Skip to main content
Log in

Time Evolution of Concentrated Vortex Rings

  • Published:
Journal of Mathematical Fluid Mechanics Aims and scope Submit manuscript

Abstract

We study the time evolution of an incompressible fluid with axisymmetry without swirl when the vorticity is sharply concentrated. In particular, we consider N disjoint vortex rings of size \(\varepsilon \) and intensity of the order of \(|\log \varepsilon |^{ -1}\). We show that in the limit \(\varepsilon \rightarrow 0\), when the density of vorticity becomes very large, the movement of each vortex ring converges to a simple translation, at least for a small but positive time.

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.

Similar content being viewed by others

Notes

  1. This mollification is obtained by modifying the stream function associated to the field. The existence of such function for axisymmetric flow without swirl is a well known fact, see, e.g., [8, Section 2].

  2. \(\lfloor a\rfloor \) denotes the integer part of the positive number a.

  3. We remark that for axisymmetric flow the integration by parts in the computation of \(\dot{E}\) is allowed since \({\bar{u}}({{\varvec{\xi }}},t) \sim |{{\varvec{\xi }}}|^{-3}\), see (2.5, 2.6).

References

  1. Ambrosetti, A., Struwe, M.: Existence of steady rings in an ideal fluid. Arch. Ration. Mech. Anal. 108, 97–108 (1989)

    Article  MathSciNet  Google Scholar 

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

    Article  Google Scholar 

  3. Brunelli, E., Marchioro, C.: Vanishing viscosity limit for a smoke ring with concentrated vorticity. J. Math. Fluid Mech. 13, 421–428 (2011)

    Article  ADS  MathSciNet  Google Scholar 

  4. Buttà, P., Marchioro, C.: Long time evolution of concentrated Euler flows with planar symmetry. SIAM J. Math. Anal. 50, 735–760 (2018)

    Article  MathSciNet  Google Scholar 

  5. Colagrossi, A., Graziani, G., Pulvirenti, M.: Particles for fluids: SPH versus vortex methods. Math. Mech. Complex Syst. 2, 45–70 (2014)

    Article  MathSciNet  Google Scholar 

  6. Cetrone, D., Serafini, G.: Long time evolution of fuids with concentrated vorticity and convergence to the point-vortex model. Rendiconti di Matematica e delle sue applicazioni 39, 29–78 (2018)

    MATH  Google Scholar 

  7. Fraenkel, L.E.: On steady vortex rings of small cross-section in an ideal fluid. Proc. Roy. Soc. Lond. A 316, 29–62 (1970)

    Article  ADS  Google Scholar 

  8. Fraenkel, L.E., Berger, M.S.: A global theory of steady vortex rings in an ideal fluid. Acta Math. 132, 13–51 (1974)

    Article  MathSciNet  Google Scholar 

  9. Gallay, T.: Interaction of vortices in weakly viscous planar flows. Arch. Ration. Mech. Anal. 200, 445–490 (2011)

    Article  MathSciNet  Google Scholar 

  10. Marchioro, C.: On the vanishing viscosity limit for two-dimensional Navier–Stokes equations with singular initial data. Math. Meth. Appl. Sci. 12, 463–470 (1990)

    Article  MathSciNet  Google Scholar 

  11. Marchioro, C.: On the inviscid limit for a fluid with a concentrated vorticity. Commun. Math. Phys. 196, 53–65 (1998)

    Article  ADS  MathSciNet  Google Scholar 

  12. Marchioro, C.: Large smoke rings with concentrated vorticity. J. Math. Phys. 40, 869–883 (1999)

    Article  ADS  MathSciNet  Google Scholar 

  13. Marchioro, C.: Vanishing viscosity limit for an incompressible fluid with concentrated vorticity. J. Math. Phys. 48, 065302 (2007)

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

    Article  MathSciNet  Google Scholar 

  15. Marchioro, C., Pulvirenti, M.: On the vortex-wave system. In: Francaviglia, M. (ed.) Mechanics, Analisis and Geometry, 200 Years After Lagrange, pp. 79–95. Elsevier, Amsterdam (1991)

