Skip to main content
Log in

Perfect fluid flows on \(\mathbb {R}^d\) with growth/decay conditions at infinity

  • Published:
Mathematische Annalen Aims and scope Submit manuscript

Abstract

We study the well-posedness and the spatial behavior at infinity of perfect fluid flows on \(\mathbb {R}^d\) with initial velocity in a scale of weighted Sobolev spaces that allow spatial growth/decay at infinity as \(|x|^\beta \) with \(\beta <1/2\). Moreover, for initial velocity with sufficient spatial decay, we show that the solution of the Euler equation generically develops an asymptotic expansion at infinity with non-vanishing asymptotic terms that depend analytically on time and the initial data. For initial data in the Schwartz space, we identify the evolution space of the fluid velocity with a certain space of symbols.

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

Availability of data and material

Not applicable.

Code availability

Not applicable.

Notes

  1. The uniqueness statements in this paper are considered within the described class of solutions.

  2. In particular, by (3b), the solution u has an asymptotic expansion at infinity of order equal to the integer part of \(\delta +d/p\) with remainder in \(W^{m,p}_\delta \).

  3. Since the asymptotic sum in (6) vanishes for \(\delta +d/p<d+1\), formula (6) continues to hold also in the case of Theorems 1.1 and 1.2(a).

  4. Recall that for \(m>d/p\) the elements of \(W^{m,p}_\delta \) are of order \(O\big (1/{|x|}^{\delta +d/p}\big )\) as \(|x|\rightarrow \infty \).

  5. By “smooth” we mean \(C^\infty \)-smooth. However, these maps will actually be shown to be real analytic.

  6. Here \(\Delta ^{-1} : W^{m-1,p}_{\kappa +1}\rightarrow W^{m+1,p}_{\kappa -1}/\mathcal {N}_0\) with \(0<\kappa +d/p<1\) (cf. (21)).

  7. This is a re-expression of the conservation of vorticity theorem.

  8. Note that \({{\mathcal {A}}}D^{m,p}_{N;0}\subseteq {{\mathcal {A}}}D^{m,p}_0\) and the inclusion is bounded.

  9. Since \(\delta +d/p\notin \mathbb {Z}\) we have that \(k<\delta +d/p\).

  10. No summation over j in the formula is assumed.

  11. As noted in [26], the completeness condition in [26, Theorem 2] can be replaces by local completeness.

  12. For the case when \(d=2\) we refer to Proposition 3.3 in [30].

  13. Note that \(g H_{k';l}\in L^1(\mathbb {R}^d,\mathbb {R})\).

  14. In particular, we obtain an alternative proof of Proposition C.1 in the case \(d=2\).

  15. Theorem 3.2 in [30] states that this map is an isomorphism; see (16) in [30] for the definition of the space \(\widetilde{\mathcal {Z}}^{m-1,p}_{M+1}\).

References

  1. Arnold, V.: Sur la geometrié differentielle des groupes de Lie de dimension infinie et ses applications à l’hydrodynamique des fluids parfaits. Ann. Inst. Fourier 16(1), 319–361 (1966)

    Article  MathSciNet  Google Scholar 

  2. Bartnik, R.: The mass of an asymptotically flat manifold. Commun. Pure Appl. Math. 39, 661–693 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  3. Benedetto, D., Marchioro, C., Pulvirenti, M.: On the Euler flow in \({\mathbb{R}}^2\). Arch. Ration. Mech. Anal. 123, 377–386 (1993)

    MATH  Google Scholar 

  4. Bondareva, I., Shubin, M.: Equations of Korteweg–de Vries type in classes of increasing functions. Trudy Sem. Petrovsk. 256(14), 45–56 (1989)

    MATH  Google Scholar 

  5. Brandolese, L., Meyer, Y.: On the instantaneous spreading for the Navier–Stokes system in the whole space. A tribute to J. L. Lions. ESAIM Control Optim. Calc. Var. 8, 273–285 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  6. Brandolese, L.: Far field geometric structures of 2D flows with localized vorticity. Math. Ann. https://doi.org/10.1007/s00208-021-02177-8 (to appear)

  7. Cantor, M.: Perfect fluid flows over \({\mathbb{R}}^n\) with asymptotic conditions. J. Funct. Anal. 18, 73–84 (1975)

    MathSciNet  MATH  Google Scholar 

  8. Cantor, M.: Some problems of global analysis on asymptotically simple manifolds. Compos. Math. 38(1), 3–35 (1979)

    MathSciNet  MATH  Google Scholar 

  9. Chemin, J.-Y.: Fluides Parfaits Incompressibles. Astérisque. SMF 230 (1995)

  10. Cozzi, E., Kelliher, J.: Well-posedness of the 2D Euler equations when the velocity grows at infinity. Discret. Contin. Dyn. Syst. 39(5), 2361–2392 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  11. Dobrokhotov, S., Shafarevich, A.: Some integral identities and remarks on the decay at infinity of the solutions of the Navier–Stokes equations in the entire space. Russ. J. Math. Phys. 2(1), 133–135 (1994)

    MathSciNet  MATH  Google Scholar 

  12. Ebin, D., Marsden, J.: Groups of diffeomorphisms and the motion of an incompressible fluid. Ann. Math. 92, 102–163 (1970)

    Article  MathSciNet  MATH  Google Scholar 

  13. Elgindi, T., Jeong, I.-J.: Symmetries and critical phenomena in fluids. Commun. Pure Appl. Math. 73, 257–316 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  14. Inci, H., Kappeler, T., Topalov, P.: On the regularity of the composition of diffeomorphisms. Mem. Am. Math. Soc. 226(1062), 1–60 (2013) (this is a small booklet)

  15. Kappeler, T., Perry, P., Shubin, M., Topalov, P.: Solutions of mKdV in classes of functions unbounded at infinity. J. Geom. Anal. 18, 443–477 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  16. Kato, T.: Nonstationary flows of viscous and ideal fluids in \({\mathbb{R}}^3\). J. Funct. Anal. 9, 296–305 (1972)

    MATH  Google Scholar 

  17. Kato, T., Ponce, G.: Well-posedness of the Euler and Navier–Stokes equations in the Lebesgue spaces \(L^p_s({\mathbb{R}}^2)\). Rev. Mat. Iberoam. 2, 73–88 (1986)

    Google Scholar 

  18. Kato, T., Ponce, G.: On nonstationary flows of viscous and ideal fluids in \(L^p_s({\mathbb{R}}^2)\). Duke Math. J. 55, 487–499 (1987)

