Log in

Closed 1/2-Elasticae in the 2-Sphere

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

We study critical trajectories in the sphere for the 1/2-Bernoulli’s bending functional with length constraint. For every Lagrange multiplier encoding the conservation of the length during the variation, we show the existence of infinitely many closed trajectories which depend on a pair of relatively prime natural numbers. A geometric description of these numbers and the relation with the shape of the corresponding critical trajectories is also given.

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 (Germany)

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Data Availability

The manuscript has no associated data.

Notes

  1. Actually, D. Bernoulli considered the unconstrained case, that is, \(\lambda =0\).

  2. The term hyperelliptic is used here in a broad sense, including also the rational and elliptic cases.

  3. In Miura and Yoshizawa (2022) there is no explicit mention to phase curves, but they are implicitly defined on page 27.

  4. Borrowing the terminology from the classical Euclidean geometry, s is an inflection point if \(\kappa (s)=0\) and \(\gamma \) is said convex if \(\kappa >0\) everywhere.

  5. In the sense specified in the Introduction.

References

  • Arroyo, J., Barros, M., Garay, O.J.: Willmore–Chen tubes on homogeneous spaces in warped product spaces. Pac. J. Math. 188–2, 201–207 (1999)

    MathSciNet  MATH  Google Scholar 

  • Arroyo, J., Garay, O.J., Mencía, J.J.: Closed generalized elastic curves in \(\textbf{S} ^2(1)\). J. Geom. Phys. 48–2, 339–353 (2003)

    MATH  Google Scholar 

  • Arroyo, J., Garay, O.J., Pámpano, A.: Constant mean curvature invariant surfaces and extremals of curvature energies. J. Math. Anal. Appl. 462, 1644–1668 (2018)

    MathSciNet  MATH  Google Scholar 

  • Arroyo, J., Garay, O.J., Pámpano, A.: Delaunay surfaces in \({\mathbb{S} }^3(\rho )\). Filomat 33–4, 1191–1200 (2019)

    MATH  Google Scholar 

  • Barros, M., Ferrández, A., Lucas, P., Meroño, M.A.: Willmore Tori and Willmore–Chen submanifolds in pseudo-Riemannian spaces. J. Geom. Phys. 28, 45–66 (1998)

    MathSciNet  MATH  Google Scholar 

  • Barros, M., Ferrández, A., Lucas, P.: Conformal tension in string theories and M-theory. Nucl. Phys. B 584, 719–748 (2000)

    MathSciNet  MATH  Google Scholar 

  • Bernoulli, J.: Curvatura laminae elasticae. Acta Eruditorum Lipsiae 262–276 (1694)

  • Blaschke, W.: Vorlesungen uber Differentialgeometrie und Geometrische Grundlagen von Einsteins Relativitatstheorie I–II: Elementare Differenntialgeometrie. Springer (1921–1923)

  • Byrd, P.F., Friedman, M.D.: Handbook of Elliptic Integrals for Engineers and Physicists. Springer-Verlag, Berlin (1954)

    MATH  Google Scholar 

  • Bohle, C., Peters, G.P., Pinkall, U.: Constrained Willmore surfaces. Calc. Var. Partial Differ. Equ. 32, 263–277 (2008)

    MathSciNet  MATH  Google Scholar 

  • Calabi, E., Olver, P.J., Tannenbaum, A.: Affine geometry, curve flows, and invariant numerical approximations. Adv. Math. 124, 154–196 (1996)

    MathSciNet  MATH  Google Scholar 

  • Calini, A., Ivey, T.: Integrable geometric flows for curves in the pseudoconformal \(\textbf{S} ^3\). J. Geom. Phys. 166, 104249 (2021)

    MATH  Google Scholar 

  • Calini, A., Ivey, T., Marí-Beffa, G.: Remarks on KdV-type flows on star-shaped curves. Phys. D 238–8, 788–797 (2009)

    MathSciNet  MATH  Google Scholar 

  • Calini, A., Ivey, T., Marí-Beffa, G.: Integrable flows for starlike curves in centroaffine space. SIGMA 9, 022 (2013)

    MathSciNet  MATH  Google Scholar 

  • Canham, P.B.: The minimum energy of bending as a possible explanation of the biconcave shape of the human red blood cell. J. Theor. Biol. 26–1, 61–76 (1970)

    Google Scholar 

  • Capovilla, R., Guven, J., Rojas, E.: Hamilton’s equations for a fluid membrane: axial symmetry. J. Phys. A: Math. Gen. 38, 8201–10 (2005)

    MathSciNet  MATH  Google Scholar 

  • Cho, M., Pember, J., Szewieczek, G.: Constrained Elastic Curves and Surfaces with Spherical Curvature Lines, ar**v: 2104.11058 [math.DG] (2021)

  • Chou, K.S., Qu, C.: The KdV equation and motion of plane curves. J. Phys. Soc. Jpn. 70, 1912–1916 (2001)

    MathSciNet  MATH  Google Scholar 

  • Chou, K.S., Qu, C.: Integrable equations arising from motions of plane curves. Phys. D 162, 9–33 (2002)

    MathSciNet  MATH  Google Scholar 

  • Chou, K.S., Qu, C.: Integrable equations arising from motions of plane curves II. J. Nonlinear Sci. 13–5, 487–517 (2003)

    MathSciNet  MATH  Google Scholar 

  • Euler, L.: De curvis elasticis. In: Methodus Inveniendi Lineas Curvas Maximi Minimive Propietate Gaudentes, Sive Solutio Problematis Isoperimetrici Lattissimo Sensu Accepti, Additamentum 1 Ser. 1, vol. 24, Lausanne (1744)

