Skip to main content
Log in

On the Interplay Between Vortices and Harmonic Flows: Hodge Decomposition of Euler’s Equations in 2d

  • Published:
Regular and Chaotic Dynamics Aims and scope Submit manuscript

Abstract

Let \(\Sigma\) be a compact manifold without boundary whose first homology is nontrivial. The Hodge decomposition of the incompressible Euler equation in terms of 1-forms yields a coupled PDE-ODE system. The \(L^{2}\)-orthogonal components are a “pure” vorticity flow and a potential flow (harmonic, with the dimension of the homology). In this paper we focus on \(N\) point vortices on a compact Riemann surface without boundary of genus \(g\), with a metric chosen in the conformal class. The phase space has finite dimension \(2N+2g\). We compute a surface of section for the motion of a single vortex (\(N=1\)) on a torus (\(g=1\)) with a nonflat metric that shows typical features of nonintegrable 2 degrees of freedom Hamiltonians. In contradistinction, for flat tori the harmonic part is constant. Next, we turn to hyperbolic surfaces (\(g\geqslant 2\)) having constant curvature \(-1\), with discrete symmetries. Fixed points of involutions yield vortex crystals in the Poincaré disk. Finally, we consider multiply connected planar domains. The image method due to Green and Thomson is viewed in the Schottky double. The Kirchhoff – Routh Hamiltonian given in C. C. Lin’s celebrated theorem is recovered by Marsden – Weinstein reduction from \(2N+2g\) to \(2N\). The relation between the electrostatic Green function and the hydrodynamic Green function is clarified. A number of questions are suggested.

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

Notes

  1. Extended version of talks given by one of the authors (JK) at two conferences in memory of Alexey Borisov: GDIS 2022, Zlatibor, Serbia, June 5–11 2022 and November 22–December 3, 2021, Steklov Mathematical Institute.

  2. How much this impacts the evolution will be discussed in the final section.

  3. So to speak, it is like the snail that carries its shell while it moves.

  4. “Fluid cohomology” (nice title!). It includes codes for the numerical implementations both in 2d and 3d and a beautiful (and impressive) video demonstration, https://yhesper.github.io/fc23/fc23.html.

  5. We refer to Arnold – Khesin [25], a comprehensive treatise about the topological-geometrical approach to Euler’s equations. Our final section, with questions, mentions recent work in 3d manifolds published by two groups, of Boris Khesin and of Eva Miranda, with their respective coauthors.

  6. See Appendix A with basic informations on Hodge theory and Appendix B on general versions of the well-known Biot – Savart formula in three dimensions. We should not confuse the Helmholtz – Hodge decomposition with Ladyzhenskaya’s for simply connected domains (on irrotational and solenoidal components), used in Chorin’s projection method [26].

  7. Aristotle already had the intuition that vorticity is what drives fluid motion. “Vortices are sinews and muscles of fluid motions” (Küchemann, 1965), see [27, 28]. In two dimensions, we will use the greek letter \(\psi=\psi_{\omega}\) to denote the stream function of \(\omega\) as a \(0\)-form, which is the traditional usage, and denote the corresponding \(2\)-form stream function by \(\Psi\). Then \(\Psi=\star\psi\).

  8. Boris Khesin (personal communication) warned us that, due to resulting coupled equations, mathematicians have mostly stayed with the space of \(\nu\) mod \(df\). Nonetheless, he and his coworkers used Hodge explicitly for 3d fluids [29] and implicitly in [30]. A recent survey by Peskin and collaborators [31] recognizes that their use of periodic boundary conditions in numerical simulations may introduce artifacts.

  9. There is a vast literature in physics and engineering exploring the Helmholtz – Hodge decomposition (see, e. g., [32, 33]) and more recently, also in Biomathematics [34]. Among recent applications we found: visualization and computer graphics, robotics, medical imaging and bio-engineering. The Hodge decomposition has been also relevant in fluids (oceanography, geophysics and astrophysics).

  10. See Section 3.9 for an additional discussion, suggested by one of the referees, about recovering the pressure.

  11. Topology breaking by motions starting impulsively were already described in Helmholtz’s 1858 paper [40]. Impulsive motions have been studied since Blasius and Lamb [41].

  12. A far-fetched analogy is the emergence of vortex pairs in 2-dimensional (i. e, thin) Bose – Einstein condensates, leading to the BKT transition and turbulence as they proliferate [42].

  13. An account on Riemann’s discovery of the bilinear period relations can be found in [45].

  14. It is not hard to show that the Hodge theorem extends to (3.13) and Euler’s equation will have the singular limit discussed here (see [1]), and for background Appendix A).

  15. He was probably the first to realize that a surface with a metric is also a Riemann surface by taking an atlas of local isothermal coordinates.

  16. We hope that experts on symplectic/Poisson structures will be intrigued by this situation. Such singular structures are now a fashionable theme (log-symplectic, b-Poisson, [49, 50]). We believe that the symplectic form (3.28) as \(\Gamma\rightarrow 0\) could be cast in these frameworks.

  17. Historical note. In the 20th century the physical and biological sciences have been revolutionized in probabilistic terms. Mathematicians began to look at differential equations, number theory and combinatorics in that light too. The applications expanded, among other areas, to financial mathematics and artificial intelligence. The most important person at the origin of probabilistic potential theory was Norbert Wiener. The seminal work is his paper from 1923 (one hundred years back from now), the first rigorous construction of a Brownian motion process [60]. Curiously, Brown made his experiment in 1827, one year before Green’s paper. The probabilistic analogies in electrostatics and ideal fluids are natural consequences of the fundamental role of the Laplacian in all these subjects. For the intuition in electrical networks in the discrete context (graphs and Markov chains), see [61], and the review in [62].

  18. The Robin mass relates with a spectral invariant, the Zeta function \(Z(p)=:{\rm Trace(\Delta}^{-p})=\sum_{j=1}^{\infty}\lambda_{j}^{-p},\mathop{\rm Re}p>1,\) that can be continued to a meromorphic function with a simple pole at \(p=1.\) \(\tilde{Z}=\lim_{p\to 1}\left(Z(p)-\frac{1}{p-1}\right)\) is the regularized trace. Morpurgo [57] proved that \(\tilde{Z}=\int_{\Sigma}R(s)\mu(s)+\frac{1}{2\pi}(\gamma-\log 2)\), \(\gamma\sim 0.5772\) (Euler constant). See also Jean Steiner [58].

  19. Vorticists is the name coined by H. Aref [75] for the community. An interesting artistic movement in the early XX century used the same name. See https://www.tate.org.uk/art/art-terms/v/vorticism.

  20. Various tori families are depicted in https://www.math.uni-tuebingen.de/user/nick/gallery/.

  21. The only constant mean curvature surface of genus zero is a round sphere.

  22. See animations in https://virtualmathmuseum.org/Surface/clifford_torus/clifford_torus.html.

  23. We found this M.Sc. thesis quite interesting and readable https://macsphere.mcmaster.ca/handle/11375/9044. We apologize for using the symbol \(\Gamma\) for these discrete groups, since it is traditional. We are sure that no confusion will arise in this section with the same symbol being used for vorticities.

  24. Translations move points along a geodesic at a constant speed. However, because of the failure of Euclid’s fifth postulate, most likely an the extra requirement is needed for a precise definition of a steady translating pattern.

  25. It is similar to the lift to the universal cover of an equilibrium position of a pair of opposite vortices on the Schottky double of a planar domain.

  26. One can make more general assumptions. All is needed is that the domain is finitely connected and no boundary component consists of just a single point. Any such domain can be mapped conformally onto a domain bounded by analytic curves. Nonetheless, to make sense to physical quantities, most authors assume them to be piecewise smooth, or just Lipschitz. It would be interesting to extend the discussion of this and the next two sections to curved surfaces (i. e., nonflat metrics) with boundaries, and the extension to dimensions \(\geqslant 3\) [14, 15, 16].

  27. The front and back faces are “morally” indistinguishable, being anticonformally equivalent. The construction goes back to F. Schottky [100]. See [101, 102] for general discussions.

  28. This formula will appear frequently in what follows. It can be viewed as a link between analysis and differential geometry. The same quantity is viewed in two different ways: \(i)\) integrating a function over a measure, \(|d\xi|\), defined on the (nonoriented) unit circle. \(ii)\) a differential form with respect to \(z\), evaluated along the oriented unit circle.

  29. Crowdy and Marshall use the name modified Green functions for those where the circulation is nonzero in one (only) of the \(g+1\) curves. Their work is highly recommended [115, 116, 117, 118, 119].

  30. Koebe constructs canonical conformal mappings by means of orthogonal series of analytic functions. From this he gets the hydrodynamic Green function with zero inner boundaries periods, which was used by C. C. Lin.

  31. It seems the term was coined in [122].

  32. There are actually a lot of indices and sums when expanding \(BQB^{\dagger}\).

  33. We know from Proposition 16 that \(Q=P^{-1}\), and \(P\) can be also called by the name of electrostatic capacity.

  34. “Stirrers”, also called “agitators” appear in studies of chaotic mixing, see, e. g., [124, 127, 125, 126].

  35. Could it be a pun? There were a lot of spy agents around in the war time when Lin was writing the paper.

  36. https://www.nist.gov/pml/sensor-science/fluid-metrology

  37. Lord Kelvin ([148], 1887): “The condition for steady motion of an incompressible inviscid fluid filling a finite fixed portion of space (that is to say, motion in which the velocity and direction of motion continue unchanged at every point of space in which the fluid is placed) is that, with given vorticity, the energy is a thorough maximum, or a thorough minimum, or a minimax. The further condition of stability is secured, by consideration of energy alone, for any case of steady motion for which the energy is a thorough maximum or a thorough minimum; because when the boundary is held fixed the energy is of necessity constant. But the mere consideration of energy does not decide the question of stability for any case of steady motion in which the energy is a minimax.”

  38. See the nice videos in https://www.youtube.com/user/thdrivas.

  39. We just mention (somewhat opportunistically): simply taking the Schottky double, these phenomena would also happen on a compact boundaryless surface.

  40. Maxwell’s Treatise on Electricity and Magnetism (1873, Vol. 1, Ch. XI, p. 245–283, 3rd Ed.) devotes 40 pages to the reflection techniques.

  41. http://www1.maths.leeds.ac.uk/pmtast/hyperbolic-surfaces/hypermodes.html, http://www1.maths.leeds.ac.uk/pmtast/publications/eigdata/datafile.html.

