Skip to main content
Log in

Global well-posedness and blow-up of solutions for the Camassa–Holm equations with fractional dissipation

  • Published:
Mathematische Zeitschrift Aims and scope Submit manuscript

Abstract

Consideration in this paper is the effect of varying fractional dissipation with the dissipative operator power \( \gamma \ge 0\) on the well-posedness of the Camassa–Holm equations with fractional dissipation. It is shown that the zero-filter limit (\(\alpha \rightarrow 0\)) of the Camassa–Holm equation with fractional dissipation is the fractal Burgers equation. It is known that in the supercritical case \(\gamma \in [0, 1)\), the fractal Burgers equation blows up in finite time in \(H^s({\mathbb {R}})\) with \(s>\frac{3}{2}-\gamma \). It is established here that the dissipative Camassa–Holm equation is globally well-posed in the critical Sobolev space \(H^{\frac{3}{2}-\gamma }({\mathbb {R}})\) with the fractional parameter \(\gamma \in [\frac{1}{2}, 1)\). Moreover, it is also demonstrated that the solution of the dissipative Camassa–Holm equation blows up in finite time in the particular supercritical case when \( \gamma = 0\).

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.

We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

References

  1. Alber, M.S., Camassa, R., Holm, D.D., Marsden, J.E.: The geometry of peaked solitons and billiard solutions of a class of integrable PDE’s. Lett. Math. Phys. 32, 137–151 (1994)

    Article  MATH  MathSciNet  Google Scholar 

  2. Bahouri, H., Chemin, J.Y., Danchin, R.: Fourier Analysis and Nonlinear Partial Differential Equations, Grundlehren der mathematischen Wissenschaften 343. Springer, Berlin (2011)

    Book  Google Scholar 

  3. Bressan, A., Chen, G., Zhang, Q.: Uniqueness of conservative solutions to the Camassa–Holm equation via characteristics. Discrete Contin. Dyn. Syst. 35(1), 25–42 (2014)

    Article  MathSciNet  Google Scholar 

  4. Bressan, A., Constantin, A.: Global conservative solutions of the Camassa–Holm equation. Arch. Ration. Mech. Anal. 183, 215–239 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  5. Bressan, A., Constantin, A.: Global dissipative solutions of the Camassa–Holm equation. Anal. Appl. 5, 1–27 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  6. Bony, J.M.: Calcul symbolique et propagation des singularités pour les q́uations aux drivées partielles non linéaires. Ann. Sci. École Norm. Sup. (4) 14, 209–246 (1981)

    MATH  MathSciNet  Google Scholar 

  7. Camassa, R., Holm, D.D.: An integrable shallow water equation with peaked solitons. Phys. Rev. Lett. 71, 1661–1664 (1993)

    Article  MATH  MathSciNet  Google Scholar 

  8. Chan, C., Czubak, M.: Regularity of solutions for the critical N-dimensional Burgers’ equation. Annales de l’Institut Henri Poincare (C) Non Linear. Analysis 27(2), 471–501 (2010)

  9. Chemin, J.-Y.: Perfect Incompressible Fluids. Oxford University Press, New York (1998)

    MATH  Google Scholar 

  10. Chemin, J.-Y.: Théorèmes d’unicité pour le système de Navier–Stokes tridimensionnel. J. Anal. Math. 77, 25–50 (1999)

    Article  MathSciNet  Google Scholar 

  11. Constantin, A.: Existence of permanent and breaking waves for a shallow water equation: a geometric approach. Ann. Inst. Fourier (Grenoble) 50, 321–362 (2000)

    Article  MATH  MathSciNet  Google Scholar 

  12. Constantin, A., Lannes, D.: The hydrodynamical relevance of the Camassa–Holm and Degasperis–Procesi equations. Arch. Ration. Mech. Anal. 192, 165–186 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  13. Constantin, A., Escher, J.: Wave breaking for nonlinear nonlocal shallow water equations. Acta Math. 181, 229–243 (1998)

    Article  MATH  MathSciNet  Google Scholar 

  14. Constantin, A., Escher, J.: Global existence and blow-up for a shallow water equation. Ann. Scuola Norm. Sup. Pisa 26, 303–328 (1998)

    MATH  MathSciNet  Google Scholar 

  15. Constantin, A., Escher, J.: On the blow-up rate and the blow-up set of breaking waves for a shallow water equation. Math. Z. 233, 75–91 (2000)