  • Evans, E.: Bending resistance and chemically induced moments in membrane bilayers. Biophys. J. 14, 923–931 (1974)

    Google Scholar 

  • Flash, T., Handzel, A.A.: Affine differential geometry analysis of human arm movements. Biol. Cybern. 96–6, 577–601 (2007)

    MathSciNet  MATH  Google Scholar 

  • Fomenko, A.T., Trofimov, V.V.: Geometric and Algebraic Mechanisms of the Integrability of Hamiltonian Systems on Homogeneous Spaces and Lie Algebras, Dinamical Systems, vol. 7. Springer-Verlag (1994)

  • Goldschmidt, H., Sternberg, S.: The Hamilton–Cartan formalism in the calculus of variations. Ann. Inst. Fourier 23, 203–267 (1973)

    MathSciNet  MATH  Google Scholar 

  • Goldstein, R.E., Petrich, D.M.: The Korteweg–de Vries hierarchy as dynamics of closed curves in the plane. Phys. Rev. Lett. 67, 3203 (1991)

    MathSciNet  MATH  Google Scholar 

  • Goldstein, R.E., Petrich, D.M.: Solitons, Euler’s equation and vortex patch dynamics. Phys. Rev. Lett. 69, 555 (1992)

    MathSciNet  MATH  Google Scholar 

  • Grant, J.D.E., Musso, E.: Coisotropic variational problems. J. Geom. Phys. 50, 303–338 (2004)

    MathSciNet  MATH  Google Scholar 

  • Guillemin, V., Sternberg, S.: Symplectic Techniques in Physics. Cambridge University Press, Cambridge (1990)

    MATH  Google Scholar 

  • Griffiths, P.A.: Exterior Differential Systems and the Calculus of Variations, Progress in Mathematics, vol. 25. Birkhauser, Boston (1982)

    Google Scholar 

  • Hasimoto, H.: Motion of a vortex filament and its relation to elastica. J. Phys. Soc. Jpn. 31, 293–294 (1971)

    Google Scholar 

  • Hasimoto, H.: A soliton on a vortex filament. J. Fluid Mech. 51, 477–485 (1972)

    MathSciNet  MATH  Google Scholar 

  • Helfrich, W.: Elastic properties of lipid bilayers: theory and possible experiments. Z. Natur. C 28, 693–703 (1973)

    Google Scholar 

  • Hsu, L.: Calculus of variations via the Griffiths formalism. J. Differ. Geom. 36, 551–589 (1992)

    MathSciNet  MATH  Google Scholar 

  • Jensen, G., Musso, E., Nicolodi, L.: The geometric Cauchy problem for the membrane shape equation. J. Phys. A Math. Theor. 47, 495201 (2014)

    MathSciNet  MATH  Google Scholar 

  • Jovanovic, B.: Noncommutative integrability and action-angle variables in contact geometry. J. Symplectic Geom. 10–4, 535–561 (2012)

    MathSciNet  MATH  Google Scholar 

  • Kida, S.: A vortex filament moving without changing of form. J. Fluid Mech. 112, 397–409 (1981)

    MathSciNet  MATH  Google Scholar 

  • Langer, J., Perline, R.: Poisson geometry of the filament equation. J. Nonlinear Sci. 1, 71–93 (1991)

    MathSciNet  MATH  Google Scholar 

  • Langer, J., Perline, R.: Curve motion inducing modified Korteweg–de Vries systems. Phys. Lett. A 239, 36–40 (1998)

    MathSciNet  MATH  Google Scholar 

  • Langer, J., Singer, D.A.: Liouville integrability of geometric variational problems. Comment. Math. Helv. 69, 272–280 (1994)

    MathSciNet  MATH  Google Scholar 

  • Langer, J., Singer, D.A.: The total squared curvature of closed curves. J. Diff. Geom. 20, 1–22 (1984)

    MathSciNet  MATH  Google Scholar 

  • Levien, R.: The Elastica: A Mathematical History, Technical Report No. UCB/EECS-2008-103, University of Berkeley (2008)

  • López, R., Pámpano, A.: Stationary soap films with vertical potentials. Nonlinear Anal. 215, 112661 (2022)

    MathSciNet  MATH  Google Scholar 

  • Miura, T., Yoshizawa, K.: Complete Classification of Planar p-Elasticae, Ar**v: 2203.08535 [math.AP] (2022)

  • Montaldo, S., Pámpano, A.: On the existence of closed biconservative surfaces in space forms. To appear in Commun. Anal. Geom. Ar**v: 2009.03233 [math.DG] (2020)

  • Montaldo, S., Oniciuc, C., Pámpano, A.: Closed biconservative hypersurfaces in spheres. J. Math. Anal. Appl. 518(1), 126697 (2023)

  • Musso, E.: Variational problems for plane curves in centro affine geometry. J. Phys A Math. Theor. 43, 1751–8113 (2010)

    MathSciNet  MATH  Google Scholar 

  • Musso, E.: Congruence Curves of the Goldstein–Petrich Flows, Harmonic Maps and Differential Geometry, Contemporary in Mathematics, vol. 542, pp. 99–113 (2011)

  • Musso, E.: Motions of curves in the projective plane inducing the Kaup–Kupershmidt hierarchy. SIGMA 8, 030 (2012)