    MathSciNet  MATH  Google Scholar 

  19. Kelliher, J.: A characterization at infinity of bounded vorticity, bounded velocity solutions to the 2D Euler equations. Indiana Math J. 64(6), 1643–1666 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  20. Kukavica, I., Reis, E.: Asymptotic expansion for solutions of the Navier–Stokes equations with potential forces. J. Differ. Equ. 250(1), 607–622 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  21. Majda, A., Bertozzi, A.: Vorticity and Incompressible Flow. Cambridge Texts in Applied Mathematics. Cambridge University Press, Cambridge (2002)

    MATH  Google Scholar 

  22. McOwen, R.: The behavior of the Laplacian on weighted Sobolev spaces. Commun. Pure Appl. Math. 32(6), 783–795 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  23. McOwen, R., Topalov, P.: Groups of asymptotic diffeomorphisms. Discret. Contin. Dyn. Syst. 36(11), 6331–6377 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  24. McOwen, R., Topalov, P.: Spatial asymptotic expansions in the incompressible Euler equation. GAFA 27, 637–675 (2017)

    MathSciNet  MATH  Google Scholar 

  25. Misiolek, G., Yoneda, T.: Continuity of the solution map of the Euler equation in Hölder spaces and weak norm inflation in Besov spaces. Trans. Am. Math. Soc. 370(7), 4709–4730 (2018)

    Article  MATH  Google Scholar 

  26. Montgomery, D.: On continuity in topological groups. Bull. Am. Math. Soc. 42, 879–882 (1936)

    Article  MathSciNet  MATH  Google Scholar 

  27. Serfati, P.: Équation d’Euler et holomorphies à faible régularité spatiale. C. R. Acad. Sci. Paris Sér. I Math. 320(2), 175–180 (1995)

  28. Serfati, P.: Solutions \(C^\infty \) en temps, \(n\)-log Lipschitz bornées en espace et équation d’Euler. C. R. Acad. Sei. Paris Sér. I Math. 320(5), 555–558 (1995)

  29. Shnirelman, A.: On the analyticity of particle trajectories in the ideal incompressible fluid. arXiv:1205.5837v

  30. Sultan, S., Topalov, P.: On the asymptotic behavior of solutions of the 2d Euler equation. J. Differ. Equ. 269(6), 5280–5337 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  31. Shubin, M.: Pseudodifferential Operators and Spectral Theory, 2nd edn. Springer, Berlin (2001)

    Book  MATH  Google Scholar 

  32. Sun, X., Topalov, P.: On the group of almost periodic diffeomorphisms and its exponential map. IMRN 2021(13), 9648–9716 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  33. Wolibner, W.: Un theorème sur l’existence du mouvement plan d’un fluide parfait, homogène, incompressible, pendant un temps infiniment long. Math. Z. 37(1), 698–726 (1933)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The second author is thankful to the IMI of the BAS for the support.

Funding

PT is partially supported by the Simons Foundation, Award #526907.

Author information

Authors and Affiliations

Authors

Contributions

Not applicable.

Corresponding author

Correspondence to P. Topalov.

Ethics declarations

Conflict of interest/Competing interests

No conflict of interests.

Additional information

Communicated by Y. Giga.

Publisher's Note

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

Appendices

Auxiliary results

Most of the results in this Appendix are only generalizations to \(\gamma +d/p>-1 \) of results in [23] that assumed \(\gamma +d/p>0 \). Rather than repeat the detailed proofs given in [23], we will simply describe how to generalization them to the case \(\gamma +d/p>-1 \). In one instance, we generalize a statement from [14]. The following lemma follows directly from (3a).

Lemma A.1

If \(w\in W^{m,p}_\gamma \) with \(m>d/p\) and \(\gamma +d/p>-1,\) then

$$\begin{aligned} C_1\langle x\rangle \le \big \langle x+w(x)\big \rangle \le C_2 \langle x\rangle \quad \hbox {for} \ x\in \mathbb {R}^d, \end{aligned}$$
(82)

where \(C_1,C_2>0\) may be chosen locally uniformly for \(w\in W^{m,p}_\gamma \).

Local uniformity means that for any \(w_0\in W^{m,p}_\gamma \) there exists an open neighborhood U of \(w_0\) in \(W^{m,p}_\gamma \) such that the inequality (82) holds for any \(w\in U\).

Our next result is analogous to Lemma 6.3 in [23]. Let \(|\mathrm{d}\varphi (x)|\) denote the sum of the absolute values of the elements of the matrix \(\mathrm{d}\varphi (x)\) for \(x\in \mathbb {R}^d\).

Lemma A.2

If \(\varphi =\mathrm{id}+w\in {{\mathcal {D}}}^{m,p}_\gamma \) with \(m>1+d/p\) and \(\gamma +d/p>-1,\) then

$$\begin{aligned} \big |\mathrm{d}\varphi (x)\big |\le C \quad \hbox {and} \quad 0<\varepsilon \le \det (\mathrm{d}\varphi (x)) \quad \hbox {for} \ x\in \mathbb {R}^d, \end{aligned}$$
(83)

where C may be chosen uniformly for \( \Vert w\Vert _{W^{m,p}_\gamma }\le M\) and \(\varepsilon \) may be chosen locally uniformly for \(w\in W^{m,p}_\gamma \).

Proof of Lemma A.2

It follows from (3a) that for \(m>1+d/p\) and \(\delta +d/p\ge -1\),

$$\begin{aligned} W^{m-1,p}_{\delta +1}\subseteq L^\infty \end{aligned}$$
(84)

is bounded. Take \(\varphi =\mathrm{id}+w\in D^{m,p}_\gamma \) with \(m>1+d/p\) and \(\gamma +d/p>-1\). Then, the first inequality in (83) follows from the boundedness of the inclusions

$$\begin{aligned} W^{m,p}_\gamma {\mathop {\hookrightarrow }\limits ^{\mathrm{d}}}W^{m-q,p}_{\gamma +1}\subseteq L^\infty \end{aligned}$$

and the fact that \(\mathrm{d}\varphi =I+\mathrm{d}w\). The second inequality in (83) follows in a similar way from (84), \(\mathrm{d}\varphi =I+\mathrm{d}w\), and (3a), since the latter implies that \(|\mathrm{d}w|=C\,\Vert \mathrm{d}w\Vert _{W^{m-1,p}_{\delta +1}}/\langle x\rangle ^{(\delta +p/d)+1}\) with \((\delta +p/d)+1>0\). The estimates above a locally uniform for \(\varphi \in D^{m,p}_\gamma \). \(\square \)

The following corollary follows from Lemma A.2 and Hadamard–Levy’s theorem.