    Chapter  Google Scholar 

  16. Marchioro, C., Pulvirenti, M.: Mathematical Theory of Incompressible Non-viscous Fluids. Applied Mathematical Sciences, vol. 96. Springer, New York (1994)

    Book  Google Scholar 

  17. Shariff, K., Leonard, A.: Vortex rings. Annu. Rev. Fluid Mech. 24, 235–279 (1972)

    Article  ADS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Paolo Buttà.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Communicated by S. Shkoller

Publisher's Note

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

Appendices

Appendix A. Planar Symmetry

In this appendix we briefly recall some results concerning the time evolution of concentrated Euler flows with planar symmetry (without giving a complete list of references on this topic). The Euler equations for an incompressible inviscid fluid in the whole space with planar symmetry and constant density reads

$$\begin{aligned} \partial _t\omega (x,t) + (u \cdot \nabla ) \omega (x,t) = 0, \quad \nabla \cdot u(x,t) = 0,\quad x\in {{\mathbb {R}}}^2. \end{aligned}$$

By assuming that u vanishes at infinity, the velocity field is reconstructed from the vorticity as

$$\begin{aligned} u(x,t) = \int {\mathrm {d}}y\, K(x-y) \, \omega (y,t), \end{aligned}$$

with \(K(\cdot )\) as in Eq. (3.7).

We assume that initially the vorticity is concentrated in N blobs of the form

$$\begin{aligned} \omega _\varepsilon (x,0) = \sum _{i=1}^N \omega _{i,\varepsilon }(x,0), \end{aligned}$$

where \(\omega _{i,\varepsilon }(x,0)\), \(i=1,\ldots , N\), are functions with definite sign such that

$$\begin{aligned} \Lambda _{i,\varepsilon }(0) := \mathop {\mathrm{supp}}\nolimits \, \omega _{i,\varepsilon }(\cdot ,0) \subset \Sigma (z_i|\varepsilon ), \quad \Sigma (z_i|\varepsilon )\cap \Sigma (z_j|\varepsilon )=\emptyset \quad \forall \, i \ne j, \end{aligned}$$

with \(\varepsilon \in (0,1)\) a small parameter and the points \(z_i\in {{\mathbb {R}}}^2\).

In this case, the solution of the Euler equations is strictly related to the point vortex model, the dynamical system defined by the following system of ordinary differential equations,

$$\begin{aligned} {\dot{z}}_i(t) = \sum _{\begin{array}{c} j=1 \\ j\ne i \end{array}}^N A_j K(z_i(t)-z_j (t)), \end{aligned}$$
(A.1)

where \(A_j\) is called the “intensity” of the jth vortex. More precisely, it has been proven that in general, for small \(\varepsilon \), the time evolution of the vorticity has the same form,

$$\begin{aligned} \Lambda _{i,\varepsilon }(t) := \mathop {\mathrm{supp}}\nolimits \, \omega _{i,\varepsilon }(\cdot ,r_\varepsilon (t)) \subset \Sigma (z_i(t),r_\varepsilon (t)), \end{aligned}$$

with

$$\begin{aligned} \Sigma (z_i(t),r_\varepsilon (t)) \cap \Sigma (z_j(t),r_\varepsilon (t)) =0\quad \forall \, i \ne j\;, \end{aligned}$$

where \(r_\varepsilon (t)\) is a positive function and \(\{z_i(t); i=1,\ldots , N\}\) is the solution to Eq. (A.1) with initial conditions \(z_i(0)=z_i\) and intensity \(A_i = \int {\mathrm {d}}x \, \omega _{i,\varepsilon }(x,0)\). The point \(z_i(t)\) in the plane thus identifies a straight line in the space around which the vorticity is concentrated.

When all the intensities \(A_i\) have the same sign (positive or negative) Eq. (A.1) has a global in time solution. Instead, if the signs are different there are examples in which collapses could occur (e.g., two vortices collide or a vortex goes to infinity in finite time). However, these pathological events are exceptional. For a review on this issue see, for instance, [16].