    MathSciNet  MATH  Google Scholar 

  • Musso, E., Nicolodi, L.: Hamiltonian flows on null curves. Nonlinearity 23, 2117 (2010)

    MathSciNet  MATH  Google Scholar 

  • Musso, E., Salis, F.: The Cauchy–Riemann strain functional for Legendrian curves in the 3-sphere. Ann. Mat. Pura Appl. 199, 2395–2434 (2020)

    MathSciNet  MATH  Google Scholar 

  • Nakayama, K., Segur, H., Wadati, M.: Integrability and the motion of curves. Phys. Rev. Lett. 69, 2603 (1992)

    MathSciNet  MATH  Google Scholar 

  • Ortega, P., Ratiu, T.: Moment Maps and Hamiltonian Reductions, Progress in Mathematics, vol. 222. Birkhauser, Boston (2004)

    Google Scholar 

  • Pámpano, A.: Critical Tori for mean curvature energies in killing submersions. Nonlinear Anal. 200, 112092 (2020)

    MathSciNet  MATH  Google Scholar 

  • Pinkall, U.: Hopf Tori in \(\textbf{S} ^3\). Invent. Math. 81, 379–386 (1985)

    MathSciNet  MATH  Google Scholar 

  • Pinkall, U.: Hamiltonian flows on the space of star-shaped curves. Results Math. 27, 328–332 (1995)

    MathSciNet  MATH  Google Scholar 

  • Raviv, D., Kimmel, R.: Affine invariant geometry for non-rigid shapes. Int. J. Comput. Vis. 111–1, 1–11 (2015)

    MathSciNet  MATH  Google Scholar 

  • Soliman, Y., Chern, A., Diamanti, O., Knoppel, F., Pinkall, U., Schroeder, P.: Constrained Willmore surfaces. ACM Trans. Graph. 40–4, 112 (2021)

    Google Scholar 

  • Truesdell, C.: The Rational Mechanics of Flexible or Elastic Bodies: 1638–1788. Leonhard Euler, Opera Omnia, Birkhauser (1960)

  • Tu, Z.C., Ou-Yang, Z.C.: A geometric theory on the elasticity of bio-membranes. J. Phys. A Math. Gen. 37, 11407 (2004)

    MathSciNet  MATH  Google Scholar 

  • Vassilev, V.M., Djondjorov, P.A., Mladenov, I.M.: Cylindrical equilibrium shapes of fluid membranes. J. Phys. A Math. Theor. 41, 435201 (2008)

    MathSciNet  MATH  Google Scholar 

  • Verpoort, S.: Curvature functionals for curves in the equi-affine plane. Czechoslov. Math. J. 61, 419–435 (2011)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The first author is partially supported by PRIN 2017 “Real and Complex Manifolds: Topology, Geometry and Holomorphic Dynamics” (protocollo 2017JZ2SW5-004) and by the GNSAGA of INDAM. The present research was also partially supported by MIUR Grant “Dipartimenti di Eccellenza” 20182022, CUP: E11G18000350001, DISMA, Politecnico di Torino. The authors would like to thank the referees for their valuable comments which have helped to improve the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Álvaro Pámpano.

Additional information

Communicated by Anthony Bloch.

Publisher's Note

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

Appendices

Appendix A. The Curvature of the Extrema and the Complete Elliptic Integral \(\Psi _\lambda \)

This appendix has two parts. In the first one we will show how to build the \(\mu \)-invariant from incomplete elliptic integrals of the third kind and how to compute its least period in terms of complete elliptic integrals. In the second part we will decompose the integral \(\Psi _\lambda \) and compute its limits as \(e_1\) approaches the boundaries of its domain.

1.1 Part I: The Curvature of the Extrema

Let \(K(\phi ,\delta )\) and \(\Pi (\zeta ,\phi ,\delta )\) be the Legendre’s incomplete elliptic integrals of the first and third kind, defined as

$$\begin{aligned}{} & {} K(\phi ,\delta )=\int _0^{\phi } \frac{1}{\sqrt{1-\delta \sin ^2 (\theta )}}\,\textrm{d}\theta ,\\{} & {} \Pi (\zeta ,\phi ,\delta )= \int _0^{\phi } \frac{1}{(1-\zeta \sin ^2(\theta ))\sqrt{1-\delta \sin ^2(\theta )}}\,\textrm{d}\theta , \end{aligned}$$

and \(K(\delta )=K(\pi /2,\delta )\), \(\Pi (\zeta ,\delta )=\Pi (\zeta , \pi /2, \delta )\) be the corresponding complete elliptic integrals. Let \(\textrm{am}(u,\delta )\) be the Jacobi’s amplitude with parameter \(\delta \) and \(\textrm{sn}(u,\delta ) =\sin (\textrm{am}(u,\delta ))\) the associated Jacobi’s elliptic function. For simplicity, we denote by

$$\begin{aligned}{} & {} \alpha \equiv \alpha (\lambda ,e_1)=\frac{e_2-e_1}{e_2-e_4},\quad \beta \equiv \beta (\lambda ,e_1)=\frac{2}{\sqrt{(e_1-e_3)(e_2-e_4)}},\\{} & {} \qquad \qquad \qquad \delta \equiv \delta (\lambda ,e_1)=\frac{(e_1-e_2)(e_3-e_4)}{(e_1-e_3)(e_2-e_4)}, \end{aligned}$$

and

$$\begin{aligned} \zeta \equiv \zeta (\lambda ,e_1)=\frac{e_4(e_2-e_1)}{e_1(e_2-e_4)}, \end{aligned}$$