Corollary A.1

If \(\varphi _0=\mathrm{id}+w_0\in {{\mathcal {D}}}^{m,p}_\gamma \) where \(m>1+d/p,\) \(\gamma +d/p>-1,\) and \({\widetilde{w}}\in W^{m,p}_\gamma \) with \(\Vert {\widetilde{w}}\Vert _{W^{m,p}_\gamma }\) sufficiently small,  then \(\varphi _0+{\widetilde{w}}\in {{\mathcal {D}}}^{m,p}_\gamma \).

In particular, the set of maps \({{\mathcal {D}}}^{m,p}_\gamma \) can be identified with an open set in \(W^{m,p}_\gamma \). Hence, \({{\mathcal {D}}}^{m,p}_\gamma \) is a Banach manifold modeled on \(W^{m,p}_\gamma \).

For \(\varphi \in {{\mathcal {D}}}^{m,p}_\gamma \), we know that \(\varphi ^{-1}\) exists but we need estimates at infinity in order to conclude that \(\varphi ^{-1}\in {{\mathcal {D}}}^{m,p}_\gamma \). The following is a first step and is analogous to Lemma 6.4 in [23].

Lemma A.3

If \(\varphi =\mathrm{id}+w\in {{\mathcal {D}}}^{m,p}_\gamma \) where \(m>1+d/p\) and \(\gamma +d/p>-1,\) then

$$\begin{aligned} \big |\mathrm{d}(\varphi ^{-1})(x)\big |\le C \quad \hbox {and} \quad 0<\varepsilon \le \det \,\big (\mathrm{d}(\varphi ^{-1})(x)\big ) \quad \hbox {for} \ x\in \mathbb {R}^d, \end{aligned}$$

where C and \(\varepsilon \) may be chosen locally uniformly for \( w\in W^{m,p}_\gamma \).

Proof of Lemma A.3

The lemma follows from Lemma A.2 and the formula

$$\begin{aligned} \mathrm{d}\big (\varphi ^{-1}\big )=\big [(\mathrm{d}\varphi )\circ \varphi ^{-1}\big ]^{-1} \end{aligned}$$

for the Jacobian matrix of \(\varphi ^{-1}\). \(\square \)

Now let us consider compositions of maps as in Theorem 2.1. We begin with an estimate.

Lemma A.4

Suppose \(m>1+d/p,\) \(\gamma +d/p>-1\) and \(\varphi \in {{\mathcal {D}}}^{m,p}_\gamma \). Then for every \(0\le k\le m\) and \(\delta \in \mathbb {R}\), we have

$$\begin{aligned} \Vert w\circ \varphi \Vert _{W^{k,p}_\delta } \le C\,\Vert w\Vert _{W^{k,p}_\delta }\quad \hbox {for all}\ f\in W^{k,p}_\delta , \end{aligned}$$

where C may be taken locally uniformly for \(\varphi \in {{\mathcal {D}}}^{m,p}_\gamma \).

Proof of Lemma A.4

This is proved by induction using Lemma A.3 for a change of integration variable when \(k=0\) and Proposition 2.2 in [23] to handle products in the induction step. For details see the proof of Lemma 6.5 in [23]. \(\square \)

Lemma A.5

Assume \(m>1+d/p,\) \(\gamma +d/p>-1,\) \(\delta \in \mathbb {R},\) and \(f\in C_c^\infty (\mathbb {R}^d)\). If \(\varphi , \varphi _k\in {{\mathcal {D}}}^{m,p}_\gamma \) with \(\varphi _k\rightarrow \varphi \) in \({{\mathcal {D}}}^{m,p}_\gamma \) as \(k\rightarrow \infty ,\) then \(f\circ \varphi _k\rightarrow f\circ \varphi \) in \(W^{m,p}_\delta (\mathbb {R}^d)\).

Proof of Lemma A.5

The only difference from the proof of Lemma 6.6 in [23] is that \(w_k(x)\) and w(x) in \(\varphi _k(x)=x+w_k(x)\) and \(\varphi (x)=x+w(x)\) are of order \(O(|x|^{1-\varepsilon })\) for some \(\varepsilon >0\) instead of just being bounded. But the rest of the proof in [23] can be used here without change. \(\square \)

Lemmas A.4 and A.5 can be combined to obtain the continuity of the composition.

Corollary A.2

Suppose \(m>1+d/p,\) \(\gamma +d/p>-1\) and \(\delta \in \mathbb {R}\). Then the composition \((f,\varphi )\mapsto f\circ \varphi ,\) \(W^{m,p}_\delta \times {{\mathcal {D}}}^{m,p}_\gamma \rightarrow W^{m,p}_\delta ,\) is continuous.

Proof of Corollary A.2

The details are the same as in the proof of Corollary 6.1 in [23]. \(\square \)

Next we investigate when the composition is \(C^1\).

Lemma A.6

Assume \(m>1+d/p,\) \(\gamma +d/p>-1,\) \(\delta \in \mathbb {R},\) and take \(\varphi _0\in {{\mathcal {D}}}^{m,p}_\gamma \). For \(f\in W^{m+1,p}_\delta \) and \(\varphi \in {{\mathcal {D}}}^{m,p}_\gamma \) sufficiently close to \(\varphi _0,\) we have

$$\begin{aligned} \Vert f\circ \varphi -f\circ \varphi _0\Vert _{W^{m,p}_{\delta +1}}\le C\,\Vert f\Vert _{W^{m+1,p}_\delta }\Vert \varphi -\varphi _0\Vert _{W^{m,p}_\gamma }, \end{aligned}$$

where \(C>0\) can be taken uniformly for all \(\varphi \) in an open neighborhood of \(\varphi _0\).

Proof of Lemma A.6

The details are the same as in the proof of Lemma 6.7 in [23]. \(\square \)

Corollary A.3

Suppose \(m>1+d/p,\) \(\gamma +d/p>-1\) and \(\delta \in \mathbb {R}\). Then the composition \((f,\varphi )\mapsto f\circ \varphi ,\) \(W^{m+1,p}_\delta \times {{\mathcal {D}}}^{m,p}_\gamma \rightarrow W^{m,p}_\delta ,\) is \(C^1\).

Proof of Corollary A.3

Using the above lemmas, Corollary A.3 can be proved following the proof of Proposition 5.1 in [23] (cf. also Appendix B in [32]). \(\square \)

This completes the proof of Theorem 2.1(a). Let us now prove Theorem 2.1(b).

Lemma A.7

If \(\varphi \in {{\mathcal {D}}}^{m,p}_\gamma \) where \(m>1+d/p\) and \(\gamma +d/p>-1,\) then \(\varphi ^{-1}\in {{\mathcal {D}}}^{m,p}_\gamma \).

