Skip to main content
Log in

Leapfrogging Vortex Rings for the Three Dimensional Gross-Pitaevskii Equation

  • Manuscript
  • Published:
Annals of PDE Aims and scope Submit manuscript

Abstract

Leapfrogging motion of vortex rings sharing the same axis of symmetry was first predicted by Helmholtz in his famous work on the Euler equation for incompressible fluids. Its justification in that framework remains an open question to date. In this paper, we rigorously derive the corresponding leapfrogging motion for the axially symmetric three-dimensional Gross-Pitaevskii equation.

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

Similar content being viewed by others

Notes

  1. For \(y\in \mathbb {C}\) and \(z=(z_1,\ldots ,z_k) \in \mathbb {C}^k\) we write \((y,z):=\big (\mathrm{Re}(y\bar{z}_1),\ldots ,\mathrm{Re}(y\bar{z}_k)\big ) \in \mathbb {R}^k\) and \(y\times z := (iy,z)\).

  2. We refer to [5, 17] for some attempts in that direction, and an account of the difficulties.

  3. The integration is actually simpler in the original cartesian coordinates. A classical reference is the book of Jackson [10], an extended analysis can be found in the 1893 paper of Dyson [6]. See Appendix A for some details.

  4. Exact traveling waves having the form of vortex rings have been constructed in [4], these are very similar in shape but not exactly equal to reference vortex rings.

  5. A similar situation is described by Hicks [9] for a simplified vortex model introduced by Love [16] in 1894.

  6. We stress that this holds at the level of the system \((\text {LF})_\varepsilon \), we do not know whether such special solutions exist at the level of equation \((\text {GP})_\varepsilon ^{c}\).

  7. Another work on the 2D inhomogeneous GP equation is a recent preprint of Kurzke et al. [15], which studies a situation where the inhomogeneity and its derivatives are of order \(|\!\log \varepsilon |^{-1}.\) This is critical in the sense that interaction of vortices with the background potential and with each other are of the same order of magnitude. In the present work, by contrast, critical coupling occurs in hard-to-resolve corrections to the leading-order dynamics.

References

  1. Abramowitz, M., Stegun, I.: Handbook of Mathematical Functions, U.S. Government Printing Office, Washington D.C. (1964)

  2. Bethuel, F., Brezis, H., Hélein, F.: Ginzburg-Landau Vortices. Birkhäuser, Boston (1994)

    Book  MATH  Google Scholar 

  3. Bethuel, F., Gravejat, P., Smets, D.: Stability in the energy space for chains of solitons of the one-dimensional Gross-Pitaevskii equation. Ann. Inst. Fourier 64, 19–70 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  4. Bethuel, F., Orlandi, G., Smets, D.: Vortex rings for the Gross-Pitaevskii equation. J. Eur. Math. Soc. (JEMS) 6, 17–94 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  5. Benedetto, D., Caglioti, E., Marchioro, C.: On the motion of a vortex ring with a sharply concentrated vorticity. Math. Methods Appl. Sci. 23, 147–168 (2000)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  6. Dyson, F.W.: The potential of an anchor ring. Philos. Trans. R. Soc. Lond. A 184, 43–95 (1893)

    Article  ADS  MATH  Google Scholar 

  7. Helmholtz, H.: Über Integrale der hydrodynamischen Gleichungen welche den Wirbelbewegungen entsprechen. J. Reine Angew. Math. 55, 25–55 (1858)

    Article  MathSciNet  Google Scholar 

  8. Helmholtz, H.: (translated by P.G. Tait) On the integrals of the hydrodynamical equations which express vortex-motion. Phil. Mag. 33, 485–512 (1867)

  9. Hicks, W.M.: On the mutual threading of vortex rings. Proc. R. Soc. Lond. A 102, 111–131 (1922)

    Article  ADS  MATH  Google Scholar 

  10. Jackson, J.D.: Classical Electrodynamics. Wiley, New York (1962)

    MATH  Google Scholar 

  11. Jerrard, R.L., Smets, D.: Vortex dynamics for the two dimensional non homogeneous Gross-Pitaevskii equation. Annali Scuola Normale Sup. Pisa Cl. Sci. 14, 729–766 (2015)