  42. See, e.g., the works by Yau and collaborators [167] and by Desbrun and associates http://fernandodegoes.org/.

REFERENCES

  1. Gustafsson, B., Vortex Pairs and Dipoles on Closed Surfaces, J. Nonlinear Sci., 2022, vol. 32, no. 5, Paper No. 62, 38 pp.

    Article  MathSciNet  Google Scholar 

  2. Grotta-Ragazzo, C., Errata and Addenda to: “Hydrodynamic Vortex on Surfaces” and “The Motion of a Vortex on a Closed Surface of Constant Negative Curvature”, J. Nonlinear Sci., 2022, vol. 32, no. 5, Paper No. 63, 10 pp.

    Article  MathSciNet  Google Scholar 

  3. Grotta Ragazzo, C., The Motion of a Vortex on a Closed Surface of Constant Negative Curvature, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci., 2017, vol. 473, no. 2206, 20170447, 17 pp.

    MathSciNet  Google Scholar 

  4. Bogatskiy, A., Vortex Flows on Closed Surfaces, J. Phys. A, 2019, vol. 52, no. 47, 475501, 23 pp.

    Article  MathSciNet  Google Scholar 

  5. Bogatskii, A., Vortex Flows on Surfaces and Their Anomalous Hydrodynamics, PhD Thesis, University of Chicago, Chicago, Ill., 2021, 47 pp.

  6. Marsden, J. and Weinstein, A., Coadjoint Orbits, Vortices, and Clebsch Variables for Incompressible Fluids, Phys. D, 1983, vol. 7, no. 1–3, pp. 305–323.

    Article  MathSciNet  Google Scholar 

  7. Hodge, W. V. D., The Theory and Applications of Harmonic Integrals, Cambridge: Cambridge Univ. Press, 1989.

    Google Scholar 

  8. Boatto, S. and Koiller, J., Vortices on Closed Surfaces, in Geometry, Mechanics, and Dynamics: The Legacy of Jerry Marsden, D. E. Chang, D. D. Holm, G. Patrick, T. Ratiu (Eds.), Fields Inst. Commun., vol. 73, New York: Springer, 2015, pp. 185–237.

    Chapter  Google Scholar 

  9. Lin, C. C., On the Motion of Vortices in Two Dimensions: 1. Existence of the Kirchhoff – Routh Function, Proc. Natl. Acad. Sci. USA, 1941, vol. 27, no. 12, pp. 570–575. Lin, C. C., On the Motion of Vortices in Two Dimensions: 2. Some Further Investigations on the Kirchoff – Routh Function, Proc. Natl. Acad. Sci. USA, 1941, vol. 27, no. 12, pp. 575–577. See also: Lin, C. C., On the Motion of Vortices in Two Dimensions, Univ. of Toronto Stud., Appl. Math. Ser., no. 5, Toronto, ON: Univ. of Toronto Press, 1943.

    Google Scholar 

  10. Gustafsson, B., On the Motion of a Vortex in Two-Dimensional Flow of an Ideal Fluid in Simply and Multiply Connected Domains, Bull. TRITA-MAT-1979-7, Stockholm: Royal Institute of Technology, 1979, 109 pp.

    Google Scholar 

  11. Flucher, M. and Gustafsson, B., Vortex Motion in Two-Dimensional Hydrodynamics, Bull. TRITA-MAT-1997-MA-02, Stockholm: Royal Institute of Technology, 1979, 24 pp.

    Google Scholar 

  12. Flucher, M., Vortex Motion in Two Dimensional Hydrodynamics, in Variational Problems with Concentration, Prog. Nonlinear Differ. Equ. Their Appl., vol. 36, Basel: Birkhäuser, 1999, pp. 131–149.

    Chapter  Google Scholar 

  13. Marsden, J. and Weinstein, A., Reduction of Symplectic Manifolds with Symmetry, Rep. Math. Phys., 1974, vol. 5, no. 1, pp. 121–130.

    Article  MathSciNet  Google Scholar 

  14. Friedrichs, K. O., Differential Forms on Riemannian Manifolds, Comm. Pure Appl. Math., 1955, vol. 8, pp. 551–590.

    Article  MathSciNet  Google Scholar 

  15. Schwarz, G., Hodge Decomposition: A Method for Solving Boundary Value Problems, Lect. Notes in Math., vol. 1607, Berlin: Springer, 1995.

    Book  Google Scholar 

  16. Morrey, Ch. B., Jr., A Variational Method in the Theory of Harmonic Integrals: 2, Amer. J. Math., 1956, vol. 78, pp. 137–170.

    Article  MathSciNet  Google Scholar 

  17. Razafindrazaka, F., Poelke, K., Polthier, K., and Goubergrits, L., A Consistent Discrete 3D Hodge-Type Decomposition: Implementation and Practical Evaluation, https://arxiv.org/abs/1911.12173 (16 Dec 2019).

  18. Saqr, K. M., Tupin, S., Rashad, S., Endo, T., Niizuma, K., Tominaga, T., and Ohta, M., Physiologic Blood Flow Is Turbulent, Sci. Rep., 2020, vol. 10, no. 1, 15492, 12 pp.

    Article  Google Scholar 

  19. Razafindrazaka, F. H., Yevtushenko, P., Poelke, K., Polthier, K., and Goubergrits, L., Hodge Decomposition of Wall Shear Stress Vector Fields Characterizing Biological Flows, R. Soc. Open Sci., 2019, vol. 6, no. 2, 181970, 14 pp.

    Article  MathSciNet  Google Scholar 

  20. Poelke, K. and Polthier, K., Boundary-Aware Hodge Decompositions for Piecewise Constant Vector Fields, Comput.-Aided Des., 2016, vol. 78, pp. 126–136.

    Article  MathSciNet  Google Scholar 

  21. Zhao, R., Debrun, M., Wei, G., and Tong, Y., 3D Hodge Decompositions of Edge- and Face-Based Vector Fields, ACM Trans. Graph., 2019, vol. 38, no. 6, Art. 181, 13 pp.

    Article  Google Scholar 

  22. Yin, H., Nabizadeh, M. S., Wu, B., Wang, S., and Chern, A., Fluid Cohomology, ACM Trans. Graph., 2023, vol. 42, no. 4, Art. 126, 25 pp.

    Article  Google Scholar 

  23. Arnold, V. I., Sur la géométrie différentielle des groupes de Lie de dimension infinie et ses applications à l’hydrodynamique des fluides parfaits, Ann. Inst. Fourier (Grenoble), 1966, vol. 16, no. 1, pp. 319–361.

    Article  MathSciNet  Google Scholar 

  24. Modin, K., Geometric Hydrodynamics: From Euler, to Poincaré, to Arnold, https://arxiv.org/abs/1910.03301 (2019).

  25. Arnold, V. I. and Khesin, B. A., Topological Methods in Hydrodynamics, Appl. Math. Sci., vol. 125, New York: Springer, 1998.

    Google Scholar 

  26. Chorin, A. J., Numerical Solution of the Navier – Stokes Equations, Math. Comp., 1968, vol. 22, no. 104, pp. 745–762.

    Article  MathSciNet  Google Scholar 

  27. Küchemann, D., Report on the I.U.T.A.M. Symposium on Concentrated Vortex Motions in Fluids, J. Fluid Mech., 1965, vol. 21, no. 1, pp. 1–20.

    Article  Google Scholar 

  28. Saffman, P. G., Vortex Dynamics, Cambridge Monogr. Mech. Appl. Math., New York: Cambridge Univ. Press, 1992.

    Google Scholar 

  29. Khesin, B., Kuksin, S., and Peralta-Salas, D., KAM Theory and the 3D Euler Equation, Adv. Math., 2014, vol. 267, pp. 498–522.

    Article  MathSciNet  Google Scholar 

  30. Khesin, B., Peralta-Salas, D., and Yang, Ch., The Helicity Uniqueness Conjecture in 3D Hydrodynamics, Trans. Amer. Math. Soc., 2022, vol. 375, no. 2, pp. 909–924.

    Article  MathSciNet  Google Scholar 

  31. Bao, Y., Donev, A., Griffith, B. E., McQueen, D. M., and Peskin, Ch. S., An Immersed Boundary Method with Divergence-Free Velocity Interpolation and Force Spreading, J. Comput. Phys., 2017, vol. 347, pp. 183–206.

    Article  MathSciNet  Google Scholar 

  32. Joseph, D. D., Helmholtz Decomposition Coupling Rotational to Irrotational Flow of a Viscous Fluid, Proc. Natl. Acad. Sci. USA, 2006, vol. 103, no. 39, pp. 14272–14277.

    Article  Google Scholar 

  33. Bhatia, H., Norgard, G., Pascucci, V., and Bremmer, P., Helmholtz – Hodge Decomposition: A Survey, IEEE Trans. Vis. Comput. Graph., 2012, vol. 19, no. 8, pp. 1386–1404.

    Article  Google Scholar 

  34. Lefèvre, J., Leroy, F., Khan, Sh., Dubois, J., Huppi, P. S., Baillet, S., and Mangin, J.-F., Identification of Growth Seeds in the Neonate Brain through Surfacic Helmholtz Decomposition, in Information Processing in Medical Imaging: Proc. of the 21st Internat. Conf. (IPMI, Williamsburg, Va., Jul 2009), J. L. Prince, D. L. Pham, K. J. Myers (Eds.), Lect. Notes in Comput. Sci., vol. 5636, Berlin: Springer, 2009, pp. 252–263.

    Chapter  Google Scholar 

  35. Marchioro, C. and Pulvirenti, M., Mathematical Theory of Incompressible Nonviscous Fluids, Appl. Math. Sci., vol. 96, New York: Springer, 1994.

    Google Scholar 

  36. Weis-Fogh, T., Quick Estimates of Flight Fitness in Hovering Animals, including Novel Mechanisms for Lift Production, J. Exp. Biol., 1974, vol. 59, pp. 169–230.

    Article  Google Scholar 

  37. Lighthill, M. J., On the Weis-Fogh Mechanism of Lift Generation, J. Fluid Mech., 1973, vol. 60, no. 1, pp. 1–17.

    Article  MathSciNet  Google Scholar 

  38. Kolomenskiy, D., Moffatt, H. K., Farge, M., and Schneider, K., The Lighthill – Weis – Fogh Clap-Fling-Sweep Mechanism Revisited, J. Fluid Mech., 2011, vol. 676, pp. 572–606.

    Article  MathSciNet  Google Scholar 

  39. Cheng, X. and Sun, M., Revisiting the Clap-and-Fling Mechanism in Small Wasp Encarsia formosa Using Quantitative Measurements of the Wing Motion, Phys. Fluids, 2019, vol. 31, no. 10, 101903.