    Article  MATH  MathSciNet  Google Scholar 

  16. Danchin, R.: A few remarks on the Camassa–Holm equation. Differ. Integr. Equ. 14, 953–988 (2001)

    MATH  MathSciNet  Google Scholar 

  17. Danchin, R.: Local and global well-posedness results for flows of inhomogeneous viscous fluids. Adv. Differ. Equ. 9(3–4), 353–386 (2004)

    MATH  MathSciNet  Google Scholar 

  18. Dong, H., Du, D., Li, D.: Finite time singularities and global well-posedness for fractal Burgers equations. Indiana Univ. Math. J. 58(2), 807–821 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  19. Foias, C., Holm, D.D., Titi, E.S.: The Navier–Stokes-alpha model of fluid turbulence. Phys. D 152/153, 505–519 (2001)

  20. Fisher, M., Schiff, J.: The Camassa Holm equation: conserved quantities and the initial value problem. Phys. Lett. A 259, 371–376 (1999)

    Article  MATH  MathSciNet  Google Scholar 

  21. Fuchssteiner, B., Fokas, A.S.: Symplectic structures, their Bäcklund transformation and hereditary symmetries. Phys. D 4, 47–66 (1981)

    Article  MATH  MathSciNet  Google Scholar 

  22. Holm, D.D., Marsden, J.E., Ratiu, T.S.: The Euler–Poincare equations and semidirect products with applications to continuum theories. Adv. Math. 137(1), 1–81 (1998)

    Article  MATH  MathSciNet  Google Scholar 

  23. Holm, D.D., Titi, E.S.: Computational models of turbulence: the LANS-alpha model and the role of global analysis. SIAM News 38, 1–5 (2005)

    Google Scholar 

  24. Johnson, R.S.: Camassa–Holm, Korteweg–de Vries and related models for water waves. J. Fluid Mech. 455, 63–82 (2002)

    Article  MATH  MathSciNet  Google Scholar 

  25. Kiselev, A., Nazarov, F., Shterenberg, R.: Blow up and regularity for fractal Burgers equation. Dyn. Partial Differ. Equ. 5(3), 211–240 (2008)

    Article  MATH  MathSciNet  Google Scholar 

  26. Li, Y.A., Olver, P.J.: Well-posedness and blow-up solutions for an integrable nonlinearly dispersive model wave equation. J. Differ. Equ. 162, 27–63 (2000)

    Article  MATH  MathSciNet  Google Scholar 

  27. Miao, C., Wu, G.: Global well-posedness of the critical Burgers equation in critical Besov spaces. J. Differ. Equ. 247, 1673–1693 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  28. Triebel, H.: Theory of Function Spaces, Monograph in Mathematics, vol. 78. Birkhauser, Basel (1983)

    Book  Google Scholar 

  29. Xin, Z., Zhang, P.: On the weak solutions to a shallow water equation. Commun. Pure Appl. Math. 53, 1411–1433 (2000)

    Article  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgments

The authors would like to express their gratitude to the referees for valuable comments and suggestions. The work of Gui is partially supported by the NNSF-China under the Grants 11331005, 2014JQ1009 and 14JK1757. The work of Liu is partially supported by the NSF Grant DMS-1207840 and the NNSF-China Grant-11271192.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yue Liu.

Appendix: Littlewood–Paley analysis

Appendix: Littlewood–Paley analysis

The proofs of Theorems 1.11.5 require the Littlewood–Paley decomposition, or a dyadic decomposition of the Fourier variables, which may be explained how it may be built in the case \(x\in {\mathbb {R}}^d\) (see e.g. [2, 9, 28]) as follows.

We recall the following proposition about the smoothing effect of the operator \(\partial _t+\Lambda ^{\gamma }\) with \(t>0\) and \(\gamma >0\), which was obtained from [10] (up to a slight modification).

Proposition 6.1

([10]) Let \({\mathcal {C}}\) be an annulus. Positive constants c and C exist such that for any p in \([1,+\infty ]\) and any couple \((t, \lambda )\) of positive real numbers, we have

$$\begin{aligned} \mathop {\mathrm{Supp}}\nolimits \,\widehat{u} \subset \lambda {\mathcal {C}} \, \Rightarrow \, \Vert e^{-t \Lambda ^{\gamma }} u\Vert _{L^p} \le C e^{-ct\lambda ^{\gamma }}\Vert u\Vert _{L^p}. \end{aligned}$$