    MathSciNet  MATH  Google Scholar 

  12. Jerrard, R.L., Soner, H.M.: The Jacobian and the Ginzburg-Landau energy. Calc. Var. Partial Differ. Equ. 14, 151–191 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  13. Jerrard, R.L., Spirn, D.: Refined Jacobian estimates for Ginzburg-Landau functionals. Indiana Univ. Math. J. 56, 135–186 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  14. Jerrard, R.L., Spirn, D.: Refined Jacobian estimates and Gross-Pitaevsky vortex dynamics. Arch. Ration. Mech. Anal. 190, 425–475 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  15. Kurzke, M., Marzuola, J.L., Spirn, D.: Gross-Pitaevskii vortex motion with critically-scaled inhomogeneities. SIAM J. Math. Anal. 49, 471–500 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  16. Love, A.E.H.: On the motion of paired vortices with a common axis. Proc. Lond. Math. Soc. 25, 185–194 (1893)

    Article  MathSciNet  MATH  Google Scholar 

  17. Marchioro, C., Negrini, P.: On a dynamical system related to fluid mechanics. NoDEA Nonlinear Differ. Equ. Appl. 6, 473–499 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  18. Martel, Y., Merle, F., Tsai, T.-P.: Stability and asymptotic stability in the energy space of the sum of N solitons for subcritical gKdV equations. Commun. Math. Phys. 231, 347–373 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The research of RLJ was partially supported by the National Science and Engineering Council of Canada under operating Grant 261955. The research of DS was partially supported by the Agence Nationale de la Recherche through the Project ANR-14-CE25-0009-01. Both authors wish to warmly thank a referee for his very careful reading of the manuscript and his judicious remarks.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Didier Smets.

Appendices

Vector Potential of Loop Currents

In the introduction we have considered the inhomogeneous Poisson equation

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle -\mathrm{div} \left( \frac{1}{r} \nabla \left( r A_a\right) \right) = 2\pi \delta _{a} &{}\qquad \text {in } \mathbb {H},\\ A_a = 0 &{} \qquad \text {on } \mathbb {H}. \end{array} \right. \end{aligned}$$

Its integration is classical (see e.g. [10]) and yields

$$\begin{aligned} A_a(r,z) = \frac{r(a)}{2} \int _0^{2\pi } \frac{\cos (t)}{\sqrt{r(a)^2+r^2 + (z-z(a))^2-2r(a)r\cos (t)}}\, dt, \end{aligned}$$

which in turn simplifies to

$$\begin{aligned} A_a(r,z) = \sqrt{\frac{r(a)}{r}} \frac{1}{k} \left[ (2-k^2)K(k^2)-2E(k^2)\right] \end{aligned}$$

where

$$\begin{aligned} k^2 = \frac{4r(a) r}{ r(a)^2 + r^2 + (z-z(a))^2 + 2 r(a) r} \end{aligned}$$

and where E and K denote the complete elliptic integrals of first and second kind respectively (see e.g. [1]). Note that \(A_{\lambda a}(\lambda r, \lambda z)= A_a(r,z)\) for any \(\lambda >0\) and that we have the asymptotic expansions [1] of the complete elliptic integrals as \(s\rightarrow 1\) :

$$\begin{aligned} K(s)= & {} -\frac{1}{2}\log (1-s)\left( 1 + \frac{1-s}{4}\right) + \log (4) + O(1-s),\\ E(s )= & {} 1 - \log (1-s)\frac{1-s}{4} + O(1-s), \end{aligned}$$

and similarly for their derivatives. For \((r,z) \in \mathbb {H}\setminus \{a\},\) direct computations therefore yield