    Article  Google Scholar 

  40. von Helmholtz, H., Über Integrale der hydrodynamischen Gleichungen, welche den Wirbelbewegungen entsprechen, J. Reine Angew. Math., 1858, vol. 55, pp. 25–55.

    MathSciNet  Google Scholar 

  41. Telionis, D. P., Impulsive Motion, in Unsteady Viscous Flows, Springer Ser. in Comput. Phys., Berlin: Springer, 1981, pp. 79-153.

    Chapter  Google Scholar 

  42. Kosterlitz, J. M. and Thouless, D. J., Early Work on Defect Driven Phase Transitions, Internat. J. Modern Phys. B, 2016, vol. 30, no. 30, 1630018, 59 pp. (See also: 40 Years of Berezinskii – Kosterlitz – Thouless Theory, J. V. Jose (Ed.), Singapore: World Sci., 2013.)

    Article  Google Scholar 

  43. Moffatt, H. K., Singularities in Fluid Mechanics, Phys. Rev. Fluids, 2019, vol. 4, no. 11, 110502, 11 pp.

    Article  Google Scholar 

  44. Farkas, H. M. and Kra, I., Riemann Surfaces, 2nd ed., Grad. Texts Math., vol. 71, New York: Springer, 1992.

    Book  Google Scholar 

  45. Chai, Ch.-L., The Period Matrices and Theta Functions of Riemann, in The Legacy of Bernhard Riemann after One Hundred and Fifty Years: Vol. 1, L. Ji, F. Oort, S.-T. Yau (Eds.), Adv. Lect. Math., vol. 35.1, Somerville, Mass.: Int. Press, 2016, pp. 79-106.

    Google Scholar 

  46. Okikiolu, K., A Negative Mass Theorem for the \(2\)-Torus, Comm. Math. Phys., 2008, vol. 284, no. 3, pp. 775–802.

    Article  MathSciNet  Google Scholar 

  47. Gustafsson, B., Vortex Motion and Geometric Function Theory: The Role of Connections, Philos. Trans. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci., 2019, vol. 377, no. 2158, 20180341, 27 pp.

    MathSciNet  Google Scholar 

  48. Klein, F., Uber Riemann’s Theorie der algebraischen Functionen und ihrer Integrale: Eine Erganzung der gewohnlichen Darstellungen, Leipzig: Teubner, 1882.

    Google Scholar 

  49. Guillemin, V., Miranda, E., and Pires, A. R., Symplectic and Poisson Geometry on \(b\)-Manifolds, Adv. Math., 2014, vol. 264, pp. 864–896.

    Article  MathSciNet  Google Scholar 

  50. Geudens, S. and Zambon, M., Deformations of Lagrangian Submanifolds in Log-Symplectic Manifolds, Adv. Math., 2022, vol. 397, Paper No. 108202, 85 pp.

    Article  MathSciNet  Google Scholar 

  51. Kimura, Y., Vortex Motion on Surfaces with Constant Curvature, R. Soc. Lond. Proc. Ser. A Math. Phys. Eng. Sci., 1999, vol. 455, no. 1981, pp. 245–259.

    Article  MathSciNet  Google Scholar 

  52. Ragazzo, C. and Viglioni, H., Hydrodynamic Vortex on Surfaces, J. Nonlinear Sci., 2017, vol. 27, no. 5, pp. 1609–1640.

    Article  MathSciNet  Google Scholar 

  53. Holcman, D. and Schuss, Z., Escape through a Small Opening: Receptor Trafficking in a Synaptic Membrane, J. Statist. Phys., 2004, vol. 117, no. 5–6, pp. 975–1014.

    Article  MathSciNet  Google Scholar 

  54. Schuss, Z., The Narrow Escape Problem: A Short Review of Recent Results, J. Sci. Comput., 2012, vol. 53, no. 1, pp. 194–210.

    Article  MathSciNet  Google Scholar 

  55. Holcman, D. and Schuss, Z., The Narrow Escape Problem, SIAM Rev., 2014, vol. 56, no. 2, pp. 213–257.

    Article  MathSciNet  Google Scholar 

  56. Doyle, P. G. and Steiner, J., Spectral Invariants and Playing Hide-and-Seek on Surfaces, https://arxiv.org/abs/1710.09857 (2017).

  57. Morpurgo, C., Zeta Functions on \(S^{2}\), in Extremal Riemann Surfaces: Papers from the AMS Special Session (held at the Annual Meeting of the American Mathematical Society in San Francisco, Calif., Jan 1995), J. R. Quine, P. Sarnak (Eds.), Contemp. Math., vol. 201, Providence, R.I.: AMS, 1997, pp. 213–226.

    Chapter  Google Scholar 

  58. Steiner, J., A Geometrical Mass and Its Extremal Properties for Metrics on \(S^{2}\), Duke Math. J., 2005, vol. 129, no. 1, pp. 63–86.

    Article  MathSciNet  Google Scholar 

  59. Grotta-Ragazzo, C., Vortex on Surfaces and Brownian Motion in Higher Dimensions: Special Metrics, J. Nonlinear Sci., 2024, vol. 34, no. 2, Paper No. 31.

    Article  MathSciNet  Google Scholar 

  60. Wiener, N., Differential-Space, J. Math. and Phys., 1923, vol. 2, pp. 131–174.

    Article  MathSciNet  Google Scholar 

  61. Doyle, P. G. and Snell, J. L., Random Walks and Electrical Networks, Carus Math. Monogr., vol. 22, Washington, D.C.: Mathematical Association of America, 1984.

    Book  Google Scholar 

  62. Stolarksy, K. B., Review on “Random Walks and Electric Networks”, Am. Math. Mon., 1987, vol. 94, no. 2, pp. 202–205.

    Article  MathSciNet  Google Scholar 

  63. Lighthill, J., Introduction. Real and Ideal Fluids, in Laminar Boundary Layers, L. Rosenhead (Ed.), Oxford: Clarendon, 1963, pp. 1–45.

    Google Scholar 

  64. Howe, M., Vorticity and the Theory of Aerodynamic Sound, J. Eng. Math., 2001, vol. 41, no. 4, pp. 367–400.

    Article  MathSciNet  Google Scholar 

  65. Tkachenko, V. K., Stability of Vortex Lattices, JETP, 1966, vol. 23, no. 6, pp. 1049–1056; see also: Zh. Èksper. Teoret. Fiz., 1966, vol. 50, no. 6, pp. 1573-1585.

    Google Scholar 

  66. O’Neil, K. A., On the Hamiltonian Dynamics of Vortex Lattices, J. Math. Phys., 1989, vol. 30, no. 6, pp. 1373–1379.

    Article  MathSciNet  Google Scholar 

  67. Stremler, M. A. and Aref, H., Motion of Three Point Vortices in a Periodic Parallelogram, J. Fluid Mech., 1999, vol. 392, pp. 101–128.

    Article  MathSciNet  Google Scholar 

  68. Stremler, M., On Relative Equilibria and Integrable Dynamics of Point Vortices in Periodic Domains, Theor. Comput. Fluid Dyn., 2010, vol. 24, no. 1, pp. 25–37.

    Article  Google Scholar 

  69. Crowdy, D., On Rectangular Vortex Lattices, Appl. Math. Lett., 2010, vol. 23, no. 1, pp. 34–38.

    Article  MathSciNet  Google Scholar 

  70. Kilin, A. A. and Artemova, E. M., Integrability and Chaos in Vortex Lattice Dynamics, Regul. Chaotic Dyn., 2019, vol. 24, no. 1, pp. 101–113.

    Article  MathSciNet  Google Scholar 

  71. Green, Ch. C. and Marshall, J. S., Green’s Function for the Laplace – Beltrami Operator on a Toroidal Surface, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci., 2013, vol. 469, no. 2149, 20120479, 18 pp.

    MathSciNet  Google Scholar 

  72. Sakajo, T. and Shimizu, Y., Point Vortex Interactions on a Toroidal Surface, Proc. Roy. Soc. London Ser. A, 2016, vol. 472, no. 2191, 20160271, 24 pp.

    MathSciNet  Google Scholar 

  73. Sakajo, T., Vortex Crystals on the Surface of a Torus, Philos. Trans. Roy. Soc. A, 2019, vol. 377, no. 2158, 20180344, 17 pp.

    Article  MathSciNet  Google Scholar 

  74. Guenther, N.-E., Massignan, P., and Fetter, A. L., Superfluid Vortex Dynamics on a Torus and Other Toroidal Surfaces of Revolution, Phys. Rev. A, 2020, vol. 101, no. 5, 053606, 11 pp.

    Article  MathSciNet  Google Scholar 

  75. Borisov, A. V., Meleshko, V. V., Stremler, M., and van Heijst, G., Hassan Aref (1950–2011), Regul. Chaotic Dyn., 2011, vol. 16, no. 6, pp. 671–684.

    Article  MathSciNet  Google Scholar 

  76. Lin, Ch.-Sh. and Wang, Ch.-L., Elliptic Functions, Green Functions and the Mean Field Equations on Tori, Ann. of Math. (2), 2010, vol. 172, no. 2, pp. 911–954.

    Article  MathSciNet  Google Scholar 

  77. Willmore, T. J., Surfaces in Conformal Geometry, Ann. Global Anal. Geom., 2000, vol. 18, no. 3–4, pp. 255–264.

    Article  MathSciNet  Google Scholar 

  78. Marques, F. C. and Neves, A., Min-Max Theory and the Willmore Conjecture, Ann. of Math. (2), 2014, vol. 179, no. 2, pp. 683–782. Marques, F. C. and Neves, A., The Willmore Conjecture, https://arxiv.org/abs/1409.7664 (2014).

  79. Pinkall, U. and Sterling, I., Willmore Surfaces, Math. Intelligencer, 1987, vol. 9, no. 2, pp. 38–43.

    Article  MathSciNet  Google Scholar 

  80. Heller, L. and Pedit, F., Towards a Constrained Willmore Conjecture, in Willmore Energy and Willmore Conjecture, M. D. Toda (Ed.), Monogr. Res. Notes Math., Boca Raton, Fla.: CRC, 2018, pp. 119–138.

    Google Scholar 

  81. Barros, M., Equivariant Tori Which Are Critical Points of the Conformal Total Tension Functional, RACSAM. Rev. R. Acad. Cienc. Exactas Fís. Nat. Ser. A Mat., 2001, vol. 95, no. 2, pp. 249–258.

    MathSciNet  Google Scholar 

  82. Barros, M., Ferrńdez, A., and Garay, Ó. J., Equivariant Willmore Surfaces in Conformal Homogeneous Three Spaces, J. Math. Anal. Appl., 2014, vol. 409, no. 1, pp. 459–477.