Let us now recall a dyadic partition of unity, which may be found in [2]. Let us first define by \({\mathcal {C}}\) the ring of center 0, of small radius 3 / 4 and great radius 8 / 3. It exists two radial functions \(\chi \) and \(\varphi \) the values of which are in the interval [0, 1], belonging respectively to \({\mathcal {D}}(B(0,4/3))\) and to \({\mathcal {D}}({\mathcal {C}})\) such that

$$\begin{aligned}&\chi (\xi ) + \sum _{j\ge 0} \varphi (2^{-j}\xi ) = 1 \quad (\forall \xi \in {\mathbb {R}}^d), \qquad \sum _{j\in {\mathbb {Z}}} \varphi \left( 2^{-j}\xi \right) = 1 \quad \left( \forall \xi \in {\mathbb {R}}^d\setminus \{0\}\right) ,\nonumber \\&|j-j'|\ge 2 \Rightarrow \mathop {\mathrm{Supp}}\nolimits \,\varphi (2^{-j}\cdot )\cap \mathop {\mathrm{Supp}}\nolimits \,\varphi (2^{-j'}\cdot )=\emptyset ,\\&j\ge 1 \Rightarrow \mathop {\mathrm{Supp}}\nolimits \,\chi \cap \mathop {\mathrm{Supp}}\nolimits \,\varphi (2^{-j}\cdot ) = \emptyset .\nonumber \end{aligned}$$
(6.1)

Setting \(h\mathop {=}\limits ^{\mathrm{def}}{{{\mathcal {F}}}}^{-1}\varphi ,\) \(\widetilde{h}\mathop {=}\limits ^{\mathrm{def}}{{{\mathcal {F}}}}^{-1}\chi \). The inhomogeneous dyadic blocks \({\Delta }_j\) and the inhomogeneous low-frequency cut-off operator \({S}_j\) are defined for all \( j \in {\mathbb {N}}\cup \{-1\}\) by

$$\begin{aligned} \begin{aligned}&{\Delta }_{j} f\mathop {=}\limits ^{\mathrm{def}}\varphi (2^{-j}D)f=2^{j d}\int _{{\mathbb {R}}^{d}}h(2^j y)f(x-y)dy \quad (\forall \quad j \ge 0),\\&{\Delta }_{-1} f\mathop {=}\limits ^{\mathrm{def}}\chi (D)f=\int _{{\mathbb {R}}^{d}}\widetilde{h}( y)f(x-y)dy,\quad {\Delta }_{j} f\mathop {=}\limits ^{\mathrm{def}}0 \quad (\forall \quad j \le -2),\\&{S}_{j} f\mathop {=}\limits ^{\mathrm{def}}\sum _{ j^{\prime } \le j-1}{\Delta }_{j^{\prime }} f=\chi (2^{-j}D)f=2^{j d}\int _{{\mathbb {R}}^{d}}\widetilde{h}(2^j y)f(x-y)dy. \end{aligned} \end{aligned}$$
(6.2)

The homogeneous dyadic blocks \(\dot{\Delta }_j\) and the homogeneous low-frequency cut-off operators \(\dot{S}_j\) are defined for all \( j \in {\mathbb {Z}}\) by

$$\begin{aligned} \begin{aligned}&\dot{\Delta }_{j} f\mathop {=}\limits ^{\mathrm{def}}\varphi (2^{-j}D)f=2^{j d}\int _{{\mathbb {R}}^{d}}h(2^j y)f(x-y)dy, \quad \dot{S}_{j} f\mathop {=}\limits ^{\mathrm{def}}\sum _{ j^{\prime } \le j-1}\dot{\Delta }_{j^{\prime }} f. \end{aligned} \end{aligned}$$
(6.3)

We should point out that all the above operators \(\Delta _j\), \(\dot{\Delta }_j\), \(S_{j}\), and \(\dot{S}_{j}\) map \(L^p\) into \(L^p\) with norms which do not depend on j.

With above notations in hand, we are in a position to define Besov spaces. Let \(s\in {\mathbb {R}},\, 1\le p,\,r\le \infty .\) The inhomogenous (or homogeneous) Besov space \({B}^s_{p,r}({\mathbb {R}}^{d})\) (or \(\dot{B}^s_{p,r}({\mathbb {R}}^{d})\)) is defined by