$$\begin{aligned} A_a(r,z) = \left( \log \left( \tfrac{r(a)}{\rho }\right) + 3 \log (2) -2\right) + O\left( \tfrac{\rho }{r(a)} \left| \log \left( \tfrac{\rho }{r(a)}\right) \right| \right) \qquad \text {as } \tfrac{\rho }{r(a)} \rightarrow 0,\nonumber \\ \end{aligned}$$
(124)

and

$$\begin{aligned} \partial _{\rho } A_a = -\frac{1}{\rho } + O\left( \tfrac{1}{r(a)} \right) \qquad \text {as } \tfrac{\rho }{r(a)} \rightarrow 0, \end{aligned}$$
(125)

where \(\rho := |a-(r,z)|.\)

Concerning the asymptotic close to \(r=0,\) we have

$$\begin{aligned} A_a(r,z) \simeq \frac{rr(a)^2}{r(a)^3+|z|^3} \qquad \text {as } \frac{r}{r(a)} \rightarrow 0. \end{aligned}$$
(126)

Singular Unimodular Maps

When \(a = \{a_1,\ldots , a_n\}\) is a family of n distinct points in \(\mathbb {H}\), we define the function \(\Psi ^*_a\) on \(\mathbb {H}_{a}:=\mathbb {H}\setminus a\) by

$$\begin{aligned} \Psi ^*_a = \sum _{i=1}^n A_{a_i}, \end{aligned}$$

so that

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle -\mathrm{div} \left( \frac{1}{r} \nabla (r \Psi ^*_a)\right) = 2\pi \sum _{i=1}^n \delta _{a_i} &{}\qquad \text {on } \mathbb {H},\\ \displaystyle \Psi ^*_a = 0 &{} \qquad \text {on } \partial \mathbb {H}. \end{array} \right. \end{aligned}$$

Up to a constant phase shift, there exists a unique unimodular map \(u^*_a \in \mathcal {C}^\infty (\mathbb {H}_a,S^1)\cap W^{1,1}_\mathrm{loc}(\mathbb {H},S^1)\) such that

$$\begin{aligned} r (iu^*_a,\nabla u^*_a) = rj(u^*_a) = -\nabla ^\perp (r\Psi ^*_a). \end{aligned}$$

In the sense of distributions in \(\mathbb {H}\), we have

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \mathrm{div}(rj(u^*_a)) &{} = 0\\ \displaystyle \mathrm{curl}(j(u^*_a)) &{} = 2\pi \sum _{i=1}^n \delta _{a_i}. \end{array} \right. \end{aligned}$$

Let

$$\begin{aligned} \rho _a :=\frac{1}{4}\min \Big ( \min _{i\ne j}|a_i-a_j|, \min _i r(a_i)\Big ), \end{aligned}$$

for \(\rho \le \rho _a\) we set

$$\begin{aligned}\mathbb {H}_{a,\rho } := \mathbb {H}\setminus \cup _{i=1}^n B(a_i,\rho ). \end{aligned}$$

Lemma A.1

Under the above assumptions we have

$$\begin{aligned} \int _{\mathbb {H}_{a,\rho }} \frac{|j(u^*_a)|^2}{2} r\,drdz= & {} \pi \sum _{i=1}^n r(a_i) \times \Big [\log \big (\tfrac{r(a_i)}{\rho }\big )+ \sum _{j\ne i}A_{a_j}(a_i)\\&+ \big (3\log (2)-2\big ) + O\big (\tfrac{\rho }{\rho _a}\log ^2\big (\tfrac{\rho }{\rho _a}\big ) \big )\Big ]. \end{aligned}$$