where \(e_1>e_2>0\) and \(e_3,e_4\) are the roots of the polynomial Q. Recall that \(e_2, e_3\) and \(e_4\) are functions of the fundamental parameters \(\lambda \) and \(e_1\). Let \(\mu \) be the solution of (4) with \(\mu (0)=e_2\) and \(\omega >0\) be its least period. By construction, \(\mu \) is strictly increasing on \([0,\omega /2]\) and \(\mu (\omega /2)=e_1\). Let \(h:[e_2,e_1]\rightarrow [0,\omega /2]\) be defined by

$$\begin{aligned} h(y)= \int _{e_2}^{y}\frac{1}{x\sqrt{-Q(x)}}\,\textrm{d}x = \frac{\omega }{2}-\int _y^{e_1}\frac{1}{x\sqrt{-Q(x)}}\,\textrm{d}x. \end{aligned}$$

Then, from (4) it follows that \(\mu |_{[0,\omega /2]}=h^{-1}\). Since \(\mu \) is even, this is enough to reconstruct \(\mu \) on the whole real axis. Using 257.12 and 340.04 of Byrd and Friedman (1954), we obtain

$$\begin{aligned} \int _y^{e_1}\frac{1}{x\sqrt{-Q(x)}}\,\textrm{d}x=\frac{\beta }{e_1}\left( \frac{\alpha }{\zeta }u(y)-\frac{\alpha -\zeta }{\zeta }\Pi (\zeta ,\textrm{am}(u(y),\delta ), \delta )\right) , \end{aligned}$$

where

$$\begin{aligned} u(y)=\textrm{sn}^{-1}\left( \sqrt{\frac{(e_2-e_4)(e_1-y)}{(e_1-e_2)(y-e_4)}},\delta \right) . \end{aligned}$$

Putting \(y=e_2\), we see that the least period of \(\mu \) is

$$\begin{aligned} \omega = 2\frac{\beta }{e_1}\left( \frac{\alpha }{\zeta }K(\delta )-\frac{\alpha -\zeta }{\zeta }\Pi (\zeta ,\delta ) \right) . \end{aligned}$$

Figure 14 reproduces the graphs of the h-function and of the \(\mu \)-function on the intervals \([e_2,e_1]\) and \([0,\omega ]\), when \(e_1=2\), \(e_2=1\), \(e_3=(-3+i\sqrt{23})/8\), \(e_4=\overline{e_3}\). The h-function is evaluated via the Mathematica library of elliptic functions while the \(\mu \) function is evaluated solving numerically (2) with initial conditions \(\mu (0)=e_2\) and \({\dot{\mu }}(0)=0\). The black-dashed portion of the graph of \(\mu \) on \([0,\omega /2]\) is obtained by symmetrizing the graph of h with respect to the bisector of the first quadrant, showing that the two methods are in agreement with each other.

Fig. 14
figure 14

On the left: the graph of the h-function for \(e_1=2\) and \(e_2=1\). On the right: the graph of the \(\mu \) function for the same values of \(e_1\) and \(e_2\)

The 1/2-Bernoulli’s bending energy of a B-string can also be evaluated in terms of the wave number and complete elliptic integrals of the first kind as

$$\begin{aligned} {{\mathcal {B}}}_{\lambda }(\gamma _{m, n})=2n\int _{e_2}^{e_1}\frac{1}{\sqrt{-Q(y)}}\,\textrm{d}y+n\lambda \omega =n\left( 2\beta K(\delta )+\lambda \omega \right) . \end{aligned}$$

1.2 Part II: The Complete Elliptic Integral \(\Psi _\lambda \)

Using (4) to make a change of variable in the definition of \(\Psi _\lambda \), (14), we have

$$\begin{aligned} \Psi _\lambda (e_1)= & {} 2\xi \int _0^\omega \frac{\mu ^2\left( \mu +2\lambda \right) }{1-4\xi ^2\mu ^2}\,\textrm{d}s\\= & {} 4\xi \int _{e_2}^{e_1}\frac{\mu \left( \mu +2\lambda \right) }{(1-4\xi ^2\mu ^2)\sqrt{-(\mu -e_1)(\mu -e_2)(\mu -e_3)(\mu -e_4)}}\,\textrm{d}\mu . \end{aligned}$$

This integral can be solved in terms of complete elliptic integrals of the first and third kind.

For simplicity, we denote by

$$\begin{aligned}{} & {} \zeta _+\equiv \zeta _+(\lambda ,e_1)=-\frac{(e_1-e_2)(-1+2e_4\xi )}{(e_2-e_4)(-1+2e_1\xi )}\,,\\{} & {} \zeta _-\equiv \zeta _-(\lambda ,e_1)=-\frac{(e_1-e_2)(1+2e_4\xi )}{(e_2-e_4)(1+2e_1\xi )}\,. \end{aligned}$$

We then have

$$\begin{aligned} \Psi _\lambda =2\pi \left( I+[1-\chi ] II+III\right) , \end{aligned}$$

where \(\chi \) is the indicator function of the exceptional locus \({\mathcal {P}}_*\) and,

$$\begin{aligned} I= & {} \frac{\beta }{4\pi \xi }\left( -2+\alpha \left[ \frac{-1-4\lambda \xi }{\zeta _+(-1+2e_1\xi )}+\frac{1-4\lambda \xi }{\zeta _-(1+2e_1\xi )}\right] \right) K(\delta )\,,\end{aligned}$$
(19)
$$\begin{aligned} II= & {} \frac{\beta (\alpha -\zeta _+)(1+4\lambda \xi )}{4\pi \zeta _+\xi (-1+2e_1\xi )}\,\Pi (\zeta _+,\delta )\,,\end{aligned}$$
(20)
$$\begin{aligned} III= & {} \frac{\beta (\alpha -\zeta _-)(-1+4\lambda \xi )}{4\pi \zeta _-\xi (1+2e_1\xi })\,\Pi (\zeta _-,\delta )\,. \end{aligned}$$
(21)