The point vortex model has been introduced in the eighteenth century by Helmholtz as a “solution” of the Euler equations, and widely analyzed in many papers. One hundred years later, it has been considered as a numerical approximation of the Euler evolution for very irregular initial data. As a numerical tool, this system is considered when N is very large and the intensity of each vortex very small (of the order of \(N^{-1}\)). On this topic there are several papers, we only quote the recent review [5].

As explained at the beginning of this appendix, a different point of view is adopted for finite N and consists in considering the point vortex model as an approximation of N very concentrated vortices, say with support of diameter 2\(\varepsilon \rightarrow 0\). It is worthwhile to emphasize that it cannot be an approximation of each evolved path, because the length of the trajectory of a fluid element diverges as \(\varepsilon \rightarrow 0\). On the other hand, by virtue of rapid rotations, a system of N disjoint concentrated patches of vorticity converges as a measure to a linear combination of Dirac measures \(\sum _{i=1}^N A_i \delta _{z_i(t)}\) for any positive time. This convergence has been proven 25 years ago. Recently, the problem on how long the sharp localization of the vorticity remains valid has been analyzed in [4].

It is possible to introduce a small viscosity \(\nu \) and study the small viscosity limit \(\nu \rightarrow 0\). For finite times the status of art has been already discussed in Sect. 4.2. The validity of the convergence on very long times (diverging as \(\varepsilon \rightarrow 0\)) is analyzed in [6].

Appendix B. Proof of Some Technical Results

Proof of Lemma 3.1

The proof of [2, Lemma 2.1] is based on the conservation along the motion of the kinetic energy,

$$\begin{aligned} E = \frac{1}{2} \int {\mathrm {d}}{{\varvec{\xi }}}\, |{{\varvec{u}}} ({{\varvec{\xi }}},t)|^2 = \frac{1}{2} \int {\mathrm {d}}z \int _0^\infty {\mathrm {d}}r \, 2\pi r \big [u_z(z,r,t)^2+u_r(z,r,t)^2\big ]. \end{aligned}$$

More precisely, the assumptions on the initial vorticity, together with Eq. (1.9) and the conservation of the quantities

$$\begin{aligned} M_0 = \int {\mathrm {d}}z \int _0^\infty {\mathrm {d}}r \, \omega _\varepsilon (z,r,t), \quad M_2 = \int {\mathrm {d}}z \int _0^\infty {\mathrm {d}}r \, r^2 \omega _\varepsilon (z,r,t) , \end{aligned}$$

allow to compute the asymptotic behavior as \(\varepsilon \rightarrow 0\) of the energy E, from which the desired concentration estimate is deduced.

In this case, since the vector field \({{\varvec{F}}}^\varepsilon = (F^\varepsilon _z,F^\varepsilon _r,F^\varepsilon _\theta ) := (F^\varepsilon _1,F^\varepsilon _2,0)\) has zero divergence, the conservation laws of the energy E and of \(M_0\) are still valid. Indeed,Footnote 3

$$\begin{aligned} \begin{aligned} \dot{E}&= \int {\mathrm {d}}{{\varvec{\xi }}}\, {{\varvec{u}}} \cdot \partial _t {{\varvec{u}}} = - \int {\mathrm {d}}{{\varvec{\xi }}}\, {{\varvec{u}}} \cdot \big [({{\varvec{u}}} + {{\varvec{F}}}^\varepsilon )\cdot \nabla {{\varvec{u}}} + \nabla p\big ] \\&= \int {\mathrm {d}}{{\varvec{\xi }}}\, \bigg \{\frac{|{{\varvec{u}}}|^2}{2} \nabla \cdot ({{\varvec{u}}} + {{\varvec{F}}}^\varepsilon ) + p\, \nabla \cdot {{\varvec{u}}} \bigg \} = 0, \end{aligned} \end{aligned}$$

while, by Liouville’s theorem and Eq. (1.9),

$$\begin{aligned} M_0(t) = \frac{1}{2\pi } \int {\mathrm {d}}{{\varvec{\xi }}}\, \frac{\omega _\varepsilon ({{\varvec{\xi }}},t)}{r} = \int {\mathrm {d}}{{\varvec{\xi }}}_0 \, \frac{\omega _\varepsilon ({{\varvec{\xi }}}_0,0)}{r_0} = M_0(0) \end{aligned}$$
(B.1)