Proof

We have the pointwise equality

$$\begin{aligned} |j(u^*_a)|^2 r = \frac{1}{r}|\nabla ^\perp (r\Psi ^*_a)|^2 = \frac{1}{r}|\nabla (r\Psi ^*_a)|^2, \end{aligned}$$

so that after integration by parts

$$\begin{aligned} \int _{\mathbb {H}_{a,\rho }} \frac{|j(u^*_a)|^2}{2} r\,drdz= & {} - \frac{1}{2} \int _{\mathbb {H}_{a,\rho }} \mathrm{div}\left( \frac{1}{r}\nabla (r\Psi ^*_a)\right) r\Psi ^*_a \,drdz\\&+ \frac{1}{2} \int _{\partial \mathbb {H}_{a,\rho }} \Psi ^*_a \nabla (r\Psi ^*_a)\cdot {\vec {n}}, \end{aligned}$$

and the first integral of the right-hand side in the previous identity vanishes by definition of \(\Psi ^*_a\) and \(\mathbb {H}_{a,\rho }.\) We next decompose the boundary integral as

$$\begin{aligned} \frac{1}{2} \int _{\partial \mathbb {H}_{a,\rho }} \Psi ^*_a \nabla (r\Psi ^*_a)\cdot {\vec {n}} = \frac{1}{2} \sum _{i,j,k=1}^n \int _{\partial B(a_i,\rho )} A_{a_j} \nabla (rA_{a_k})\cdot \vec {n}, \end{aligned}$$

and for fixed ijk we write

$$\begin{aligned} A_{a_j} \nabla (rA_{a_k})\cdot \vec {n} = \left( -A_{a_j} \partial _\rho A_{a_k} r + A_{a_j}A_{a_k}n_r\right) . \end{aligned}$$

Using (124), we have

$$\begin{aligned} \left| \int _{\partial B(a_i,\rho )} A_{a_j}A_{a_k}n_r \right| \le r(a_i)O\Big (\tfrac{\rho }{\rho _a}\log ^2(\tfrac{\rho }{\rho _a})\Big ). \end{aligned}$$

When \(i=j=k,\) we have by (124) and (125)

$$\begin{aligned} -\frac{1}{2} \int _{\partial B(a_i,\rho )} A_{a_j} \partial _\rho A_{a_k} r = \pi r(a_i) \left( \log \left( \frac{r(a_i)}{\rho }\right) + 3\log (2) -2 + O\Big (\tfrac{\rho }{\rho _a}\log (\tfrac{\rho }{\rho _a})\Big )\right) \end{aligned}$$

while when \(i=k\ne j\) we have

$$\begin{aligned} -\frac{1}{2} \int _{\partial B(a_i,\rho )} A_{a_j} \partial _\rho A_{a_k} r = \pi r(a_i) \left( A_{a_j}(a_i) + O\Big (\tfrac{\rho }{\rho _a}\Big )\right) . \end{aligned}$$

Finally, when \(i\ne k\) we have

$$\begin{aligned} \left| \frac{1}{2} \int _{\partial B(a_i,\rho )} A_{a_j} \partial _\rho A_{a_k} r \right| \le r(a_i) O\Big ((\tfrac{\rho }{\rho _a})^2 \log (\tfrac{\rho }{\rho _a})\Big ). \end{aligned}$$

The conclusion follows by summation. \(\square \)

If we next fix some constant \(K_0>0\) and we assume that the points \(a_i\) are of the form

$$\begin{aligned} a_{i} := \Big (r_0+ \frac{r(b_i)}{\sqrt{|\!\log \varepsilon |}} , z_0 + \frac{z(b_i)}{\sqrt{|\!\log \varepsilon |}} \Big ),\qquad i=1,\ldots ,n, \end{aligned}$$

for some \(r_0>0\), \(z_0\in \mathbb {R}\) and n points \(\{b_1,\ldots ,b_n\} \in \mathbb {R}^2\) which satisfy

$$\begin{aligned} \max _i |b_i| \le K_0,\qquad \text {and}\qquad \min _{i\ne j} \mathrm{dist}(b_i,b_j)\ge \frac{1}{K_0}, \end{aligned}$$

we directly deduce from Lemma A.1, (124) and (125):