Proof of Lemma A.7

Let \(\varphi =\mathrm{id}+w\) and \(\varphi ^{-1}=\mathrm{id}+u.\) To show \(\varphi ^{-1}\in {{\mathcal {D}}}^{m,p}_\gamma \), we need to show \(\partial ^\alpha u\in L^p_{\gamma +|\alpha |}\) for all \(|\alpha |\le m\). For \(\alpha =0 \) we use the change of variables \(x=\varphi (y)\) to compute

$$\begin{aligned} \begin{aligned} \int _{\mathbb {R}^d}\langle x\rangle ^{\gamma p}|u(x)|^p\,dx&=\int _{\mathbb {R}^d} \langle x\rangle ^{\gamma p}|\varphi ^{-1}(x)-x|^p\,dx =\int _{\mathbb {R}^d} \langle \varphi (y)\rangle ^{\gamma p}|w(y)|^p\,\det (\mathrm{d}\varphi (y))\,dy \\&\le C\,\int \langle y\rangle ^{\gamma p}|w(y)|^p\,dy <\infty , \end{aligned} \end{aligned}$$

where we have used (82), (83), and \(w=u\circ \varphi \in W^{m,p}_\gamma \) (cf. Lemma A.4). For \(1\le |\alpha |\le m\), we can proceed as in (28) in [14] to show

$$\begin{aligned} \partial ^\alpha (\varphi ^{-1}-\mathrm{id})=F^{(\alpha )}\circ \varphi ^{-1}, \end{aligned}$$
(85)

where \(F^{(\alpha )} : \mathbb {R}^d\rightarrow \mathbb {R}^d\) is in \(W^{m-|\alpha |,p}_{\gamma +|\alpha |}\). Then, by (85), for any \(\alpha \) with \(0\le |\alpha |\le m\),

$$\begin{aligned} \begin{aligned} \int \langle x\rangle ^{(\gamma +|\alpha |)p}\,|\partial ^\alpha (\varphi ^{-1}(x)-x)|^p\,dx&= \int \langle \varphi (y)\rangle ^{(\gamma +|\alpha |)p}\, |F^{(\alpha )}(y)|^p\,\det (\mathrm{d}\varphi (y))\,dy \\&\le C\,\int \langle y\rangle ^{(\gamma +|\alpha |)p}\,|F^{(\alpha )}(y)|^p\,dy <\infty . \end{aligned} \end{aligned}$$

Hence \(\varphi ^{-1}-\mathrm{id}\in W^{m,p}_\gamma \). This implies that \(\varphi \in {{\mathcal {D}}}^{m,p}_\gamma \). \(\square \)

The continuity of the map \(\varphi \mapsto \varphi ^{-1}\), \({{\mathcal {D}}}^{m,p}_\delta \rightarrow {{\mathcal {D}}}^{m,p}_\delta \), now follows from Corollary A.2 and Theorem 2 in [26].Footnote 11 The last statement in Theorem 2.1 (b) can be proved by following the proof of Proposition 2.13 in [14].

Remark A.1

Note that the regularity assumption \(m>1+d/p\) in Lemma A.7 above is weaker than the regularity assumption \(m>3+d/p\) in the analogous [23, Lemma 7.2] and [30, Lemma 2.5]. This happens since in [23, 30] one deals with the asymptotic expansion of \(\varphi ^{-1}\), that complicates the situation. Note however, that the results in [23, 30] can be extended to the case when \(m>1+d/p\) by expanding their proofs.

Inverting the Laplace operator

In this Appendix we present, in an extended form, a basic result about the inversion of the Laplace operator in weighted Sobolev spaces (see Lemma A.3 in [24]). Denote by \(\mathcal {S}'\) the space of tempered distributions in \(\mathbb {R}^d\).

Proposition B.1

Assume that \(d\ge 3\) and \(m\ge 0\) with \(1<p<\infty \).Footnote 12 Then,  for any \(g\in W^{m.p}_{\delta +2}(\mathbb {R}^d,\mathbb {R})\) with weight \(\delta \in \mathbb {R}\) such that \(\delta +d/p>0,\) \(\delta +d/p\notin \mathbb {Z},\) there exists a unique (up to adding a constant term) solution u in \(\mathcal {S}'\cap L^\infty \) of the Poisson equation

$$\begin{aligned} \Delta u=g \end{aligned}$$

such that

$$\begin{aligned} u:=\Delta ^{-1}g=\chi (r)\sum _{d-2\le k<\delta +d/p}\frac{a_k(\theta )}{r^k}+f,\quad f\in W^{m+2,p}_{\delta }, \end{aligned}$$
(86)

where \(a_k(\theta )\) is an eigenfunction of the Laplace operator \(-\Delta _S\) on the unit sphere \(S^{d-1}\) in \(\mathbb {R}^d\) with eigenvalue \(\lambda _{k-d+2}=k(k-d+2)\). If we fix for any \(k':=k-d+2\ge 0\) with \(d-2\le k<\delta +d/p\) an orthonormal basis

$$\begin{aligned} \big \{h_{k';l}(\theta )\,\big |\, 1\le l\le \nu (k')\big \},\quad \nu (k'):=\dim \mathcal {H}_{k'}, \end{aligned}$$

of the eigenspace \(\mathcal {H}_{k'}\) of \(-\Delta _S\) with eigenvalue \(\lambda _{k'}=k'(k'+d-2)\) and expand

$$\begin{aligned} a_k(\theta )=\sum _{l=1}^{\nu (k')}\widehat{a}_{k';l}\,h_{k';l}(\theta ) \end{aligned}$$

in the Fourier modes, then the Fourier coefficient \(\widehat{a}_{k';l}\) equals

$$\begin{aligned} \widehat{a}_{k';l}=-\frac{1}{2k'+d-2}\int _{\mathbb {R}^d}g(x) H_{k';l}(x)\,\mathrm{d}x,\quad 1\le l\le \nu (k'), \end{aligned}$$
(87)

where \(H_{k';l}(x):=h_{k';l}(\theta ) r^{k'}\) is the homogeneous harmonic polynomial (of degree \(k'\ge 0\)) that corresponds to the Fourier mode \(h_{k';l}(\theta )\).Footnote 13 If we coordinatize the linear space \(\mathcal {I}^{m,p}_\delta \) of functions of the form (86) by the Fourier coefficients of \(a_k(\theta ),\) \(d-2\le k<\delta +d/p,\) and the reminder \(f\in W^{m,p}_\delta \) then the map

$$\begin{aligned} \Delta ^{-1} : W^{m+2,p}_{\delta +2}\rightarrow \mathcal {I}^{m,p}_\delta \end{aligned}$$

is an isomorphism of Banach spaces.