(above, the coordinate transformation is \({{\varvec{\xi }}}=\phi ^t({{\varvec{\xi }}}_0)\) with \(\phi ^t\) the flow generated by \(\dot{{{\varvec{\xi }}}} = {{\varvec{u}}}({{\varvec{\xi }}},t) + {{\varvec{F}}}^\varepsilon ({{\varvec{\xi }}},t)\)).

Concerning the variation of \(M_2\), since \(\omega _\varepsilon (z,r,t)\) has compact support, we can apply Eq. (2.10) with \(f(x,t) = x_2^2\), so that

$$\begin{aligned} \dot{M}_2 = \int {\mathrm {d}}x \, \omega _\varepsilon (x,t) \,2 x_2 F^\varepsilon _2(x,t) , \end{aligned}$$

which implies \(|\dot{M}_2| \le 2C_F |\log \varepsilon |^{-3/2} \sqrt{M_2}\) in view of Assumption 2.1, item (b). Therefore, \(M_2 \le 2(|{\bar{\zeta }}_2| + \varepsilon )^2 M_0 + 4C_F^2T^2 |\log \varepsilon |^{-3} \le {({\mathrm{const.}})\,}|\log \varepsilon |^{-1}\). This is the same estimate, but for a larger constant, which is obtained in absence of \(F^\varepsilon \). Since the particular value of this constant is easily seen to be irrelevant in the proof of [2, Thm. 1], the lemma follows. \(\square \)

Proof of Lemma 3.3

Letting

$$\begin{aligned} a(x,y) := \frac{|x-y|}{\sqrt{x_2y_2}}, \end{aligned}$$
(B.2)

we have,

$$\begin{aligned} H(x,y) = - \frac{I_1(a(x,y))}{2\pi } \frac{(x-y)^\perp }{x_2\sqrt{x_2y_2}} + \frac{I_2(a(x,y))}{2\pi }\frac{1}{x_2}\sqrt{\frac{y_2}{x_2}} \begin{pmatrix} 1 \\ 0 \end{pmatrix}, \end{aligned}$$

where

$$\begin{aligned} I_1(a) = \int _0^\pi {\mathrm {d}}\theta \, \frac{\cos \theta }{[a^2+2(1-\cos \theta )]^{3/2}} , \quad I_2(a) = \int _0^\pi {\mathrm {d}}\theta \, \frac{1-\cos \theta }{[a^2+2(1-\cos \theta )]^{3/2}}. \end{aligned}$$

By an explicit computation, see, e.g., the Appendix in [12], for any \(a>0\),

$$\begin{aligned} I_1(a) = \frac{1}{a^2} + \frac{1}{4} \log \frac{a}{1+a} + \frac{c_1(a)}{1+a}, \quad I_2(a) = -\frac{1}{2} \log \frac{a}{1+a} + \frac{c_2(a)}{1+a}, \end{aligned}$$

with \(c_1(a)\), \(c_1'(a)\), \(c_2(a)\), \(c_2'(a)\) uniformly bounded for \(a\in (0,+\infty )\). Therefore, the kernel R(xy) defined by (3.8) is given by

$$\begin{aligned} {{\mathcal {R}}} (x,y) = \sum _{j=1}^6 R^j(x,y), \end{aligned}$$