Lemma A.2

Under the above assumptions we have

$$\begin{aligned}&\int _{\mathbb {H}_{a,\rho }} \frac{|j(u^*_a )|^2}{2} r\,drdz\nonumber \\&\quad =\ \pi n r_0 \Big ( |\log \rho | + n\log r_0 + n\big (3\log (2)-2\big ) + \tfrac{n-1}{2}\log |\log \varepsilon |\Big ) \nonumber \\&\quad \quad + \pi r_0 \Big ( \sum _{i} \frac{r(b_i)}{r_0}\frac{|\log \rho |}{\sqrt{|\log \varepsilon |}} - \sum _{i\ne j} \log |b_i-b_j|\Big )\nonumber \\&\quad \quad + O_{K_0,r_0} \Big ( \frac{1}{\sqrt{|\log \varepsilon |}}\Big ). \end{aligned}$$
(127)

Jacobian and Excess for 2D Ginzburg-Landau Functional

For the ease of reading, we recall in this appendix a few results from [12, 13] and [14] which we use in our work.

Theorem B.1

(Thm 1.3 in [13]—Lower energy bound) There exists an absolute constant \(C>0\) such that for any \(u \in H^1(B_r,\mathbb {C})\) satisfying \(\Vert Ju-\pi \delta _0\Vert _{\dot{W}^{-1,1}(B_r)} < r/4\) we have

$$\begin{aligned} {\mathcal E}_{\varepsilon }(u,B_r) \ge \pi \log \frac{r}{\varepsilon } + \gamma -\frac{C}{r} \left( \varepsilon \sqrt{\log \tfrac{r}{\varepsilon }} + \Vert Ju-\pi \delta _0\Vert _{\dot{W}^{-1,1}(B_r)}\right) . \end{aligned}$$

Theorem B.2

(from Thm 1.1 in [13]—Jacobian estimate without vortices) There exists an absolute constant \(C>0\) with the following property. If \(\Omega \) is a bounded domain, \(u \in H^1(\Omega ,\mathbb {C})\), \(\varepsilon \in (0,1]\) and \({\mathcal E}_{\varepsilon }(u,\Omega )< \pi |\!\log \varepsilon |\), then

$$\begin{aligned} \Big \Vert Ju \Big \Vert _{\dot{W}^{-1,1}(\Omega )} \le \varepsilon C {\mathcal E}_{\varepsilon }(u,\Omega )\exp \Big ( \frac{1}{\pi }{\mathcal E}_{\varepsilon }(u,\Omega )\Big ). \end{aligned}$$

Theorem B.3

(Thm 2.1 in [12]—Jacobian estimate with vortices) There exists an absolute constant \(C>0\) with the following property. If \(\Omega \) is a bounded domain, \(u \in H^1(\Omega , \mathbb {C})\), and \(\varphi \in \mathcal {C}^{0,1}_c(\Omega )\), then for any \(\lambda \in (1,2]\) and any \(\varepsilon \in (0,1)\),

$$\begin{aligned} \Big | \int _\Omega \varphi Ju \, dx \Big | \le \pi d_\lambda \Vert \varphi \Vert _\infty + \Vert \varphi \Vert _{\mathcal {C}^{0,1}}h^\varepsilon (\varphi ,u,\lambda ) \end{aligned}$$

where

$$\begin{aligned} d_\lambda = \Big \lfloor \frac{\lambda }{\pi } \frac{{\mathcal E}_{\varepsilon }(u,\mathrm{spt}(\varphi ))}{|\!\log \varepsilon |} \Big \rfloor , \end{aligned}$$

\(\lfloor x \rfloor \) denotes the greatest integer less than or equal to x, and

$$\begin{aligned} h^\varepsilon (\varphi ,u,\lambda ) \le C \varepsilon ^\frac{\lambda -1}{12\lambda }\big (1+\frac{{\mathcal E}_{\varepsilon }(u,\mathrm{spt}(\varphi ))}{|\!\log \varepsilon |}\big ) \big (1+\mathcal {L}^2(\mathrm{spt}(\varphi )\big ). \end{aligned}$$