Remark B.1

Note that the eigenvalues of the Laplace operator on the unit sphere \(S^{d-1}\) in \(\mathbb {R}^d\) are highly degenerate: we have that \(\dim \mathcal {H}_0=1\) (these are all constants), \(\dim \mathcal {H}_1=d\) (all linear polynomials restricted to \(S^{d-1}\)), and

$$\begin{aligned} \dim \mathcal {H}_{k'}=\left( {\begin{array}{c}d-1+k'\\ d-1\end{array}}\right) -\left( {\begin{array}{c}d-3+k'\\ d-1\end{array}}\right) ,\quad k'\ge 2, \end{aligned}$$

where the second binomial coefficient above vanishes when \(d=2\).

Proof of Proposition B.1

The fact that there exists a unique solution \(u\in \mathcal {S}'\cap L^\infty \) of the form (86) follows from [24, Lemma A.3(b)] (cf. also [22]) and the fact that there is a bijective correspondence between the eigenfunctions of the Laplace operator \(-\Delta _S\) on the unit sphere \(S^{d-1}\) in \(\mathbb {R}^d\) with eigenvalue \(\lambda _{k'}=k'(k'+d-2)\) and the restriction to \(S^{d-1}\) of harmonic polynomials of degree \(k'\ge 0\) ( [31, §22.2]. Let us now prove the integral relation (87). To this end, take \(d-2\le k<\delta +d/p\), \(d-2\le n<\delta +d/p\), \(1\le l_1\le \nu (k')\), \(1\le l_2\le d(n')\) with \(n':=n-d+2\), and consider the \(k'\)-th asymptotic term \(A_{k';l}:=\chi h_{k';l}/r^k\) in (86). Then, it follows from the second Green’s identity and the fact that the eigenspaces \(\mathcal {H}_{k'}\) and \(\mathcal {H}_{n'}\) are \(L^2\)-orthogonal on \(S^{d-2}\) for \(k'\ne n'\), that

$$\begin{aligned}&\int _{\mathbb {R}^d}\Delta \big (A_{k';l_1}\big ) H_{n';l_2}\,\mathrm{d}x=\lim _{R\rightarrow \infty }\int _{S^{d-1}_R} \Big (\frac{\partial A_{k'l_1}}{\partial r}H_{n';l_2}-A_{k';l_1}\frac{\partial H_{n';l_2}}{\partial r}\Big ) \,d\sigma _R\nonumber \\&\quad =C(k,n)\lim _{R\rightarrow \infty }R^{n'-k'}\int _{S^{d-1}}h_{k';l_1} h_{n';l_2}\,d\sigma _1 =C(k,n)\,\delta _{kn} \end{aligned}$$
(88)

where \(C(k,n):=-(k'+n'+d-2)\ne 0\), \(S^{d-1}_R\) is the sphere of radius R, and \(\delta _{kn}\) is the Kronecker delta. Now, consider the remainder \(f\in W^{m+2,p}_\delta \) of the solution u in (86). For any \(d-2\le n<\delta +d/p\) and \(1\le l\le d(n')\) we have that \(\Delta \big (f\big ) H_{n';l}\in W^{m,p}_{\delta +2-n'}\). This, together with the estimate (3a) implies that \(\Delta \big (f\big ) H_{n';l}=O\big (1/r^{\delta +(d/p)+2-n'}\big )\), and hence \(\Delta \big (f\big ) H_{n';l}\in L^1(\mathbb {R}^d)\) by the estimate \(\delta +(d/p)+2-n'>d\). By arguing in the same way as above, we also have

$$\begin{aligned}&\int _{\mathbb {R}^d}\Delta \big (f\big ) H_{n';l}\,\mathrm{d}x=\lim _{R\rightarrow \infty } \int _{S^{d-1}_R}\Big (\frac{\partial f}{\partial r}\,H_{n';l}-f\,\frac{\partial H_{n';l}}{\partial r}\Big )\,d\sigma _R\nonumber \\&\quad =\lim _{R\rightarrow \infty }\int _{S^{d-1}_R}O\big (1/R^{\delta +(d/p)-n'+1}\big )\,d\sigma _R= \lim _{R\rightarrow \infty }O\big (1/R^{\delta +(d/p)-n}\big )=0, \end{aligned}$$
(89)

where we used that \(n<\delta +d/p\) and the estimate (3a) on the decay of f at infinity. Finally, the integral formula (87) follows from (88), (89), and (86). \(\square \)

Asymptotic spaces \({{\mathcal {A}}}^{m,p}_{n,N;\ell }\)

In this section we will discuss the asymptotic spaces with log terms that are used in the proof of Theorem 1.2. For integers \(m>d/p\), \(0\le n\le N\), and \(\ell \ge -n\), let \({{\mathcal {A}}}^{m,p}_{n,N;\ell }\) denote functions u on \(\mathbb {R}^d\) of the form

$$\begin{aligned} \chi (r)\left( \frac{a_n^0(\theta )+\cdots +a_n^{n+\ell }(\theta )(\log r)^{n+\ell }}{r^n}+\cdots + \frac{a_N^0(\theta )+\cdots +a_N^{N+\ell }(\theta )(\log r)^{N+\ell }}{r^N}\right) +f(x),\nonumber \\ \end{aligned}$$
(90a)

where \(a_k^j\in H^{m+1+N-k,p}(\mathrm{S}^{d-1},\mathbb {R}^d)\) for \(0\le j\le k+\ell \), \(0\le n\le k\le N\), and \(f\in W_{\gamma _N}^{m,p}\) with \(N\le \gamma _N+d/p<N+1\) so, by (3b), \(f(x)=o\big (r^{-N}\big )\) as \(r\rightarrow \infty \). This is a Banach space with norm

$$\begin{aligned} \Vert u\Vert _{{{\mathcal {A}}}^{m,p}_{n,N;\ell }}=\sum _{0\le j\le k+\ell , \, n\le k\le N} \Vert a_k^j\Vert _{H^{m+1+N-k,p}} \ + \ \Vert f\Vert _{W^{m,p}_{\gamma _N}}. \end{aligned}$$
(90b)