with, for \(a = a(x,y)\) as in (B.2),

$$\begin{aligned} R^1(x,y)&= \frac{1}{2\pi } \bigg (1 - \sqrt{\frac{y_2}{x_2}}\bigg ) \frac{(x-y)^\perp }{|x-y|^2},\;\; R^2(x,y) = \frac{1}{8\pi } \bigg (\log \frac{1+a}{a}\bigg ) \frac{(x-y)^\perp }{x_2\sqrt{x_2y_2}}, \\ R^3(x,y)&= \frac{1}{4\pi x_2} \sqrt{\frac{y_2}{x_2}} \bigg (\log \frac{|x-y|}{1+|x-y|} - \log \frac{a}{1+a}\bigg ) \begin{pmatrix} 1 \\ 0 \end{pmatrix} , \\ R^4(x,y)&= \frac{1}{4\pi x_2} \bigg (1-\sqrt{\frac{y_2}{x_2}}\bigg ) \log \frac{|x-y|}{1+|x-y|} \begin{pmatrix} 1 \\ 0 \end{pmatrix} , \\ R^5(x,y)&= -\frac{c_1(a)}{2\pi (1+a)} \frac{(x-y)^\perp }{x_2\sqrt{x_2y_2}}, \quad R^6(x,y) = \frac{c_2(a)}{2\pi (1+a)x_2}\sqrt{\frac{y_2}{x_2}} \begin{pmatrix} 1 \\ 0 \end{pmatrix}. \end{aligned}$$

Using that

$$\begin{aligned} \bigg |1-\sqrt{\frac{y_2}{x_2}} \bigg | = \frac{|y_2-x_2|}{x_2+\sqrt{x_2y_2}} \le \frac{|x-y|}{x_2} \end{aligned}$$

and

$$\begin{aligned} \bigg |\log \frac{|x-y|}{1+|x-y|} - \log \frac{a}{1+a}\bigg | = \bigg | \log \frac{1+a}{(x_2y_2)^{-1/2}+a}\bigg | \le \frac{1}{2} |\log (x_2y_2)|, \end{aligned}$$

we have,

$$\begin{aligned}&|R^1(x,y)| = \frac{1}{2\pi } \bigg |1-\sqrt{\frac{y_2}{x_2}} \bigg | \frac{1}{|x-y|}\le \frac{1}{2\pi x_2}, \\&|R^2(x,y)| = \frac{1}{8\pi x_2} \bigg (\log \frac{1+a}{a}\bigg ) \frac{|x-y|}{\sqrt{x_2y_2}} \le \frac{1}{8\pi x_2} \, \sup _{\rho>0}\bigg ( \rho \log \frac{1+\rho }{\rho }\bigg ), \\&|R^3(x,y)| + |R^6(x,y)| \le \frac{1}{4\pi x_2} \sqrt{\frac{y_2}{x_2}} \bigg (|\log (x_2y_2)| + \sup _{\rho>0} \frac{2c_2(\rho )}{1+\rho }\bigg ), \\&|R^4(x,y)| = \frac{1}{4\pi x_2} \bigg |1-\sqrt{\frac{y_2}{x_2}} \bigg | \log \frac{1+|x-y|}{|x-y|} \le \frac{1}{4\pi x_2^2} \,\sup _{\rho>0}\bigg ( \rho \log \frac{1+\rho }{\rho }\bigg ), \\&|R^5(x,y)| = \frac{|c_1(a)|}{2\pi (1+a)} \frac{|x-y|}{x_2\sqrt{x_2y_2}} \le \frac{1}{2\pi x_2} \, \sup _{\rho >0}\frac{\rho c_1(\rho )}{1+\rho }. \end{aligned}$$

In conclusion, there is \(C_0>0\) such that

$$\begin{aligned} \begin{aligned}&|R^1(x,y)| + |R^2(x,y)| + |R^5(x,y)| \le \frac{C_0}{x_2}, \quad |R^4(x,y)| \le \frac{C_0}{x_2^2}, \\&|R^3(x,y)| + |R^6(x,y)| \le \frac{C_0}{x_2} \sqrt{\frac{y_2}{x_2}} \bigg (1+|\log (x_2y_2)|\bigg ). \end{aligned} \end{aligned}$$

The lemma is thus proven. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Buttà, P., Marchioro, C. Time Evolution of Concentrated Vortex Rings. J. Math. Fluid Mech. 22, 19 (2020). https://doi.org/10.1007/s00021-020-0482-x

Download citation

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00021-020-0482-x

Keywords

Mathematics Subject Classification

Navigation