These formulas follow from three standard elliptic integrals. The first one (cf. 340.01 and 341.03 of Byrd and Friedman 1954) is

$$\begin{aligned} \int _0^{K(\delta )}\frac{1-a\, \textrm{sn}^2(u,\delta )}{1-b\, \mathrm{sn^2}(u,\delta )}\,\textrm{d}u=\frac{a}{b}K(\delta )-\frac{a-b}{b}\Pi (b,\delta ). \end{aligned}$$

The second elliptic integral (cf. 257 and 259 of Byrd and Friedman (1954)) is

$$\begin{aligned} \int _{e_2}^{e_1}\frac{\textrm{d}\mu }{\sqrt{-(\mu -e_1)(\mu -e_2)(\mu -e_3)(\mu -e_4)}}=\beta K(\delta )\,, \end{aligned}$$

where \(\beta \) and \(\delta \) are as above. The third relevant elliptic integral (cf. 257.39 and 259.04 of Byrd and Friedman 1954) is

$$\begin{aligned}{} & {} \int _{e_2}^{e_1}\frac{\textrm{d}\mu }{(p-\mu )\sqrt{-(\mu -e_1)(\mu -e_2)(\mu -e_3)(\mu -e_4)}}\\{} & {} \quad =\frac{\beta }{p-e_1}\int _0^{K(\delta )}\frac{1-\alpha \, \textrm{sn}^2(u,\delta )}{1-{\widetilde{\zeta }}\,\mathrm{sn^2}(u,\delta )}\,\textrm{d}u\\{} & {} \quad =\frac{\beta }{p-e_1}\left( \frac{\alpha }{{\widetilde{\zeta }}}K(\delta )-\frac{\alpha -{\widetilde{\zeta }}}{{\widetilde{\zeta }}}\Pi ({\widetilde{\zeta }},\delta )\right) , \end{aligned}$$

where \(p\ne e_1\), \(\alpha \), \(\beta \) and \(\delta \) are as above, and

$$\begin{aligned} {\widetilde{\zeta }}=\frac{(p-e_4)(e_1-e_2)}{(e_1-p)(e_2-e_4)}. \end{aligned}$$

We begin proving that \(\Psi _\lambda \rightarrow 2\pi p(\lambda )\) when \(e_1\) approaches \(\eta _\lambda \) from the right. By construction we have that \(e_2\rightarrow \eta _\lambda \) too, and, hence, it follows that

$$\begin{aligned} \left\{ \begin{aligned}&\lim _{e_1\rightarrow \eta _\lambda ^+} e_3(\lambda ,e_1)=\frac{-1+\sqrt{1-\eta _\lambda ^4}}{\eta _\lambda ^3} \\&\lim _{e_1\rightarrow \eta _\lambda ^+} e_4(\lambda ,e_1)=\frac{-1-\sqrt{1-\eta _\lambda ^4}}{\eta _\lambda ^3} \end{aligned}\right. , \end{aligned}$$

and

$$\begin{aligned} \lambda =\frac{1-\eta _\lambda ^4}{2\eta _\lambda ^3}\,. \end{aligned}$$

We now see that the coefficients in (19)–(21) tend to, respectively,

$$\begin{aligned} \lim _{e_1\rightarrow \eta _\lambda ^+} \alpha (\lambda ,e_1)=0\,,\quad \quad \quad \lim _{e_1\rightarrow \eta _\lambda ^+} \beta (\lambda ,e_1)=\frac{2\eta _\lambda }{\sqrt{3+\eta _\lambda ^4}}\,,\quad \quad \quad \lim _{e_1\rightarrow \eta _\lambda ^+} \delta (\lambda ,e_1)=0\,, \end{aligned}$$

while

$$\begin{aligned} \lim _{e_1\rightarrow \eta _\lambda ^+} \xi (\lambda ,e_1)=\frac{\sqrt{1+\eta _\lambda ^4}}{2\eta _\lambda ^3}\,,\quad \quad \quad \lim _{e_1\rightarrow \eta _\lambda ^+} \zeta _+(\lambda ,e_1)=0\,,\quad \quad \quad \lim _{e_1\rightarrow \eta _\lambda ^+} \zeta _-(\lambda ,e_1)=0\,, \end{aligned}$$

and

$$\begin{aligned}{} & {} \lim _{e_1\rightarrow \eta _\lambda ^+} \frac{\alpha }{\zeta _+}=\frac{\eta _\lambda ^4\left( \eta _\lambda ^2-\sqrt{1+\eta _\lambda ^4}\right) }{\sqrt{1+\eta _\lambda ^4}+\sqrt{1-\eta _\lambda ^8}+\eta _\lambda ^6},\,\,\\{} & {} \lim _{e_1\rightarrow \eta _\lambda ^+} \frac{\alpha }{\zeta _-}=-\frac{\eta _\lambda ^4\left( \eta _\lambda ^2+\sqrt{1+\eta _\lambda ^4}\right) }{\sqrt{1+\eta _\lambda ^4}+\sqrt{1-\eta _\lambda ^8}-\eta _\lambda ^6}\,. \end{aligned}$$