Note that \(W_{\gamma _N}^{m,p}\) is a closed subspace of \({{\mathcal {A}}}^{m,p}_{n,N;\ell }\). It is easy to confirm that the following inclusions are bounded for \(N\le \gamma _N+d/p<N=1\):

$$\begin{aligned} {{\mathcal {A}}}^{m,p}_{n_1,N_1;\ell _1} \subseteq {{\mathcal {A}}}^{m,p}_{n,N;\ell } \quad \hbox {if}\ n_1\ge n,\ N_1\ge N,\ \ell \ge \ell _1\ge -n, \end{aligned}$$
(91a)
$$\begin{aligned} \partial _{x_J} : {{\mathcal {A}}}^{m,p}_{n,N;\ell } \rightarrow {{\mathcal {A}}}^{m-1,p}_{n+1,N+1;\ell -1} \quad \hbox {if}\ m>1+d/p,\quad 1\le j\le d. \end{aligned}$$
(91b)

Moreover, if \(n=n_1+n_2\), \(\ell =\ell _1+\ell _2\), and \(N<\min (N_1+n_2,N_2+n_1)\), then pointwise multiplication \((u,v)\mapsto uv\),

$$\begin{aligned} {{\mathcal {A}}}^{m,p}_{n_1,N_1;\ell _1} \times {{\mathcal {A}}}^{m,p}_{n_2,N_2;\ell _2} \rightarrow {{\mathcal {A}}}^{m,p}_{n,N;\ell }, \end{aligned}$$
(91c)

is a bounded bilinear map. When \(\ell _i=-n_i\) there are no log terms in the leading asymptotic and we have the sharper version with \(N=\min (N_1+n_2,N_2+n_1)\), \(n=n_1+n_1\):

$$\begin{aligned} {{\mathcal {A}}}^{m,p}_{n_1,N_1;-n_1} \times {{\mathcal {A}}}^{m,p}_{n_2,N_2;-n_2} \rightarrow {{\mathcal {A}}}^{m,p}_{n,N;\ell }. \end{aligned}$$
(91d)

These may be combined to conclude that

$$\begin{aligned} {{\mathcal {A}}}^{m,p}_{n,N;\ell } \ \hbox {is a Banach algebra} \end{aligned}$$
(91e)

in the case when \(n\ge 1\), or when \(\ell =-n\). For more details, see Appendix B in [23] (for the case when \(N<\gamma _N<N+1\)).

Analogous to (14), for \(m>1+d/p\) we introduce asymptotic diffeomorphisms

$$\begin{aligned} {{{\mathcal {A}}}{\mathcal {D}}}^{m,p}_{n,N;\ell }:=\big \{\varphi : \mathbb {R}^d\rightarrow \mathbb {R}^d\,\big |\,\varphi =\mathrm{id}+w,\,w\in {{\mathcal {A}}}^{m,p}_{n,N;\ell }\, \text { and}\,\det (\mathrm{d}\varphi )>0\big \}. \end{aligned}$$
(92)

Analogous to Theorem 4.1 above, Theorem 6.1 in [24] shows for \(m>3+d/p\) and \(0\le n\le \min (d+1,N)\) that the Euler vector field

$$\begin{aligned} {{\mathcal {E}}}: {{{\mathcal {A}}}{\mathcal {D}}}^{m,p}_{n,N;0} \times {{{\mathcal {A}}}}^{m,p}_{n,N;0} \rightarrow {{{\mathcal {A}}}{\mathcal {D}}}^{m,p}_{n,N;0} \times {{{\mathcal {A}}}}^{m,p}_{n,N;0} \end{aligned}$$
(93)

is smooth. Theorem 1.1 in [24] uses this vector field as we did in Sect. 4 to show the existence of a unique solution as in (43). However, all results in [24] are under the assumption \(N<\gamma _N+d/p<N+1\).

We now show that this solvability of the Euler equations in \({{\mathcal {A}}}^{m,p}_{n,N;0}\) also holds for \(N\le \gamma _N+d/p<N+1\), at least when \(n=1\).

Proposition C.1

Assume \(m>3+d/p,\) \(N\ge 1,\) and \(N\le \gamma _N<N+1\). Then,  for any given \(\rho >0\) there exists \(\tau >0\) such that for any with \(\Vert u_0\Vert _{{{\mathcal {A}}}^{m,p}_{1,N;0}}<\rho \) there exists a unique solution of the Euler equation

(94)

that depends continuously on the initial data \(u_0\).

Proof of Proposition C.1

As previously stated, this was proved in [24] when \(N<\gamma _N+d/p<N+1\), so we need only consider the case when \(\gamma _N+d/p=N\). To do this, let us change notation within this proof and denote by \(\mathfrak {A}^{m,p}_{n,N;0}\) the asymptotic space \({{\mathcal {A}}}^{m,p}_{n,N;\ell }\) with \(\gamma _N+d/p=N\); we reserve the notation \({{\mathcal {A}}}^{m,p}_{n,N;0}\) for an asymptotic space with a remainder \(f\in W^{m,p}_{{\tilde{\gamma }}_N}\) with \(N<{\tilde{\gamma }}_N<N+1\). In particular, referring to (91a), we have the following bounded inclusions for any \(\ell \ge -n\):

$$\begin{aligned} {{\mathcal {A}}}^{m,p}_{n,N;\ell }\subseteq \mathfrak {A}^{m,p}_{n,N;\ell }\subseteq {{\mathcal {A}}}^{m,p}_{n,N-1;\ell }. \end{aligned}$$
(95)

We will need to take the square of elements in \(\mathfrak {A}^{m,p}_{n,N;\ell }\). For \(n+\ell >0\), we know from (91c) and (91d) that the log terms in the leading asymptotic prevent us from keeping the same order of decay; but decreasing this order of decay by one enables us to also embed in a space with \(0<\gamma _0+d/p<1\):

$$\begin{aligned} u\mapsto u^2 \quad \hbox {is smooth} \quad \mathfrak {A}^{m,p}_{n,N;\ell } \rightarrow {{\mathcal {A}}}^{m,p}_{2n,N+n-1;2\ell }. \end{aligned}$$
(96)