    Article  MathSciNet  Google Scholar 

  83. Wente, H. C., Counterexample to a Conjecture of H. Hopf, Pacific J. Math., 1986, vol. 121, no. 1, pp. 193–243.

    Article  MathSciNet  Google Scholar 

  84. Abresch, U., Constant Mean Curvature Tori in Terms of Elliptic Functions, J. Reine Angew. Math., 1987, vol. 374, pp. 169–192.

    MathSciNet  Google Scholar 

  85. Andrews, B. and Li, H., Embedded Constant Mean Curvature Tori in the Three-Sphere, J. Differential Geom., 2015, vol. 99, no. 2, pp. 169–189.

    Article  MathSciNet  Google Scholar 

  86. Hauswirth, L., Kilian, M., and Schmidt, M. U., Mean-Convex Alexandrov Embedded Constant Mean Curvature Tori in the \(3\)-Sphere, Proc. Lond. Math. Soc. (3), 2016, vol. 112, no. 3, pp. 588–622.

    Article  MathSciNet  Google Scholar 

  87. Lawson, H. Blaine, Jr., Complete Minimal Surfaces in \(S^{3}\), Ann. of Math. (2), 1970, vol. 92, no. 3, pp. 335–374.

    Article  MathSciNet  Google Scholar 

  88. Penskoi, A. V., Generalized Lawson Tori and Klein Bottles, J. Geom. Anal., 2015, vol. 25, no. 4, pp. 2645–2666.

    Article  MathSciNet  Google Scholar 

  89. Pinkall, U., Hopf Tori in \(S^{3}\), Invent. Math., 1985, vol. 81, no. 2, pp. 379–386.

    Article  MathSciNet  Google Scholar 

  90. Mironov, A. E., On a Family of Conformally Flat Minimal Lagrangian Tori in \(CP^{3}\), Math. Notes, 2007, vol. 81, no. 3–4, pp. 329–337; see also: Mat. Zametki, 2007, vol. 81, no. 3, pp. 374-384.

    Article  MathSciNet  Google Scholar 

  91. Aref, H., Newton, P. K., Stremler, M. A., Tokieda, T., and Vainchtein, D., Vortex Crystals, Adv. Appl. Math., 2003, vol. 39, pp. 1–79.

    Google Scholar 

  92. Koiller, J., Getting into the Vortex: On the Contributions of James Montaldi, J. Geom. Mech., 2020, vol. 12, no. 3, pp. 507–523.

    MathSciNet  Google Scholar 

  93. Bolza, O., On Binary Sextics with Linear Transformations into Themselves, Am. J. Math., 1887, vol. 10, no. 1, pp. 47–70.

    Article  MathSciNet  Google Scholar 

  94. Magnus, W., Noneuclidean Tesselations and Their Groups, Pure Appl. Math., vol. 61, New York: Acad. Press, 1974.

    Google Scholar 

  95. Balazs, N. L. and Voros, A., Chaos on the Pseudosphere, Phys. Rep., 1986, vol. 143, no. 3, pp. 109–240.

    Article  MathSciNet  Google Scholar 

  96. Gilman, J., Compact Riemann Surfaces with Conformal Involutions, Proc. Amer. Math. Soc., 1973, vol. 37, no. 1, pp. 105–107.

    Article  MathSciNet  Google Scholar 

  97. Schmutz Schaller, P., Involutions and Simple Closed Geodesics on Riemann Surfaces, Ann. Acad. Sci. Fenn. Math., 2000, vol. 25, no. 1, pp. 91–100.

    MathSciNet  Google Scholar 

  98. Haas, A. and Susskind, P., The Geometry of the Hyperelliptic Involution in Genus Two, Proc. Amer. Math. Soc., 1989, vol. 105, no. 1, pp. 159–165.

    Article  MathSciNet  Google Scholar 

  99. Costa, A. F. and Parlier, H., A Geometric Characterization of Orientation-Reversing Involutions, J. Lond. Math. Soc. (2), 2008, vol. 77, no. 2, pp. 287–298.

    Article  MathSciNet  Google Scholar 

  100. Schottky, F., Ueber die conforme Abbildung mehrfach zusammenhängender ebener Flächen, J. Reine Angew. Math., 1877, vol. 83, pp. 300–351.

    MathSciNet  Google Scholar 

  101. Schiffer, M. and Spencer, D. C., Functionals of Finite Riemann Surfaces, Princeton, N.J.: Princeton Univ. Press, 1954.

    Google Scholar 

  102. Hawley, N. S. and Schiffer, M. M., Riemann Surfaces Which Are Doubles of Plane Domains, Pacific J. Math., 1967, vol. 20, pp. 217–222.

    Article  MathSciNet  Google Scholar 

  103. Davis, Ph. J., The Schwarz Function and Its Applications, The Carus Math. Monogr., vol. 1, Buffalo, N.Y.: Mathematical Association of America, 1974.

    Google Scholar 

  104. Cohn, H., Conformal Mapping on Riemann Surfaces, Dover Books on Adv. Math., New York: Dover, 1980.

    Google Scholar 

  105. Gustafsson, B. and Roos, J., Partial Balayage on Riemannian Manifolds, J. Math. Pures Appl. (9), 2018, vol. 118, pp. 82–127.

    Article  MathSciNet  Google Scholar 

  106. Rodrigues, A. R., Castilho, C., and Koiller, J., On the Linear Stability of a Vortex Pair Equilibrium on a Riemann Surface of Genus Zero, Regul. Chaotic Dyn., 2022, vol. 27, no. 5, pp. 493–524.

    Article  MathSciNet  Google Scholar 

  107. Alling, N. L. and Greenleaf, N., Foundations of the Theory of Klein Surfaces, Lect. Notes in Math., vol. 219, Berlin: Springer, 1971.

    Google Scholar 

  108. Vanneste, J., Vortex Dynamics on a Möbius Strip, J. Fluid Mech., 2021, vol. 923, Paper No. A12, 12 pp.

    Article  MathSciNet  Google Scholar 

  109. Balabanova, N., Algebraic and Geometric Methods in Mechanics, PhD Dissertation, The University of Manchester, Manchester, UK, 2021, 178 pp.

  110. Gustafsson, B. and Tkachev, V. G., On the Exponential Transform of Multi-Sheeted Algebraic Domains, Comput. Methods Funct. Theory, 2011, vol. 11, no. 2, pp. 591–615.

    Article  MathSciNet  Google Scholar 

  111. Gustafsson, B. and Sebbar, A., Critical Points of Green’s Function and Geometric Function Theory, Indiana Univ. Math. J., 2012, vol. 61, no. 3, pp. 939–1017.

    Article  MathSciNet  Google Scholar 

  112. Krichever, I., Marshakov, A., and Zabrodin, A., Integrable Structure of the Dirichlet Boundary Problem in Multiply-Connected Domains, Comm. Math. Phys., 2005, vol. 259, no. 1, pp. 1–44.

    Article  MathSciNet  Google Scholar 

  113. Yamada, A., Positive Differentials, Theta Functions and Hardy \(H^{2}\) Kernels, Proc. Amer. Math. Soc., 1999, vol. 127, no. 5, pp. 1399–1408.

    Article  MathSciNet  Google Scholar 

  114. Green, G., Essay on the Application of Mathematical Analysis to the Theories of Electricity and Magnetism, Nottingham: Wheelhouse, 1828.

    Google Scholar 

  115. Crowdy, D. and Marshall, J., Green’s Functions for Laplace’s Equation in Multiply Connected Domains, IMA J. Appl. Math., 2007, vol. 72, no. 3, pp. 278–301.

    Article  MathSciNet  Google Scholar 

  116. Crowdy, D., Solving Problems in Multiply Connected Domains, CBMS-NSF Region. Conf. Ser. Appl. Math., vol. 97, Philadelphia, Penn.: Society for Industrial and Applied Mathematics (SIAM), 2020.

    Book  Google Scholar 

  117. Crowdy, D. and Marshall, J., Analytical Formulae for the Kirchhoff – Routh Path Function in Multiply Connected Domains, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci., 2005, vol. 461, no. 2060, pp. 2477–2501.

    MathSciNet  Google Scholar 

  118. Crowdy, D. G. and Marshall, J. S., The Motion of a Point Vortex around Multiple Circular Islands, Phys. Fluids, 2005, vol. 17, no. 5, 056602, 13 pp.

    Article  MathSciNet  Google Scholar 

  119. Crowdy, D., The Schottky – Klein Prime Function on the Schottky Double of Planar Domains, Comput. Methods Funct. Theory, 2010, vol. 10, no. 2, pp. 501–517.

    Article  MathSciNet  Google Scholar 

  120. Koebe, P., Abhandlungen zur Theorie der konformen Abbildung: 4. Abbildung mehrfach zusammenhängender schlichter Bereiche auf Schlitzbereiche, Acta Math., 1916, vol. 41, pp. 305–344.

    Article  MathSciNet  Google Scholar 

  121. Koebe, P., Abhandlungen zur Theorie der konformen Abbildung: 5. Abbildung mehrfach zusammenhängender schlichter Bereiche auf Schlitzbereiche (Fortsetzung), Math. Z., 1918, vol. 2, pp. 198–236.

    Article  MathSciNet  Google Scholar 

  122. Bandle, C. and Flucher, M., Harmonic Radius and Concentration of Energy; Hyperbolic Radius and Liouville’s Equations \(\Delta U=e^{U}\) and \(\Delta U=U^{(n+2)/(n-2)}\) SIAM Rev., 1996, vol. 38, no. 2, pp. 191–238.

    MathSciNet  Google Scholar 

  123. Ahlfors, L. V., Conformal Invariants: Topics in Geometric Function Theory, McGraw-Hill Ser. in Higher Math., New York: McGraw-Hill, 1973.

    Google Scholar 

  124. Aref, H., Stirring by Chaotic Advection, J. Fluid Mech., 1984, vol. 143, pp. 1–21.

    Article  MathSciNet  Google Scholar 

  125. Ottino, J. M., The Kinematics of Mixing: Stretching, Chaos, and Transport, Cambridge Texts in Appl. Math., Cambridge: Cambridge Univ. Press, 1989.

    Google Scholar 

  126. Daitche, A. and Tél, T., Dynamics of Blinking Vortices, Phys. Rev. E (3), 2009, vol. 79, no. 1, 016210, 9 pp.

    Article  MathSciNet  Google Scholar 

  127. Khakhar, D. V., Rising, H., and Ottino, J. M., Analysis of Chaotic Mixing in Two Model Systems, J. Fluid Mech., 1986, vol. 172, pp. 419–451.