Finally, recalling that \(K(0)=\Pi (0,0)=\pi /2\), using the above limits and (15) we conclude that

$$\begin{aligned} \lim _{e_1\rightarrow \eta _\lambda ^+} \Psi _\lambda =-2\pi \sqrt{\frac{1+\eta _\lambda ^4}{3+\eta _\lambda ^4}}=2\pi p(\lambda ). \end{aligned}$$
(22)

In what follows, we prove the limit when \(e_1\rightarrow \infty \). This limit will depend on the sign of \(\lambda \). More precisely, we will see that

$$\begin{aligned} \lim _{e_1\rightarrow \infty } I(\lambda ,e_1)=\lim _{e_1\rightarrow \infty } III(\lambda ,e_1)=0\,, \end{aligned}$$

and

$$\begin{aligned} \left\{ \begin{aligned}&\lim _{e_1\rightarrow \infty } II(\lambda ,e_1)=-\frac{1}{2},&\quad \text {if}\,\lambda \ge 0\\&\lim _{e_1\rightarrow \infty } II(\lambda ,e_1)=\frac{1}{2},&\quad \text {if}\,\lambda <0 \end{aligned}\right. \,. \end{aligned}$$

In order to prove these limits we observe that, as \(e_1\rightarrow \infty \), the following asymptotic estimates hold true:

$$\begin{aligned}&e_2\sim 1/e_1\,,\quad e_3\sim -1/e_1\,,\quad e_4\sim -e_1\,,\quad \xi \sim e_1/2\,,\quad \beta \sim 2/e_1\,,\\&\delta \sim 1-4/e_1^2\,,\quad \alpha \sim (1-e_1^2)/(1+e_1^2)\,,\quad \zeta _-\sim 1-4/e_1^2\,, \end{aligned}$$

and

$$\begin{aligned} \left\{ \begin{aligned}&\zeta _+\sim 1-\frac{4\lambda ^2}{e_1^4},&\quad \text {if}\,\lambda \ne 0\\&\zeta _+\sim 1-\frac{1}{e_1^6},&\quad \text {if}\,\lambda =0 \end{aligned}\right. . \end{aligned}$$
(23)

Moreover, recall that \(K(\delta )\sim -\log (1-\delta )/2\) as \(\delta \rightarrow 1^-\) and so, in our case, we have

$$\begin{aligned} K\left( \delta (\lambda ,e_1)\right) \sim \frac{-1}{2}\log \frac{4}{e_1^2}, \end{aligned}$$

as \(e_1\rightarrow \infty \). Combining this and the above estimates we conclude that

$$\begin{aligned} I(\lambda ,e_1)\sim \frac{1}{\pi e_1^2}\log \frac{4}{e_1^2}, \end{aligned}$$

as \(e_1\rightarrow \infty \). This proves the first limit.

For the other limits we need some basic properties of the complete elliptic integral of the third kind \(\Pi (\zeta ,\delta )\). Let \(\Lambda =\{(\zeta ,\delta )\in [0,1)\times [0,1)\,|\,\zeta \ge \delta \}\) and consider the function

$$\begin{aligned} f:(\zeta ,\delta )\in \Lambda \longmapsto \frac{2}{\pi }\sqrt{1-\zeta }\sqrt{1-\delta }\,\Pi (\zeta ,\delta )\in {\mathbb {R}}\,. \end{aligned}$$

This function f is bounded below by \(2/\pi \) and above by 1. In addition, \(f(\zeta ,0)=1\) for every \(\zeta \in [0,1)\) and \(f(\zeta ,\delta )\rightarrow 1\) when \(\zeta \rightarrow 1^-\), for every value of \(\delta \in [0,1)\). Moreover, for every \(\zeta \in [0,1)\),

$$\begin{aligned} f(\zeta ,\zeta )=\frac{2}{\pi }\sqrt{1-\zeta }\,E\left( \frac{\zeta }{\zeta -1}\right) \end{aligned}$$

where E is the complete elliptic integral of the second kind. From this we infer that

$$\begin{aligned} \lim _{\zeta \rightarrow 1^-} f(\zeta ,\zeta )=\frac{2}{\pi }\,. \end{aligned}$$

From these properties we deduce the following facts:

  1. 1.

    If \(\gamma :(a,\infty )\longrightarrow \Lambda \) is a smooth curve such that \(\gamma (t)\rightarrow (1,1)\) when \(t\rightarrow \infty \) and \(\zeta =\delta \) is an asymptote of \(\gamma \) as \(t\rightarrow \infty \), then

    $$\begin{aligned} \lim _{t\rightarrow \infty }f(\gamma (t))=\frac{2}{\pi }\,. \end{aligned}$$
  2. 2.