Using (91b), for \(u\in \mathfrak {A}^{m,p}_{1,N;0}\) we have \(\mathrm{d}u\in \mathfrak {A}^{m-1,p}_{2,N+1;-1}\) so (96) implies that \(Q(u)\equiv \mathrm{tr}\,\big (\mathrm{d}u\big )^2\) defines a bounded quadratic polynomial map \(Q :\mathfrak {A}^{m,p}_{1,N;0}\rightarrow {{\mathcal {A}}}^{m-1,p}_{4,N+2;-2}\subseteq {{\mathcal {A}}}^{m-1,p}_{2,N+2;-2}\). Consequently, the maps

$$\begin{aligned} Q : \mathfrak {A}^{m,p}_{1,N;0}\rightarrow {{\mathcal {A}}}^{m-1,p}_{2,N+2;-2}\quad \text { and}\quad \nabla \circ Q : \mathfrak {A}^{m,p}_{1,N;0}\rightarrow {{\mathcal {A}}}^{m-2,p}_{3,N+3;-3} \end{aligned}$$
(97)

are smooth. We can now apply Proposition 3.1 and (17b) in [24] to conclude that the linear map \(\Delta ^{-1}: {{\mathcal {A}}}^{m-2,p}_{3,N+3;-3}\rightarrow {{\mathcal {A}}}^{m,p}_{1,N+1;0}\) is bounded and injective. Combined with (97) and the embeddings \({{\mathcal {A}}}^{m,p}_{1,N+1;0}\subseteq {{\mathcal {A}}}^{m,p}_{1,N;0} \subseteq \mathfrak {A}^{m,p}_{1,N;0}\) we see that

$$\begin{aligned} \Delta ^{-1}\circ \nabla \circ Q : \mathfrak {A}^{m,p}_{1,N;0}\rightarrow \mathfrak {A}^{m,p}_{1,N;0} \end{aligned}$$

is smooth. The arguments in [24, Section 4, 5, and 6] (cf. Lemma 5.2, Proposition 5.1, Lemma 6.1, and Theorem 6.1 in [24]) then show that the associated conjugate map

$$\begin{aligned}&(\varphi ,f)\mapsto \big (R_\varphi \circ \Delta ^{-1}\circ R_{\varphi ^{-1}}\big )\circ \big (R_{\varphi }\circ \nabla \circ Q\circ R_{\varphi ^{-1}}\big )(f),\\&\quad \mathfrak {A}{{\mathcal {D}}}^{m,p}_{1,N;0}\times \mathfrak {A}^{m,p}_{1,N;0}\rightarrow \mathfrak {A}{{\mathcal {D}}}^{m,p}_{1,N;0}\times \mathfrak {A}^{m,p}_{1,N;0}, \end{aligned}$$

is smooth. This implies that the Euler vector field \(\mathcal {E}\) is smooth as a map

$$\begin{aligned} \mathcal {E} : {\mathcal {A}D}^{m,p}_{1,N;0}\times {{\mathcal {A}}}^{m,p}_{1,N;0}\rightarrow {\mathcal {A}D}^{m,p}_{1,N;0}\times {{\mathcal {A}}}^{m,p}_{1,N;0}. \end{aligned}$$
(98)

The arguments in the proof of Theorem 1.1 above then complete the proof of the proposition (cf. also Section 7 in [24]). \(\square \)

Global existence in the case when \(d=2\)

In this section we generalize Theorem 1.1 in [30] and prove that for \(d=2\) the solution of the Euler equation (1) has a unique global in time solution in the asymptotic space \({{\mathcal {Z}}}^{m,p}_N\) with weight \(\gamma _N\) such that \(\gamma _N+d/p>0\) is integer.Footnote 14 For a given \(a\in \mathbb {R}\) denote by \(\lfloor a \rfloor \) the integer part of a. We will follow the notation introduced in [30, Section 2].

For a given \(1<p<\infty \), \(m>2/p\), and \(\delta +2/p>0\) we set \(N:=\lfloor \delta +2/p \rfloor \), \(\gamma _N:=\delta \), and consider the space of complex valued functions of \(z\in \mathbb {C}\),

$$\begin{aligned} \mathcal {Z}^{m,p}_{n,N}:=\Big \{\chi \sum _{n\le k+l\le N}\frac{a_{kl}}{z^k{{\bar{z}}}^l}+f\,\Big |\,f\in W^{m,p}_{\gamma _N}, a_{kl}\in \mathbb {C}\Big \}, \end{aligned}$$
(99)

where \(0\le n\le N+1\) and where we omit the summation term if \(n=N+1\) and set \(\mathcal {Z}^{m,p}_{n,N}\equiv W^{m,p}_{\gamma _N}\). We also set \(\mathcal {Z}^{m,p}_N\equiv \mathcal {Z}^{m,p}_{0,N}\). The space (99) is a closed subspace in the asymptotic space \({{\mathcal {A}}}^{m,p}_N\) of vector fields on \(\mathbb {R}^2\) that satisfies Proposition 2.1 and 2.2 in [30]. Note however, that Proposition 3.3 and Theorem 3.2 in [30] does not hold for integer \(\delta +2/p\). As a consequence, the proof of the global well-posedness of the Euler equation for \(d=2\) in [30, Section 5] does not apply for integer values of \(\delta +2/p\). Following [30, Section 2] we denote the group of diffeomorphisms of \(\mathbb {R}^2\) modeled on \(\mathcal {Z}^{m,p}_N\) by \({{\mathcal {Z}}}D^{m,p}_N\). First, we prove the following lemma.

Lemma D.1

Take \(m>3+2/p,\) a non-integer \(\delta +2/p>0,\) and let \({\hat{\delta }}\) be the lowest integer \({\hat{\delta }}>\delta \) such that \({\hat{\delta }}+2/p\in \mathbb {Z}\). Then,  for a given volume preserving \(\varphi \in {{\mathcal {Z}}}D^{m,p}_M\) (with \(\gamma _M:=\delta ,\) \(M:=\lfloor \delta +2/p \rfloor )\) and \(u_0\in \mathcal {Z}^{m,p}_N\) (with \(\gamma _N:={\hat{\delta }},\) \(N:={{\hat{\delta }}}+2/p)\) we have that

$$\begin{aligned} \big (R_\varphi \circ \partial _z^{-1}\circ R_{\varphi ^{-1}}\big )(\partial _z u_0)=u_0+\mathcal {R}(\varphi ,u_0),\quad \mathcal {R}(\varphi ,u_0)\in \mathcal {Z}^{m,p}_{1,M+1}, \end{aligned}$$
(100)

where the map \(\mathcal {R} : {{\mathcal {Z}}}D^{m,p}_M\times {{\mathcal {Z}}}^{m,p}_N\rightarrow {{\mathcal {Z}}}^{m,p}_{M+1}\) is analytic and \(\partial _z\) denotes the Cauchy operator \(\partial _z : \mathcal {Z}^{m,p}_{1,M}\rightarrow \widetilde{\mathcal {Z}}^{m-1,p}_{M+1}\).Footnote 15