    Article  MathSciNet  Google Scholar 

  128. Courant, R. and Hilbert, D., Methods of Mathematical Physics: Vol. 1, New York: Wiley, 1989.

    Book  Google Scholar 

  129. Vaskin, V. V. and Erdakova, N. N., On the Dynamics of Two Point Vortices in an Annular Region, Nelin. Dinam., 2010, vol. 6, no. 3, pp. 531–547 (Russian).

    Article  Google Scholar 

  130. Kurakin, L. G., Influence of Annular Boundaries on Thomson’s Vortex Polygon Stability, Chaos, 2014, vol. 14, no. 2, 023105, 12 pp.

    Article  MathSciNet  Google Scholar 

  131. Erdakova, N. N. and Mamaev, I. S., On the Dynamics of Point Vortices in an Annular Region, Fluid Dyn. Res., 2014, vol. 46, no. 3, 031420, 7 pp.

    Article  MathSciNet  Google Scholar 

  132. Flucher, M., Variational Problems with Concentration, Progr. Nonlinear Differ. Equ. Their Appl., vol. 36, Basel: Birkhäuser, 1999.

    Book  Google Scholar 

  133. Richardson, S., Vortices, Liouville’s Equation and the Bergman Kernel Function, Mathematika, 1980, vol. 27, no. 2, pp. 321–334.

    Article  MathSciNet  Google Scholar 

  134. Borah, D., Haridas, P., and Verma, K., Comments on the Green’s Function of a Planar Domain, Anal. Math. Phys., 2018, vol. 8, no. 3, pp. 383–414.

    Article  MathSciNet  Google Scholar 

  135. Solynin, A. Yu., A Note on Equilibrium Points of Green’s Function, Proc. Amer. Math. Soc., 2008, vol. 136, no. 3, pp. 1019–1021.

    Article  MathSciNet  Google Scholar 

  136. Gustafsson, B., On the Convexity of a Solution of Liouville’s Equation, Duke Math. J., 1990, vol. 60, no. 2, pp. 303–311.

    Article  MathSciNet  Google Scholar 

  137. Nehari, Z., Conformal Mapping, New York: McGraw-Hill, 1952.

    Google Scholar 

  138. Sario, L. and Oikawa, K., Capacity Functions, Grundlehren Math. Wiss., vol. 149, New York: Springer, 1969.

    Book  Google Scholar 

  139. Gianni, P., Seppälä, M., Silhol, R., and Trager, B., Riemann Surfaces, Plane Algebraic Curves and Their Period Matrices. Symbolic Numeric Algebra for Polynomials, J. Symbolic Comput., 1998, vol. 26, no. 6, pp. 789–803.

    Article  MathSciNet  Google Scholar 

  140. Luo, W., Error Estimates for Discrete Harmonic \(1\)-Forms over Riemann Surfaces, Comm. Anal. Geom., 2006, vol. 14, no. 5, pp. 1027–1035.

    Article  MathSciNet  Google Scholar 

  141. Nasser, M. M. S., Fast Computation of Hydrodynamic Green’s Function, Rev. Cuba Fís., 2015, vol. 32, no. 1, pp. 26–32.

    Google Scholar 

  142. Nasser, M., Fast Solution of Boundary Integral Equations with the Generalized Neumann Kernel, Electron. Trans. Numer. Anal., 2015, vol. 44, pp. 189–229.

    MathSciNet  Google Scholar 

  143. Yudovich, V. I., Eleven Great Problems of Mathematical Hydrodynamics, Mosc. Math. J., 2003, vol. 3, no. 2, pp. 711–737.

    Article  MathSciNet  Google Scholar 

  144. Khesin, B., Misiołek, G., and Shnirelman, A., Geometric Hydrodynamics in Open Problems, Arch. Ration. Mech. Anal., 2023, vol. 247, no. 2, Paper No. 15, 43 pp.

    Article  MathSciNet  Google Scholar 

  145. Yushutin, V., On Stability of Euler Flows on Closed Surfaces of Positive Genus, https://arxiv.org/abs/1812.08959 (2019).

  146. Davidson, P. A., Incompressible Fluid Dynamics, Oxford: Oxford Univ. Press, 2022.

    Google Scholar 

  147. Vladimirov, V. A. and Ilin, K. I., On Arnold’s Variational Principles in Fluid Mechanics, in The Arnoldfest: Proc. of a Conf. in Honour of V. I. Arnold for His Sixtieth Birthday (Toronto, ON, 1997), E. Bierstone, B. Khesin, A. Khovanskii, J. E. Marsden (Eds.), Fields Inst. Commun., vol. 24, Providence, R.I.: AMS, 1999, pp. 471–495.

    Google Scholar 

  148. Thomson, W., (1st Baron Kelvin), On the Stability of Steady and of Periodic Fluid Motion. Maximum and Minimum Energy in Vortex Motion, Philos. Mag. (5), 1887, vol. 23, no. 145, pp. 529–539.

    Article  Google Scholar 

  149. Khesin, B., Symplectic Structures and Dynamics on Vortex Membranes, Mosc. Math. J., 2012, vol. 12, no. 2, pp. 413–434, 461–462.

    Article  MathSciNet  Google Scholar 

  150. Izosimov, A. and Khesin, B., Characterization of Steady Solutions to the 2D Euler Equation, Int. Math. Res. Not. IMRN, 2017, vol. 2017, no. 24, pp. 7459–7503.

    MathSciNet  Google Scholar 

  151. Izosimov, A. and Khesin, B., and Mousavi, M., Coadjoint Orbits of Symplectic Diffeomorphisms of Surfaces and Ideal Hydrodynamics, Ann. Inst. Fourier (Grenoble), 2016, vol. 66, no. 6, pp. 2385–2433.

    Article  MathSciNet  Google Scholar 

  152. Izosimov, A. and Khesin, B., Classification of Casimirs in 2D Hydrodynamics, Mosc. Math. J., 2017, vol. 17, no. 4, pp. 699–716.

    Article  MathSciNet  Google Scholar 

  153. Iftimie, D., Lopes Filho, M. C., and Nussenzveig Lopes, H. J., Weak Vorticity Formulation of the Incompressible 2D Euler Equations in Bounded Domains, Comm. Partial Differential Equations, 2020, vol. 45, no. 2, pp. 109–145.

    Article  MathSciNet  Google Scholar 

  154. Dekeyser, J. and Van Schaftingen, J., Vortex Motion for the Lake Equations, Comm. Math. Phys., 2020, vol. 375, no. 2, pp. 1459–1501.

    Article  MathSciNet  Google Scholar 

  155. Grote, M. J., Majda, A. J., and Grotta Ragazzo, C., Dynamic Mean Flow and Small-Scale Interaction through Topographic Stress, J. Nonlinear Sci., 1999, vol. 9, no. 1, pp. 89–130.

    Article  MathSciNet  Google Scholar 

  156. Modin, K. and Viviani, M., A Casimir Preserving Scheme for Long-Time Simulation of Spherical Ideal Hydrodynamics, J. Fluid Mech., 2020, vol. 884, A22, 27 pp.

    Article  MathSciNet  Google Scholar 

  157. Shnirelman, A., On the Long Time Behavior of Fluid Flows, Procedia IUTAM, 2013, vol. 7, pp. 151–160.

    Article  Google Scholar 

  158. Yudovich, V. I., On the Loss of Smoothness of the Solutions of the Euler Equations and the Inherent Instability of Flows of an Ideal Fluid, Chaos, 2000, vol. 10, no. 3, pp. 705–719.

    Article  MathSciNet  Google Scholar 

  159. Morgulis, A., Shnirelman, A., and Yudovich, V., Loss of Smoothness and Inherent Instability of 2D Inviscid Fluid Flows, Comm. Partial Differential Equations, 2008, vol. 33, no. 4–6, pp. 943–968.

    Article  MathSciNet  Google Scholar 

  160. Kiselev, A. and Šverák, V., Small Scale Creation for Solutions of the Incompressible Two-Dimensional Euler Equation, Ann. of Math. (2), 2014, vol. 180, no. 3, pp. 1205–1220.

    Article  MathSciNet  Google Scholar 

  161. Samavaki, M. and Tuomela, J., Navier – Stokes Equations on Riemannian Manifolds, J. Geom. Phys., 2020, vol. 148, 103543, 15 pp.

    Article  MathSciNet  Google Scholar 

  162. Shashikanth, B. N., Dynamically Coupled Rigid Body-Fluid Flow Systems, Cham: Springer, 2021.

    Book  Google Scholar 

  163. Borisov, A., Mamaev, I. S., and Ramodanov, S. M., Coupled Motion of a Rigid Body and Point Vortices on a Two-Dimensional Spherical Surface, Regul. Chaotic Dyn., 2010, vol. 15, no. 4–5, pp. 440–461.

    Article  MathSciNet  Google Scholar 

  164. Avelin, H., Computations of Green’s Function and Its Fourier Coefficients on Fuchsian Groups, Experiment. Math., 2010, vol. 19, no. 3, pp. 317–334.

    Article  MathSciNet  Google Scholar 

  165. Jorgenson, J. and Kramer, J., Bounds on Canonical Green’s Functions, Compos. Math., 2006, vol. 142, no. 3, pp. 679–700.

    Article  MathSciNet  Google Scholar 

  166. Strohmaier, A. and Uski, V., An Algorithm for the Computation of Eigenvalues, Spectral Zeta Functions and Zeta-Determinants on Hyperbolic Surfaces, Comm. Math. Phys., 2013, vol. 317, no. 3, pp. 827–869.

    Article  MathSciNet  Google Scholar 

  167. Gu, X. D. and Yau, Sh.-T., Computational Conformal Geometry, Adv. Lect. in Math., vol. 3, Somerville, Mass.: International Press, 2008.

    Google Scholar 

  168. Dix, O. M. and Zieve, R. J., Vortex Simulations on a \(3\)-Sphere, Phys. Rev. Res., 2019, vol. 1, no. 3, 033201, 13 pp.

    Article  Google Scholar 

  169. DeTurck, D. and Gluck, H., Linking Integrals in the \(n\)-Sphere, Mat. Contemp., 2008, vol. 34, pp. 239–249.

    MathSciNet  Google Scholar 

  170. DeTurck, D. and Gluck, H., Electrodynamics and the Gauss Linking Integral on the \(3\)-Sphere and in Hyperbolic \(3\)-Space, J. Math. Phys., 2008, vol. 49, no. 2, 023504, 35 pp.