Theorem B.4

(Thm 1.2’ in [14]—Jacobian localization for a vortex in a ball) There exists an absolute constant \(C>0\), such that for any \(u \in H^1(B_r,\mathbb {C})\) satisfying

$$\begin{aligned}\Vert Ju-\pi \delta _0\Vert _{\dot{W}^{-1,1}(B_r)} < r/4,\end{aligned}$$

if we write

$$\begin{aligned} \Xi = {\mathcal E}_{\varepsilon }(u,B_r) - \pi \log \frac{r}{\varepsilon } \end{aligned}$$

then there exists a point \(\xi \in B_{r/2}\) such that

$$\begin{aligned} \left\| Ju - \pi \delta _\xi \right\| _{\dot{W}^{-1,1}(B_r)} \le \varepsilon C(C+\Xi )\left[ (C+\Xi )e^{\Xi /\pi } + \sqrt{\log \frac{r}{\varepsilon }}\right] . \end{aligned}$$

Theorem B.5

(Thm 3 in [14]—Jacobian localization for many vortices) Let \(\Omega \) be a bounded, open, simply connected subset of \(\mathbb {R}^2\) with \(\mathcal {C}^1\) boundary. There exists constants C and K, depending on \(\mathrm{diam}(\Omega )\), with the following property: For any \(u \in H^1(\Omega ,\mathbb {C})\), if there exists \(n\ge 0\) distinct points \(a_1,\ldots ,a_n\) in \(\Omega \) and \(d \in \{\pm 1\}^{n}\) such that

$$\begin{aligned} \Vert Ju-\pi \sum _{i=1}^n d_i \delta _{a_i}\Vert _{\dot{W}^{-1,1}(\Omega )} \le \frac{\rho _a}{Kn^5}, \end{aligned}$$

where

$$\begin{aligned} \rho _a := \frac{1}{4}\mathrm{min}_i\left\{ \mathrm{min}_{j\ne i}|a_i-a_j|, \mathrm{dist}(a_i,\partial \Omega )\right\} , \end{aligned}$$

and if in addition \({\mathcal E}_{\varepsilon }(u,\Omega ) \ge 1\) and

$$\begin{aligned} \frac{n^5}{\rho _a}{\mathcal E}_{\varepsilon }(u,\Omega ) + \frac{n^{10}}{\rho _a^2}\sqrt{{\mathcal E}_{\varepsilon }(u,\Omega )} \le \frac{1}{\varepsilon }, \end{aligned}$$

then there exist \(\xi _1,\ldots ,\xi _d\) in \(\Omega \) such that

$$\begin{aligned}&\left\| Ju - \pi \sum _{i=1}^n d_i \delta _{\xi _i} \right\| _{\dot{W}^{-1,1}(\Omega )}\\&\quad \le C\varepsilon \left[ n(C+\Xi _\Omega ^\varepsilon )^2e^{\Xi _\Omega ^\varepsilon /\pi } + (C+\Xi _\Omega ^\varepsilon )\frac{n^5}{\rho _a} + {\mathcal E}_{\varepsilon }(u,\Omega )\right] , \end{aligned}$$

where

$$\begin{aligned} \Xi _\Omega ^\varepsilon:= & {} {\mathcal E}_{\varepsilon }(u,\Omega )-n(\pi \log \frac{1}{\varepsilon } + \gamma ) \\&+\,- \pi \Big ( \sum _{i\ne j} d_id_j \log |a_i-a_j| + \sum _{i,j} d_id_j H_\Omega (a_i,a_j)\Big )\end{aligned}$$

and \(H_\Omega \) is the Robin function of \(\Omega .\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jerrard, R.L., Smets, D. Leapfrogging Vortex Rings for the Three Dimensional Gross-Pitaevskii Equation. Ann. PDE 4, 4 (2018). https://doi.org/10.1007/s40818-017-0040-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s40818-017-0040-x

Keywords

Navigation