.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to V. V. Zhmur.

Ethics declarations

The authors declare that they have no conflicts of interest.

APPENDIX

APPENDIX

1.1 SIGNIFICANCE ASSESSMENT OF THE β-EFFECT IN THE DESCRIPTION OF THE GRS

In our theory, we used the law of conservation of potential vorticity (2) without taking into account the β effect. If we consider the β effect, then the expression for the potential vorticity will take the form

$$\sigma = {{\Delta }_{h}}\psi + \frac{\partial }{{\partial z}}\frac{{{{f}^{2}}}}{{{{N}^{2}}}}\frac{{\partial \psi }}{{\partial z}} + ~\beta y,$$
(A1)

where y is the deviation of the liquid particle in the northern direction from its initial position. Parameter value \(\beta = 0.4742\,\, \times \,\,{{10}^{{ - 11}}}\) (m s)–1 for latitude 22°. The maximum deviation from a latitude of 22° to the north in different periods is \(b = \left( {6{\kern 1pt} - {\kern 1pt} 7} \right)\,\, \times \,\,{{10}^{3}}\) km. Hence, it follows that the change in potential vorticity due to the \(\beta \) effect in the extreme case, when the particle is deflected from a latitude of 22° to a distance \(b = 6\,\, \times \,\,{{10}^{3}}\) km to the north, will be

$${{\left( {\delta \sigma } \right)}_{{{\text{max}}}}} = \beta b = 2.84\,\, \times \,\,{{10}^{{ - 5}}}\,\,{{{\text{s}}}^{{ - 1}}}{\text{.}}$$
(A2)

In this case, we estimate the relative change in the potential vorticity in the core as

$$\frac{{{{{\left( {\delta \sigma } \right)}}_{{{\text{max}}}}}}}{{\sigma + {{\Gamma }}}} = \frac{{\beta b}}{{\sigma + {{\Gamma }}}} = \frac{{0.284\,\, \times \,\,{{{10}}^{{ - 4}}}~}}{{1.05\,\, \times \,\,{{{10}}^{{ - 4}}}}} = 0.27.$$
(A3)

Here, \(\left( {\sigma + {{\Gamma }}} \right) = 1.05\,\, \times \,\,{{10}^{{ - 4}}}\) s–1 is the experimental estimation of the potential vorticity of particles of the vortex core. This estimate is valid only for one particle that has shifted from a latitude of 22° and reached the northernmost point of the vortex core. On the average, over the volume of the ellipsoid, the particles deviate to the north by an amount of \(\frac{3}{8}b\) (this is the position of the center of gravity of the semiellipsoid). Then the average error over the volume of the nucleus

$$\left| {\delta \sigma } \right| = \frac{3}{8}\beta b = 1.07\,\, \times \,\,{{10}^{{ - 5}}}\,\,{{{\text{s}}}^{{ - 1}}}$$
(A4)

and the relative error in determining the value of the potential vorticity of the core when neglecting β effect on average will be the value

$$\frac{{\left| {\delta \sigma } \right|}}{{\sigma + {{\Gamma }}}} = \frac{{1.07\,\, \times \,\,{{{10}}^{{ - 5}}}}}{{1.05\,\, \times \,\,{{{10}}^{{ - 4}}}}} = 0.1.$$
(A4)

As we see, neglecting the \(\beta \) effect in terms of potential vorticity will give a maximum error of 27% and an average of 10% for vorticity. This is a significant amount. Therefore, all of the above GRS models use a more accurate approximation theory of \(\beta \) planes. The error of the applied theory in both the f-plane approximation and in the \(\beta \)-plane approximation is on the order of the Rossby number, the estimates of which at different times were in the range from 0.15 to 0.30. Thus, the error of both theories covers the additional effect from the refinement of the potential vorticity in the approximation β-plane. As a result, almost due to the error, it equalizes the approaches to the f- and \(\beta \)-planes. However, in fairness, we note that the \(\beta \)-plane approximation from a physical point of view describes a richer spectrum of phenomena, for example, the westward drift of a vortex and the Rossby waves that occur when flowing around a vortex. Neglecting the \(\beta \) effect, we discard these phenomena as well.

It should also be noted that with time the vortex decreases in size and intensifies in terms of potential vorticity values, and this process is accompanied by an increase in the Rossby number.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Suetin, B.P., Zhmur, V.V. & Chkhetiani, O.G. Baroclinic Model of Jupiter’s Great Red Spot. Izv. Atmos. Ocean. Phys. 59, 243–254 (2023). https://doi.org/10.1134/S0001433823030088

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S0001433823030088

Keywords:

Navigation