Proof of Lemma D.1

Since \(u_0\in \mathcal {Z}^{m,p}_{N}\) (with \(\gamma _N={\hat{\delta }}\)) we have that \(\partial _z u_0\in \widetilde{\mathcal {Z}}^{m-1,p}_{1,M+1}\) and, by Proposition 3.4 in [30], \(\big (R_\varphi \circ \partial _z^{-1}\circ R_{\varphi ^{-1}}\big )(\partial _z u_0)\) is well defined and belongs to \(\mathcal {Z}^{m,p}_{1,M}\). By setting \(w:=\big (R_\varphi \circ \partial _z^{-1}\circ R_{\varphi ^{-1}}\big )(\partial _z u_0)\) we then obtain from Lemma 2.4 in [30] that

$$\begin{aligned} \big (R_\varphi \circ \partial _z\circ R_{\varphi ^{-1}}\big )(w)=\partial _z u_0. \end{aligned}$$

This, together with formula (54) in [30] and the fact that \(\varphi =\mathrm{id}_\mathbb {C}+u\in {{\mathcal {Z}}}D^{m,p}_M\) is volume preserving, then implies that \(\partial _z w+(\partial _z w)(\partial _{{\bar{z}}}{{\bar{u}}}_0)-(\partial _{{\bar{z}}}w)(\partial _z{{\bar{u}}}_0)=\partial _z u_0\), or equivalently,

$$\begin{aligned} w=u_0+\partial _z^{-1}\big [(\partial _{{\bar{z}}}w)(\partial _z{{\bar{u}}}_0)-(\partial _z w)(\partial _{{\bar{z}}}{{\bar{u}}}_0)\big ]. \end{aligned}$$
(101)

Since, by Lemma 3.5 in [30], \(w\equiv \big (R_\varphi \circ \partial _z^{-1}\circ R_{\varphi ^{-1}}\big )(\partial _z u_0)\in \mathcal {Z}^{m,p}_{1,M}\) depends analytically on \((\varphi ,u_0)\in {{\mathcal {Z}}}D^{m,p}_M\times {{\mathcal {Z}}}D^{m,p}_N\), we obtain from Proposition 2.2 in [30] that

$$\begin{aligned} (\partial _{{\bar{z}}}w)(\partial _z{{\bar{u}}}_0)-(\partial _z w)(\partial _{{\bar{z}}}{{\bar{u}}}_0)\in \widetilde{\mathcal {Z}}^{m-1,p}_{M+2} \end{aligned}$$

and depends analytically on \((\varphi ,u_0)\in {{\mathcal {Z}}}D^{m,p}_M\times {{\mathcal {Z}}}^{m,p}_N\). By combining this with Theorem 3.2 in [30] we then see that

$$\begin{aligned} \mathcal {R}(\varphi ,u_0):=\partial _z^{-1}\big [(\partial _{{\bar{z}}}w)(\partial _z{{\bar{u}}}_0)-(\partial _z w)(\partial _{{\bar{z}}}{{\bar{u}}}_0)\big ] \in \mathcal {Z}^{m,p}_{1,M+1} \end{aligned}$$

and depends analytically on \((\varphi ,u_0)\in {{\mathcal {Z}}}D^{m,p}_M\times {{\mathcal {Z}}}^{m,p}_N\). This completes the proof of the lemma. \(\square \)

Now, we are ready to prove

Proposition D.1

Assume that \(m>3+2/p,\) \(\delta +2/p>0\) is an integer,  and \(d=2\). Then,  for any (with \(\gamma _N:=\delta \) and \(N:=\delta +2/p)\) the Euler equation (1) has a unique global in time solution that depends continuously on the initial data (cf. [30, Theorem 1.1] for the case when \(\gamma _N+2/p\) is not integer).

Proof of Proposition D.1

Assume that \(\delta +2/p>0\) is an integer and choose \(\delta ^-\in \mathbb {R}\) such that \(0<\delta -\delta ^-<1\) and \(\delta ^-+2/p>0\). Take \(u_0\in {{\mathcal {Z}}}^{m,p}_N\) (with \(\gamma _N=\delta \)). Since \({{\mathcal {Z}}}^{m,p}_N\) is a subspace in \({{\mathcal {Z}}}^{m,p}_M\) (with \(\gamma _M:=\delta ^-\) and \(M:=\lfloor \delta ^-+2/p \rfloor =N-1\)) and since \(\delta ^-+2/p\) is not integer, we conclude from [30, Theorem 1.1] that there exists a unique solution of the Euler equation

$$\begin{aligned} u\in C\big ([0,\infty ),\mathcal {Z}^{m,p}_M\big )\cap C^1\big ([0,\infty ),\mathcal {Z}^{m-1,p}_M\big ) \end{aligned}$$

that depends continuously on the initial data \(u_0\in \mathcal {Z}^{m,p}_N\). By [30, Proposition 4.2], \(\varphi \in C^1\big ([0,\infty ),\mathcal {Z}D^{m,p}_M\big )\) where , \(\varphi |_{t=0}=u_0\), and it depends continuously on the initial data \(u_0\in \mathcal {Z}^{m,p}_N\). The preservation of vorticity (cf. formula (76) in [30]) and Lemma D.1 then imply that

$$\begin{aligned} u&=\partial _z^{-1}\big ((\partial _z u_0)\circ \varphi ^{-1}\big )= R_{\varphi ^{-1}}\circ \big (R_\varphi \circ \partial _z^{-1}\circ R_{\varphi ^{-1}}\big )(\partial _z u_0)\nonumber \\&=u_0\circ \varphi ^{-1}+\mathcal {R}(\varphi ,u_0)\circ \varphi ^{-1} \end{aligned}$$
(102)

where \(\mathcal {R}(\varphi ,u_0)\in \mathcal {Z}^{m,p}_{1,M+1}\) and it depends analytically on \((\varphi ,u_0)\in {{\mathcal {Z}}}D^{m,p}_M\times {{\mathcal {Z}}}^{m,p}_N\). Since \(\gamma _M+1>\delta \) we have that \(\mathcal {Z}^{m,p}_{M+1}\subseteq \mathcal {Z}^{m,p}_N\). By Proposition 2.3 and Proposition 2.4 in [30] we then obtain that

$$\begin{aligned} u\in C\big ([0,\infty ),\mathcal {Z}^{m,p}_N\big )\cap C^1\big ([0,\infty ),\mathcal {Z}^{m-1,p}_N\big ) \end{aligned}$$

and it depends continuously on the initial data \(u_0\in \mathcal {Z}^{m,p}_N\). This completes the proof of the proposition. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

McOwen, R., Topalov, P. Perfect fluid flows on \(\mathbb {R}^d\) with growth/decay conditions at infinity. Math. Ann. 383, 1451–1488 (2022). https://doi.org/10.1007/s00208-021-02248-w

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00208-021-02248-w

Mathematics Subject Classification

Navigation