    If \({\widetilde{\gamma }}:(a,\infty )\longrightarrow \Lambda \) is a smooth curve such that \({\widetilde{\gamma }}(t)\rightarrow (1,1)\) when \(t\rightarrow \infty \) and \(\zeta =1\) is an asymptote of \({\widetilde{\gamma }}\) as \(t\rightarrow \infty \), then

    $$\begin{aligned} \lim _{t\rightarrow \infty }f({\widetilde{\gamma }}(t))=1\,. \end{aligned}$$

Combining both things, it follows that as \(t\rightarrow \infty \),

$$\begin{aligned} \left\{ \begin{aligned}&\Pi (\gamma (t))\sim \frac{1}{\sqrt{1-\gamma _1(t)}\sqrt{1-\gamma _2(t)}}\\&\Pi ({\widetilde{\gamma }}(t))\sim \frac{\pi }{2\sqrt{1-{\widetilde{\gamma }}_1(t)}\sqrt{1-{\widetilde{\gamma }}_2(t)}} \end{aligned}\right. \,. \end{aligned}$$

In view of these properties, fix \(\tau _\lambda \) sufficiently large and consider the curves

$$\begin{aligned} \left\{ \begin{aligned}&\gamma _\lambda :e_1\in (\tau _\lambda ,\infty )\longmapsto (\zeta _-(\lambda ,e_1),\delta (\lambda ,e_1))\in \Lambda \\&{\widetilde{\gamma }}_\lambda :e_1\in (\tau _\lambda ,\infty )\longmapsto (\zeta _+(\lambda ,e_1),\delta (\lambda ,e_1))\in \Lambda \end{aligned}\right. \,. \end{aligned}$$

From above estimates when \(e_1\rightarrow \infty \), we have the following asymptotic behavior for \(\gamma _\lambda \),

$$\begin{aligned} \gamma _\lambda \sim \left( 1-\frac{4}{e_1^2},1-\frac{4}{e_1^2}\right) , \end{aligned}$$

while for \({\widetilde{\gamma }}_\lambda \), it depends on the value of \(\lambda \),

$$\begin{aligned} \left\{ \begin{aligned}&{\widetilde{\gamma }}_\lambda \sim \left( 1-\frac{4\lambda ^2}{e_1^4},1-\frac{4}{e_1^2}\right) ,&\quad \text {if}\,\lambda \ne 0\\&{\widetilde{\gamma }}_0\sim \left( 1-\frac{1}{e_1^6},1-\frac{4}{e_1^2}\right) ,&\quad \text {if}\,\lambda =0 \end{aligned}\right. \,. \end{aligned}$$

Thus, \(\gamma _\lambda \rightarrow (1,1)\) and \(\zeta =\delta \) is an asymptote of \(\gamma _\lambda \) as \(e_1\rightarrow \infty \). Similarly, \({\widetilde{\gamma }}_\lambda \rightarrow (1,1)\) as \(e_1\rightarrow \infty \). Hence, it follows from above facts that

$$\begin{aligned} \left\{ \begin{aligned}&\Pi \left( \zeta _-(\lambda ,e_1),\delta (\lambda ,e_1)\right) \sim \frac{e_1^2}{e_1^2-4}\\&\Pi \left( \zeta _+(\lambda \ne 0,e_1),\delta (\lambda \ne 0,e_1)\right) \sim \frac{\pi e_1^3}{8|\lambda |}\\&\Pi \left( \zeta _+(\lambda =0,e_1),\delta (\lambda =0,e_1)\right) \sim \frac{\pi e_1^4}{4}\\ \end{aligned}\right. \,, \end{aligned}$$

when \(e_1\rightarrow \infty \).

It is then clear, combining this and above estimates, that

$$\begin{aligned} \left\{ \begin{aligned}&\lim _{e_1\rightarrow \infty }III(\lambda ,e_1)=0\\&\lim _{e_1\rightarrow \infty }II(\lambda \ge 0,e_1)=-\frac{1}{2}\\&\lim _{e_1\rightarrow \infty }II(\lambda <0,e_1)=\frac{1}{2}\\ \end{aligned}\right. \,. \end{aligned}$$

This completes the proof about the claimed limits for \(\Psi _\lambda \) when \(e_1\rightarrow \infty \).

Appendix B. Closed 1/2-Elasticae in the Plane

We briefly comment about 1/2-elasticae in \({{\mathbb {R}}}^2\) in order to clarify some assertions made in the Introduction. We begin with the non-existence of closed convex 1/2-elasticae other than circles. In \({{\mathbb {R}}}^2\) the phase curves for convex 1/2-elasticae are the singular rational curves (see the picture on the left of Fig. 16)

$$\begin{aligned} {{\mathcal {C}}}_{e_1,e_2}\,:\, y^2+x^4\left( x^2-[e_1+e_2]x + e_1e_2\right) =0\,, \end{aligned}$$

where \(e_1>e_2>0\). Then, following the general argument of the Introduction, \(\mu \) is a solution of \({\dot{\mu }}^2+\mu ^4(\mu ^2-[e_1+e_2]\mu +e_1e_2)=0\) and \((\mu ,{\dot{\mu }})\) is a periodic, regular parameterization of the smooth component of the phase curve lying in the positive half-plane \({{\mathbb {H}}}^2=\{(x,y)\,|\, x>0\}\).