    Article  MathSciNet  Google Scholar 

  171. Parsley, R. J., The Biot – Savart Operator and Electrodynamics on Bounded Subdomains of the Three- Sphere, PhD Dissertation, University of Pennsylvania, Philadelphia,Penn., 2004, 131 pp.

  172. Arnol’d, V., Sur la topologie des écoulements stationnaires des fluides parfaits, C. R. Acad. Sci. Paris, 1965, vol. 261, pp. 17–20.

    MathSciNet  Google Scholar 

  173. Gromeka, I. S., Some Cases of Incompressible Fluid Motion, in Collected Papers, P. Ya. Kochina (Ed.), Moscow: AN SSSR, 1952, pp. 76–148 (Russian).

    Google Scholar 

  174. Dombre, T., Frisch, U., Greene, J. M., Hénon, M., Mehr, A., and Soward, A. M., Chaotic Streamlines in the ABC Flows, J. Fluid Mech., 1986, vol. 167, pp. 353–391.

    Article  MathSciNet  Google Scholar 

  175. Zhao, X. H., Kwek, K. H., Li, J. B., and Huang, K. L., Chaotic and Resonant Streamlines in the ABC Flow, SIAM J. Appl. Math., 1993, vol. 53, no. 1, pp. 71–77.

    Article  MathSciNet  Google Scholar 

  176. Galloway, D., ABC Flows Then and Now, Geophys. Astrophys. Fluid Dyn., 2012, vol. 106, no. 4–5, pp. 450–467.

    Article  Google Scholar 

  177. Etnyre, J. and Ghrist, R., Contact Topology and Hydrodynamics: 1. Beltrami Fields and the Seifert Conjecture, Nonlinearity, 2000, vol. 13, no. 2, pp. 441–458.

    Article  MathSciNet  Google Scholar 

  178. Etnyre, J. B. and Ghrist, R. W., Stratified Integrals and Unknots in Inviscid Flows, in Geometry and Topology in Dynamics (Winston-Salem, NC, 1998/San Antonio, TX, 1999), M. Barge, K. Kuperberg (Eds.), Contemp. Math., vol. 246, Providence, R.I.: AMS, 1999, pp. 99-111.

    Google Scholar 

  179. Etnyre, J. and Ghrist, R., Contact Topology and Hydrodynamics: 3. Knotted Orbits, Trans. Amer. Math. Soc., 2000, vol. 352, no. 12, pp. 5781–5794.

    Article  MathSciNet  Google Scholar 

  180. Cardona, R., Miranda, E., and Peralta-Salas, D., Computability and Beltrami Fields in Euclidean Space, J. Math. Pures Appl. (9), 2023, vol. 169, pp. 50–81.

    Article  MathSciNet  Google Scholar 

  181. Cardona, R., Miranda, E., Peralta-Salas, D., and Presas, F., Constructing Turing Complete Euler Flows in Dimension \(3\), Proc. Natl. Acad. Sci. USA, 2021, vol. 118, no. 19, e2026818118, 9 pp.

    Article  MathSciNet  Google Scholar 

  182. Enciso, A. and Peralta-Salas, D., Knots and Links in Steady Solutions of the Euler Equation, Ann. of Math. (2), 2012, vol. 175, no. 1, pp. 345–367. See also: Procedia IUTAM, 2013, vol. 7, pp. 13–20.

    MathSciNet  Google Scholar 

  183. Warner, F. W., Foundations of Differentiable Manifolds and Lie Groups, Grad. Texts in Math., vol. 94, New York: Springer, 1983.

    Google Scholar 

  184. Friedrichs, K. O., Differential Forms on Riemannian Manifolds, Comm. Pure Appl. Math., 1955, vol. 8, pp. 551–590.

    Article  MathSciNet  Google Scholar 

  185. Schwarz, G., Hodge Decomposition: A Method for Solving Boundary Value Problems, Lect. Notes in Math., vol. 1607, Berlin: Springer, 1995. viii, 164.

    Book  Google Scholar 

  186. Morrey, Ch. B., Jr., A Variational Method in the Theory of Harmonic Integrals: 2, Amer. J. Math., 1956, vol. 78, pp. 137–170.

    Article  MathSciNet  Google Scholar 

  187. Poelke, K. and Polthier, K., Boundary-Aware Hodge Decompositions for Piecewise Constant Vector Fields, Comput.-Aided Des., 2016, vol. 78, pp. 126–136.

    Article  MathSciNet  Google Scholar 

  188. Zhao, R., Debrun, M., Wei, G., and Tong, Y., 3D Hodge Decompositions of Edge- and Face-Based Vector Fields, ACM Trans. Graph., 2019, vol. 38, no. 6, Art. 181, 13 pp.

    Article  Google Scholar 

  189. Razafindrazaka, F., Poelke, K., Polthier, K., and Goubergrits, L., A Consistent Discrete 3D Hodge-Type Decomposition: Implementation and Practical Evaluation, https://arxiv.org/abs/1911.12173 (2019).

  190. Saqr, K. M., Tupin, S., Rashad, S., Endo, T., Niizuma, K., Tominaga, T., and Ohta, M., Physiologic Blood Flow Is Turbulent, Sci. Rep., 2020, vol. 10, no. 1, 15492, 12 pp.

    Article  Google Scholar 

  191. Razafindrazaka, F. H., Yevtushenko, P., Poelke, K., Polthier, K., and Goubergrits, L., Hodge Decomposition of Wall Shear Stress Vector Fields Characterizing Biological Flows, R. Soc. Open Sci., 2019, vol. 6, no. 2, 181970, 14 pp.

    Article  MathSciNet  Google Scholar 

  192. Glötzl, E. and Richters, O., Helmholtz Decomposition and Potential Functions for \(n\)-Dimensional Analytic Vector Fields, J. Math. Anal. Appl., 2023, vol. 525, no. 2, 127138, 19 pp.

    Article  MathSciNet  Google Scholar 

  193. Gustafsson, B., On Quadrature Domains and an Inverse Problem in Potential Theory, J. Analyse Math., 1990, vol. 55, pp. 172–216.

    Article  MathSciNet  Google Scholar 

  194. Shashikanth, B. N., Vortex Dynamics in \(R^{4}\), J. Math. Phys., 2012, vol. 53, no. 1, 013103, 21 pp.

    Article  MathSciNet  Google Scholar 

  195. Weyl, H., Die Idee der Riemannschen Fläche, R. Remmert (Ed.), Wiesbaden: Teubner, 2013.

    Google Scholar 

  196. Young, J., On the Cauchy Integral and Jump Decomposition, https://arxiv.org/abs/2301.12287 (2023).

  197. Solomentsev, E., Cauchy Integral, https://encyclopediaofmath.org/wiki/Cauchy_integral (see also: Encyclopaedia of Mathematics: Vol. 1, M. Hazewinkel (Ed.), Boston, Mass.: Springer, 1995).

Download references

ACKNOWLEDGMENTS

(JK) wishes to thank Vladimir Dragovic, Valery Kozlov, Ivan Mamaev, Zoran Rakic and Dmitry Treschev, the organizers of GDIS 2022, Zlatibor, Serbia, June 5–11 2022, and the organizers of the conference dedicated to the memory of Alexey Borisov, November 22 – December 3, 2021, Steklov Institute.

(BG) and (CGR) wish to thank Stefanella Boatto and other organizers of the conference in November 2012 in Rio de Janeiro, “N-vortex and N-body dynamics: common properties and approaches”. This is what made (BG) return to working on vortex motion after many years of absence from this field.

The three authors thank Boris Khesin, Edriss Titi and Albert Chern for references and insights, as well as the referee for many recommendations to clarify the text.

Funding

This work was supported by ongoing institutional funding. No additional grants to carry out or direct this particular research were obtained.

Author information

Authors and Affiliations

Authors

Contributions

All authors participated in discussing the results and in writing the article. BG was Salviati, CGR was Sagredo and JK was Simplicio.

Corresponding authors

Correspondence to Clodoaldo Grotta-Ragazzo, Björn Gustafsson or Jair Koiller.

Ethics declarations

The authors declare that they have no conflicts of interest.

Additional information

Dedicated to the memory of Alexey Borisov

PUBLISHER’S NOTE

Pleiades Publishing remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

MSC2010

76B47, 76M60, 34C23, 37E35

APPENDIX A. HODGE DECOMPOSITION

Hodge Theory for Compact Manifolds \(\Sigma^{n}\) Without Boundary

See [A1] for a complete exposition. A metric \(g\) in \(T\Sigma\) extends to tensors and forms via Gramians:

$$\langle\nu_{2},\nu_{2}\rangle_{Hodge}:=\int_{\Sigma}\nu_{1}\wedge\star\nu_{2},$$
(A.1)
where the operator \(\star\nu\) from \(k\) to \((n-k)\) forms such that \((\star)^{2}=(-1)^{k(n-k)}\) can be defined by
$$\xi\wedge\star\nu=g(\xi,\nu)\mu \quad(\mu=\text{volume form}).$$
(A.2)

Theorem (Hodge).

There is an orthogonal decomposition of any k-form \(\nu\in\Omega^{k}(\Sigma)\) as a sum

$$\nu=d\alpha\oplus\delta\Psi\oplus\eta,$$
$$\alpha\in\Omega^{k-1}(\Sigma),\quad\Psi\in\Omega^{k+1}(\Sigma),\quad\eta\in\text{ kernel}(\Delta),$$
  1. i)

    \(\delta:\Omega^{k+1}(\Sigma)\to\Omega^{k}(\Sigma)\), given by \(\delta=(-1)^{nk+1}\star d\star\) is the adjoint of \(d\): \(\langle d\nu_{1},\nu_{2}\rangle=\langle\nu_{1},\delta\nu_{2}\rangle\).

  2. ii)

    \(\Delta=d\delta+\delta d\) is an elliptic operator.

    $$\text{ kernel}(\Delta)={\rm Harm}^{k}(\Sigma)\textit{ is of finite dimension, and determines the cohomology}.$$
  3. iii)

    \(\Delta\eta=0\) iff both \(d\eta=0\) and \(\delta\eta=0\) and \(\left\{d\alpha\oplus\delta\omega\right\}={\rm Image}(\Delta)(\perp\quad{\rm to}\quad{\rm kernel}(\Delta)).\)

Schematically,

$$\Omega^{k}(\Sigma)={\rm im}d\oplus{\rm ker}\delta={\rm im}\delta\oplus{\rm ker}d={\rm im}d\oplus{\rm im}\delta\oplus{\rm ker\Delta}={\rm im}\Delta\oplus{\rm ker}\Delta.$$

Manifolds with Boundary: Hodge – Morrey – Friedrichs [A2, A3, A4]. For Computational Implementation, See [A5, A6]. For Application in Blood Physiology, See [A7, A8, A9]

We have not explored this situation in this paper. In order to show why the Hodge theorem has to be refined, observe that, if \(\Sigma\) has boundary, then it is possible for a 1-form to be in the kernel of the Laplacian without being both closed and co-closed.

A simple example: Let \(\Sigma=A\) be the annulus \(A:0<a\leqslant r\leqslant b\). The 2-form \(\omega=(1/2)\log(x^{2}+y^{2})dx\wedge dy\) is harmonic (since \(\log r\) is harmonic in the annulus). It is actually also exact: \(\omega=d[\int^{r}r\log r^{2} d\theta].\) Notice that it is not co-closed: \(\phi:=\delta\omega=(-ydx+xdy)/(x^{2}+y^{2})\)represents the 1-dimensional cohomology \({\rm Harm}^{1}(A)\).

Hodge Theory in 2d: Formalism

For brevity we will omit the wedge \(\wedge\) and sometimes denote by \(\mu\) the area form. The total area will be denoted by \(V\). All coordinate systems \((x,y)\) will be isothermal, with \(ds^{2}=\lambda^{2}(dx^{2}+dy^{2}).\)

We will circumvent the complex notation \(dz\) and \(d\bar{z}\), and write

$$\nu=\nu_{x}dx+\nu_{y}dy,\quad v=\nu^{\sharp}=(1/\lambda^{2})(\nu_{x}\partial_{x}+\nu_{y}\partial_{y}).$$
(A.3)

The main rules of the game are \(\star\nu=-v_{y}dx+\nu_{x}dy\) (90 degrees rotation), together with

$$\begin{gathered}\displaystyle\star 1=\lambda^{2}dx\wedge dy=\mu,\quad\star dx=dy,\quad\star dy=-dx,\quad\star\mu=1,\\ \displaystyle\star\star\nu=-\nu\ \text{(for 1-forms)},\quad\delta=-\star d\star\ \text{(any degree)},\quad\Delta=d\delta+\delta d\ \text{commutes with $d$ and $\delta$},\\ \displaystyle\Delta f=-\frac{1}{\lambda^{2}}(f_{xx}+f_{yy})\quad\text{(for functions)}.\end{gathered}$$

The following formulas are useful in computations with 1-forms:

$$\star\nu=i_{v}\mu,L_{v}=di_{v}+i_{v}d\Rightarrow d(\star\nu)=d(i_{v}\mu)=L_{v}\mu=:({\rm div}v)\mu.$$
(A.4)

All our vector fields will be incompressible: \(d\star\nu=0\). This assumption is used to say that locally

$$\star\nu=d\psi.$$
(A.5)

Proposition.

The inner product of \(1\)-forms requires just the complex structure.

Proof.

By definition, \(\langle\nu_{1},\nu_{2}\rangle=\int_{\Sigma}\nu_{1}\wedge\star\nu_{2}.\) Locally, \(\nu_{1}=a_{1}dx+a_{2}dy,\nu_{2}=b_{1}dx+b_{2}dy\), then

$$\langle\nu_{1},\nu_{2}\rangle=\int(a_{1}b_{1}+a_{2}b_{2})dxdy.$$
If one goes back to the vector fields, \(\nu_{1}^{\sharp}=(a_{1}\partial_{x}+a_{2}\partial_{y})/\lambda^{2},\nu_{2}^{\sharp}=(b_{1}\partial_{x}+b_{2}\partial_{y})/\lambda^{2}\) and
$$\langle\nu_{1},\nu_{2}\rangle=(a_{1}b_{1}+a_{2}b_{2})\lambda^{2}/\lambda^{4},$$
because \(|\partial_{x}|=\lambda\), etc. So
$$ \langle \nu_1^\sharp, \nu_2^\sharp \rangle = \int \frac{a_1 b_1 + a_2 b_2}{{\hspace{3mm}\big{/}}{\hspace{-3mm}\lambda^4}} {\hspace{2mm}\big{/}}{\hspace{-3mm}\lambda^2} {\hspace{2mm}\big{/}}{\hspace{-5mm}(\lambda^2} dx dy). $$
Thus, there is no need to refer to a specific metric for forms, although for the inner product between vector fields one has to specify the metric tensor. As we have seen, things miraculously compensate:

Proposition.

The harmonicity of \(1\)-forms, namely, the two conditions \(d\nu=0,d\star\nu=0\), does not depend on the chosen metric in the conformal class.

APPENDIX B. BIOT – SAVART

Two Dimensions

Recall that \(\Psi\) is the Hodge star of the Green function \(G^{\omega}\) (notations in Section 2.1) and the Green functions can in principle be constructed by potential theoretic methods (minimization of energy, Perron’s method of subharmonic functions, etc).

For 2-forms in dimension two one has \(\Delta = d \delta + {\hspace{2mm}\big{/}}{\hspace{-4mm}\delta d } = d \delta \), since the second term disappears by default: \(d\) is applied in the maximal dimension of the manifold.

Given a vorticity 2-form \(\omega\), write Poisson’s equation (for functions), \(\Delta\psi=\star\omega\) where \(\psi\) is a function. By abuse of language one can omit the \(\star\) in front of the 2-form \(\omega\) (so \(\omega\) is interpreted as a density multiplying \(\mu\), just to conform with the fluid mechanics notation).

The requirement for the solution \(\psi\) to exist is \(\omega\) to be exact (it is a vorticity 2-form)

$$\Bigl{(}\omega\text{ is exact }\Leftrightarrow\omega\text{ is }\perp\text{ to harmonic 2-forms (constant multiples of $\mu$)}\Leftrightarrow\int_{\Sigma}\omega=0\Bigr{)}.$$

Let \(\Psi=\star\psi=\psi\mu.\) Consider now \(\delta\Psi\) (which is purely vortical by definition). Then \(d (\delta \Psi) = \Delta \Psi - \delta {\hspace{2mm}\big{/}}{\hspace{-4mm} d \Psi} = \omega . \)

Note that the vortical part of the energy can be rewritten as

$$H_{vort}(\omega):=H(\delta\Psi)=\frac{1}{2}\langle\delta\Psi,\delta\Psi\rangle=\frac{1}{2}\langle d\delta\Psi,\Psi\rangle=\frac{1}{2}\langle\Delta\Psi,\Psi\rangle=\frac{1}{2}\langle\omega,\Delta^{-1}\omega\rangle.$$

Note that \(\Psi\) is not unique. One can add any \(\Psi_{o}\) such that \(\delta\Psi_{o}=0\). This means that \(\Psi_{o}\) is a constant multiple of \(\mu\). One may normalize \(\Psi=\Psi_{\omega}\) by requiring that it be orthogonal to all harmonic functions (namely, the constants), i. e., by requiring that

$$\int_{\Sigma}\Psi_{\omega}=0.$$
Then, by identification,
$$\nu=\delta\Psi_{\omega}+\eta=-\star d\star\Psi_{\omega}+\eta=-\star dG^{\omega}+\eta$$
so the Hodge star of \(\Psi_{\omega}\) is the Green function: \(G^{\omega}=\star\Psi_{\omega}\) with normalization
$$\int_{\Sigma}G^{\omega}\mu=0.$$
Extending the Hodge decomposition to certain singular forms gives, for example,
$$G(s,r)=G^{\delta_{r}\mu}(s)=\star\Psi_{\delta_{r}\mu}(s).$$

Biot – Savart in \({\mathbb{R}}^{n},n\geqslant 3\)

In dimension greater than two, for 2-forms, the term \(\delta d\) in \(\Delta=d\delta+\delta d\) no longer vanishes, so we cannot use Poisson’s equation anymore. The Biot – Savart law in \({\mathbb{R}}^{3}\) is taught in basic electromagnetism classes. For \({\mathbb{R}}^{4}\), see [B3]. For \({\mathbb{R}}^{n},n\geqslant 4\) we just found a very recent reference [B1]. One of the authors (BG) formulated a version of his own in [B2], Lemma 1.3.

Let \(f=(f_{1},\dots,f_{n})\) be a vector field in \({\mathbb{R}}^{n}\), vanishing sufficiently fast at infinity (in order for the convolutions to make sense). No regularity of \(f\) is assumed, we rather think of a distributional setting, allowing situations in which \(f\) is singular, such as having support on lower-dimensional manifolds (lines, surfaces, etc). Let

$$E(x)=\frac{c_{n}}{|x|^{n-2}}\quad(n\geqslant 3)$$
(B.1)
be the standard fundamental solution of the Laplacian, \(c_{N}\) chosen so that \(\Delta E=\delta_{o}\).

Proposition 3.

With \(\star=\star_{\rm conv}\) denoting convolution, we have the Biot – Savart type representation:

$$f=-({\rm div}f)\star\nabla E-({\rm curl}f)\star\nabla E.$$
(B.2)

The componentwise interpretation of this will be clear from the derivation below, where \(E_{j}=\partial E/\partial x_{j}\) \((\)the components of \(\nabla E)\). For each \(k=1,\dots,n\) we have, using that the derivative of a convolution can be moved to an arbitrary factor in it, and also that \(\partial E_{j}/\partial x_{k}=\partial E_{k}/\partial x_{j}\):

$$ \begin{aligned} f_k & =-(\Delta f_k)*E= -\sum_{j=1}^n \frac{\partial^2 f_k}{\partial x_j^2}\star E= -\sum_{j=1}^n \frac{\partial f_k}{\partial x_j}\star E_j \\ & = -\sum_{j=1}^n \frac{\partial}{\partial x_j}(f_k\star E_j)+\sum_{j=1}^n \frac{\partial}{\partial x_k}(f_j\star E_j)-\sum_{j=1}^n \frac{\partial}{\partial x_j}(f_j\star E_k)\\ & =-\sum_{j=1}^n \frac{\partial f_j}{\partial x_j}\star E_k-\sum_{j=1}^n \Big(\frac{\partial f_k}{\partial x_j}-\frac{\partial f_j}{\partial x_k}\Big)\star E_j. \end{aligned} $$
(B.3)
The \({\rm curl}\) of the vector field \(f\) is implicit in the second term of this formula. It is an antisymmetric tensor corresponding to the exterior differential acting on a one-form.