Integrating by quadratures, the arc-length parameterization of a critical curve, up to rigid motions, is

$$\begin{aligned} \gamma _{\lambda ,d}(s)=\sqrt{\frac{1}{d}}\left( \lambda s +\frac{1}{2}\int \mu (s)\textrm{d}s,-\frac{1}{2\mu (s)}\right) . \end{aligned}$$

Let \(\omega \) be the least period of \(\mu \). Then,

$$\begin{aligned} \gamma (\omega )-\gamma (0)=\sqrt{\frac{1}{d}}\left( \lambda \omega +\frac{1}{2}\int _0^{\omega }\mu (s)\textrm{d}s,0\right) . \end{aligned}$$

On the other hand,

$$\begin{aligned} \omega = 2\int _{e_2}^{e_1}\frac{\textrm{d}\mu }{\mu ^2\sqrt{-(\mu -e_1)(\mu -e_2)}}=\pi \frac{e_1+e_2}{(e_1e_2)^{3/2}}\,, \end{aligned}$$

and

$$\begin{aligned} \frac{1}{2}\int _0^{\omega }\mu (s)\textrm{d}s = \int _{e_2}^{e_1}\frac{\textrm{d}\mu }{\mu \sqrt{-(\mu -e_1)(\mu -e_2)}}=\frac{\pi }{\sqrt{e_1e_2}}\,. \end{aligned}$$

Hence

$$\begin{aligned} \gamma (\omega )-\gamma (0)=-\frac{\pi (e_1-e_2)}{(e_1e_2)^{3/2}}(1,0). \end{aligned}$$

This implies that the trajectory of \(\gamma _{\lambda ,d}\) is invariant by the subgroup generated by a non-trivial translation along the Ox-axis. In particular, it is unbounded.

Next we focus on the non-existence of non-convex critical curves with periodic curvature. In this case \(e_1>0>e_2\). By contradiction, suppose that \(\kappa \) is periodic and non-constant. Without loss of generality \(\kappa (0)=\textrm{max}(\kappa )>0\). Let \(J=(a,b)\), \(a<0<b\) be the connected component of \(\{s\in {{\mathbb {R}}}\,|\, \kappa (s)>0\}\) containing the origin. Since \(\kappa \) is not strictly positive, at least one among a or b is finite. Put \(\mu =\sqrt{\kappa |_{J}}\). Then, \(m:s\in J\rightarrow (\mu ,\mu ')\in {{\mathbb {H}}}^2\) is a parameterization of an open arc contained in \({{\mathcal {C}}}^+_{e_1,e_2}:= {{\mathcal {C}}}_{e_1,e_2}\cap {{\mathbb {H}}}^2\) (see the picture on the right of Fig. 15; \({{\mathcal {C}}}^+_{e_1,e_2}\) is represented in the black part). By construction, \(\mu (0)=e_1\) and, in addition, there exist \(\epsilon >0\) such that \(\mu '>0\) on \((-\epsilon ,0)\) and \(\mu '<0\) on \((0,\epsilon )\). Taking into account that \(\mu \) is a solution of the Euler–Lagrange equation, m is an integral curve of the vector field

$$\begin{aligned} \textbf{X}|_{(x,y)}=y\,\partial _{y}+\frac{1}{x}\left( 2y^2+\frac{1}{2}[e_1+e_2]x^5-x^6\right) \partial _{y}\in {{\mathfrak {X}}}({{\mathbb {H}}}^2)\,. \end{aligned}$$
Fig. 15
figure 15

Phase curves of convex planar 1/2-elasticae. On the left, \(e_1=2\) and \(e_2=1\); while, on the right, \(e_1=3\) and \(e_2=-2\) (Color Figure Online)

Fig. 16
figure 16

On the left, the functions \(h_+\) (black) and \(h_-\) (red). On the right, the graph of the \(\mu \)-invariant of a planar 1/2-elastic curve with \(e_1=3\) and \(e_2=-2\) (Color Figure Online)

Thus, m(s) cannot invert his motion along \({{\mathcal {C}}}^+_{e_1,e_2}\). Since \({{\mathcal {C}}}^+_{e_1,e_2}\) is homeomorphic to \({{\mathbb {R}}}\), this implies that m is a homeomorphism onto its image. Therefore, m((ab)) intersects the Ox-axis at one point, namely at \(m(0)=(e_1,0)\). Hence, \(\mu '>0\) on (a, 0) and \(\mu '<0\) on (0, b). Let \(h_{+}\) and \(h_{-}\) be the inverses of \(\mu |_{(a,0)}\) and \(\mu |_{(0,b)}\) respectively (see the picture on the left of Fig. 16; the graph of \(h_+\) is shown in black while the graph of \(h_-\) in red). Then,

$$\begin{aligned} \begin{aligned} h_{\pm }(m)&=\int _{e1}^{m}\frac{1}{\mu ^2\sqrt{-\mu ^2+(e_1+e_2)\mu -e_1e_2}}\textrm{d}\mu \\&\quad = \mp \frac{1}{(e_1^2e_2^2)^{3/4}m}\\&\qquad \times \, \left( \sqrt{e_1|e_2|(e_1-m)(m-e_2)} -(e_1+e_2)m\,\textrm{arctanh} \sqrt{\frac{|e_2|(e_1-m)}{e_1(m-e_2)}} \right) . \end{aligned}\end{aligned}$$

Consequently, \(\mu \) can be extended to a function \({\hat{\mu }}:{{\mathbb {R}}}\rightarrow (0,e_1]\) attaining its maximum at \(s=0\), strictly increasing on \((-\infty ,0)\) and strictly decreasing on \((0,+\infty )\) (see the picture on the right of Fig. 16) such that

$$\begin{aligned} \lim _{s \rightarrow \infty }{\hat{\mu }}(s)= \lim _{s \rightarrow +\infty }{\hat{\mu }}(s)=0\,. \end{aligned}$$

This implies \(a=-\infty \) and \(b=+\infty \), which is a contradiction.

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

Musso, E., Pámpano, Á. Closed 1/2-Elasticae in the 2-Sphere. J Nonlinear Sci 33, 3 (2023). https://doi.org/10.1007/s00332-022-09860-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00332-022-09860-3

Keywords

Mathematics Subject Classification

Navigation