Case \(n=3\). The above reduces to the ordinary Biot – Savart law. We have \(f=\left(f_{1},f_{2},f_{3}\right)\):

$$ \begin{aligned} {\rm curl} f = & \left( \frac{\partial f}{\partial x_2} - \frac{\partial f}{\partial x_3} , \frac{\partial f}{\partial x_3} - \frac{\partial f}{\partial x_1} , \frac{\partial f}{\partial x_1} - \frac{\partial f}{\partial x_2} \right) \\ {\rm div} f = & \frac{\partial f}{\partial x_1} + \frac{\partial f}{\partial x_2} + \frac{\partial f}{\partial x_3}, \nabla E = (E_1,E_2,E_3) = - \frac{1}{4\pi} \frac{x}{|x|^3} \end{aligned} $$
as vector and scalar fields in \({\mathbb{R}}^{3}\). Spelling out formula (B.3) in this vector analysis language results in
$$f(x)=\frac{1}{4\pi}\int_{{\mathbb{R}}^{3}}({\rm div}f)(y)\frac{x-y}{|x-y|^{3}}d^{3}y+\frac{1}{4\pi}\int_{{\mathbb{R}}^{3}}({\rm curl}f)(y)\times\frac{x-y}{|x-y|^{3}}d^{3}y,$$
(B.4)
where \(d^{3}y=dy_{1}dy_{2}dy_{3}\) and \(\times\) denotes the ordinary vector product. When \(f\) is a stationary magnetic field, so that \({\rm div}f=0\) and \({\rm curl}f=\) current density, by Maxwell’s equations, the above formula agrees with Biot – Savart’s law found in textbooks. If \(f\) instead represents a stationary electric field, then \({\rm curl}f=0\) and we obtain Coulomb’s law for an extended charge distribution \({\rm div}f.\)

Formulation for One-Forms

Next, we wish to reformulate (B.3) in a way which in principle opens up for generalizations to curved manifolds. We then consider \(f\) as a differential form rather than a vector field. Thus,

$$f=f_{1}dx_{1}+\cdots+f_{n}dx_{n}.$$

Proposition.

We have

$$f=dE\star_{\rm conv}\delta f-i_{(dE)^{\sharp}}\star_{\rm conv}df.$$
(B.5)
The second term shall be interpreted as “interior convolution product”.

Proof.

Noting that, in our Cartesian context,

$$ \begin{aligned} df & = \sum_{k,j=1}^n \frac{\partial f_k}{\partial x_j} dx_j \wedge dx_k = \frac{1}{2} \left(\frac{\partial f_k}{\partial x_j} - \frac{\partial f_j}{\partial x_k} \right) dx_j \wedge dx_k \\ \delta f & = - d \star d \star f = - \sum_{k=1}^n \frac{\partial f_k}{\partial x_k } = - {\rm div} f \\ & \hspace{-3mm} i_{(dE)^{\sharp}} (dx_j \wedge dx_k ) = E_j dx_k - E_k dx_j, \end{aligned} $$
(B.6)

we have by (B.3):

$$ \begin{aligned} F &= - \delta f \star_{\rm conv} dE - \sum_{k,j=1}^n \left(\frac{\partial f_k}{\partial x_j} - \frac{\partial f_j}{\partial x_k} \right) \star_{\rm conv} E_j dx_k \\ &= \delta f \star_{\rm conv} dE - \frac{1}{2} \left(\frac{\partial f_k}{\partial x_j} - \frac{\partial f_j}{\partial x_k} \right) \star_{\rm conv} \left( E_j dx_k - E_k dx_j \right) \\ &= \delta f \star_{\rm conv} dE - \frac{1}{2} \left(\frac{\partial f_k}{\partial x_j} - \frac{\partial f_j}{\partial x_k} \right) \left( i_{(dE)^{\sharp}} \star_{\rm conv} dx_j \wedge dx_k \right) \\ &= dE \star_{\rm conv} \delta f - i_{(dE)^{\sharp}} \star_{\rm conv} df. \end{aligned} $$
(B.7)

For curved manifolds one has to replace the convolutions above, which can be interpreted as different kinds of Newtonian potentials, by expressions involving Green potentials. For example, with a \(0\)-form \(\rho,E\star_{\rm conv}\rho\) would be written \(E^{\rho}\), or \(G^{\rho}\) with \(G\) some Green function.

APPENDIX C. DIRECT PROOF OF LEMMA 3

We first show that the function

$$U_{\gamma}(s)=\oint_{\gamma}\star_{r}d_{r}G(r,s)\quad(\text{integral in }r)$$
(C.1)
is harmonic if \(s\not\in\gamma\). The Laplace operator with respect to \(s\) can be written as
$$\Delta_{s}G(r,s)=-\star_{s}d_{s}\star_{s}d_{s}G(r,s)\quad(\text{see Appendices A and B).}$$

The use of local isothermal coordinates \(s=(x,y)\) implies:

$$\mu_{s}=\lambda^{2}(x,y)dx\wedge dy,\quad\star\mu_{s}=1,\quad\Delta_{s}G(r,s)=-\lambda^{2}(x,y)\bigl{(}\partial_{x}^{2}+\partial_{y}^{2}\bigr{)}G(r,x,y).$$
Clearly,
$$\Delta_{s}\star_{r}d_{r}G(r,s)=\star_{r}d_{r}\Delta_{s}G(r,s){\rm if}r\neq s.$$

The symmetry of the Green function \(G(r,s)=G(s,r)\) and Eq. (3.10) implies that

$$\Delta_{s}G(r,s)=-V^{-1}{\rm for}r\neq s.$$
Therefore, if \(s\not\in\gamma\), then
$$\Delta_{s}U_{\gamma}(s)=\oint_{\gamma}\Delta_{s}\star_{r}d_{r}G(r,s)=\oint_{\gamma}\star_{r}d_{r}\Delta_{s}G(r,s)=-\oint_{\gamma}\star_{r}d_{r}(1/V)=0.$$
(C.2)

We now study \(U_{\gamma}(s)\) in a neighborhood of a point in \(\gamma\). We assume that \(\gamma\) is a real analytic curve. In this case, any \(s_{0}\in\gamma\) has a neighborhood \(N\) in which there are isothermal coordinates \((x,y):N\to{\mathbb{C}}\) such that

$$\gamma\cap N=\{|x|<a,y=0\}{\rm and}(x,y)(s_{0})=(0,0).$$

The orientation of \(\gamma\) coincides with the orientation of the \(x\)-axis. Let \(s\) be a point in \(N\) with coordinates \((x,y)(s)=(\xi,\eta)\) such that

$$U_{\gamma}(s)=\int_{\gamma\cap N}\star_{r}d_{r}G(r,s)+\int_{\gamma-\gamma\cap N}\star_{r}d_{r}G(r,s).$$
Since \(G(r,s)\) is analytic for \(r\in\{\gamma-\gamma\cap N\}\) and \(s\in N\), the second integral is an analytic function of \(s\). Therefore, the singular behavior of \(U_{\gamma}(s)\) is determined by
$$\int_{\gamma\cap N}\star_{r}d_{r}G(r,s)=\int_{-a}^{a}-\partial_{y}G(r,s)dx.$$
In the isothermal coordinates
$$G(r,s)=-\frac{1}{4\pi}\log\big{(}(x-\xi)^{2}+(y-\eta)^{2}\big{)}+h(x,y,\xi,\eta),$$
where \(h\) is an analytic function on \(N\). The singular behavior of \(U_{\gamma}(s)\) is, therefore, determined by the limit
$$ \begin{aligned} & \frac{1}{4\pi} \int_{-a}^a \partial_y \log\big((x-\xi)^2+(y-\eta)^2\big) \Big|_{y=0}dx \\ {}& \quad = - \frac{1}{2\pi} \int_{-a}^a\frac{\eta }{(x-\xi)^2+\eta^2}dx\\ {}& \quad = -\frac{1}{2\pi}\bigg\{ \arctan\left(\frac{a-\xi}{\eta}\right)+ \arctan \left(\frac{a+\xi}{\eta}\right)\bigg\}. \end{aligned} $$
We conclude that for \(s\in N\)
$$U_{\gamma}(\xi,\eta)=-\frac{1}{2\pi}\bigg{\{}\arctan\left(\frac{a-\xi}{\eta}\right)+\arctan\left(\frac{a+\xi}{\eta}\right)\bigg{\}}+U_{R}(\xi,\eta),$$
(C.3)
where \(U_{R}\) is an analytic function.

Let \((\xi,0)\) be a point in \(\gamma\cap N\), which implies \(|\xi|<a\), and make \((\xi,\eta)\to(\xi,0_{-})\) from the right-hand side of \(\gamma\), namely, with \(\eta<0\), then

$$U(\xi,0_{-})=\lim_{\eta\to 0_{-}}U_{\gamma}(\xi,\eta)=\frac{1}{2}+U_{R}(\xi,0).$$
(C.4)
If the limit is taken from the left-hand side of \(\gamma\), namely, with \(\eta>0\), then
$$U(\xi,0_{+})=\lim_{\eta\to 0_{+}}U_{\gamma}(\xi,\eta)=-\frac{1}{2}+U_{R}(\xi,0).$$
(C.5)

Therefore \(U(\xi,0_{-})-U(\xi,0_{+})=1\) and the function \(U_{\gamma}(s)\) jumps by one when \(\gamma\) is crossed from the left to the right. The differential of \(U_{\gamma}\) as given in Eq. (C.3) is \(dU_{R}\) plus

$$-2a\frac{2\xi\eta d\xi+\left(a^{2}-\xi^{2}+\eta^{2}\right)d\eta}{\left((a-\xi)^{2}+\eta^{2}\right)\left((a+\xi)^{2}+\eta^{2}\right)},$$
which is analytic over \(\gamma\). Therefore, \(dU_{\gamma}\) is a harmonic differential that is regular over \(\gamma\).     \(\square\)

This was presented just as a guide for the nonexpert. In different degrees of sophistication such an analysis can be found, for example, in Hermann Weyl’s classical 1913 treatise [C1], and it is also similar to jump formulas for Cauchy integrals [C2, C3]. The Cauchy integral of an analytic \(f\) in a domain \(D\) equals \(f\) inside the domain and zero outside, hence the jump across the boundary is exactly \(f\). With \(f=1\) we are integrating just the Cauchy kernel, and the singularity of this is of the same type as that of \(dG\) (or rather \(dG+i\star dG\)). In complex analysis, for the Cauchy integral, the jump formula often goes under the name Sokhotski – Plemelj jump formula (tracing back to 1868).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Grotta-Ragazzo, C., Gustafsson, B. & Koiller, J. On the Interplay Between Vortices and Harmonic Flows: Hodge Decomposition of Euler’s Equations in 2d. Regul. Chaot. Dyn. 29, 241–303 (2024). https://doi.org/10.1134/S1560354724020011

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S1560354724020011

Keywords

Navigation