$$\begin{aligned}&{B}^s_{p,r}({\mathbb {R}}^{d})\mathop {=}\limits ^{\mathrm{def}}\{f\in {\mathcal {S}}^{\prime }({\mathbb {R}}^{d})| \, \Vert f\Vert _{{B}^s_{p,r}}<\infty \}\,\bigg (\text{ or }\quad \dot{B}^s_{p,r}({\mathbb {R}}^{d})\mathop {=}\limits ^{\mathrm{def}}\{f\in {\mathcal S}^{\prime }_{h}({\mathbb {R}}^{d})| \, \Vert f\Vert _{\dot{B}^s_{p,r}}<\infty \}\bigg ) \, \text{ with }\\&\Vert f\Vert _{{B}^s_{p,r}}\mathop {=}\limits ^{\mathrm{def}}{\left\{ \begin{array}{ll} \bigg (\sum _{j \in {\mathbb {Z}}} 2^{j s r}\Vert {\Delta }_j f\Vert _{L^p}^r\bigg )^{\frac{1}{r}},\quad \hbox {if}\quad r<\infty ,\\ \sup _{j \in {\mathbb {Z}}}2^{j s}\Vert {\Delta }_j f\Vert _{L^p}, \quad \quad \ \quad \hbox {if} \quad r=\infty , \end{array}\right. } (\text{ or }\\&\Vert f\Vert _{\dot{B}^s_{p,r}}\mathop {=}\limits ^{\mathrm{def}}{\left\{ \begin{array}{ll} \bigg (\sum _{j \in {\mathbb {Z}}} 2^{j s r}\Vert \dot{\Delta }_j f\Vert _{L^p}^r\bigg )^{\frac{1}{r}},\quad \hbox {for}\quad r<\infty ,\\ \sup _{j \in {\mathbb {Z}}}2^{j s}\Vert \dot{\Delta }_j f\Vert _{L^p}, \quad \quad \ \quad \hbox { for} \quad r=\infty , \end{array}\right. }\text{ and }\\&{\mathcal {S}}^{\prime }_{h}({\mathbb {R}}^{d}) \mathop {=}\limits ^{\mathrm{def}}\{f \in {\mathcal S}^{\prime }({\mathbb {R}}^{d})| \, \lim _{j \rightarrow -\infty }\dot{S}_{j} f =0 \quad \hbox {in} \quad {\mathcal {S}}^{\prime }({\mathbb {R}}^{d}) \}). \end{aligned}$$

If \(s=\infty \), \(B^{\infty }_{p, r} \mathop {=}\limits ^{\mathrm{def}}\bigcap _{s \in {\mathbb {R}}} B^{s}_{p, r} .\)

Remark 6.1

  1. (i)

      Let \(s\in {\mathbb {R}}, \,1\le p,\,r\le \infty \), and \(u \in {\mathcal S}'_{h}.\) Then u belongs to \(\dot{B}^{s}_{p, r}\) if and only if there exists \(\{c_{j, r}\}_{j \in {\mathbb {Z}}} \) such that \(\Vert c_{j, r}\Vert _{\ell ^{r}} =1\) and \( \Vert \dot{\Delta }_{j}u\Vert _{L^{p}}\le C c_{j, r} 2^{-j s } \Vert u\Vert _{\dot{B}^{s}_{p, r}}. \)

  2. (ii)

      Let \(s\in {\mathbb {R}},\, 1\le p,\,r\le \infty \). Then u belongs to \({B}^{s}_{p, r}\) if and only if there exists \(\{c_{j, r}\}_{j \in {\mathbb {N}} \cup \{-1\}} \) such that \(\Vert c_{j, r}\Vert _{\ell ^{r}} =1\) and \( \Vert {\Delta }_{j}u\Vert _{L^{p}}\le C c_{j, r} 2^{-j s } \Vert u\Vert _{{B}^{s}_{p, r}}. \)

We should point out that if \(s>0\) then \(B^s_{p,r}({\mathbb {R}}^d)=\dot{B}^s_{p,r}({\mathbb {R}}^d)\cap L^p({\mathbb {R}}^d)\) and \( \Vert u\Vert _{B^s_{p,r}}\approx \Vert u\Vert _{\dot{B}^s_{p,r}}+\Vert u\Vert _{L^p}. \) It is easy to verify that the homogeneous Besov space \(\dot{B}^s_{2,2}({\mathbb {R}}^d)\) coincides with the classical homogeneous Sobolev space \(\dot{H}^{s}({\mathbb {R}}^d)\). Similar assertions may be easily understood for the inhomogeneous Besov spaces.

In order to obtain a better description of the regularizing effect of the transport-diffusion equation, we will use Chemin–Lerner type spaces \(\widetilde{L}^{\lambda }_T(B^s_{p,r}({\mathbb {R}}^d))\) from [10].

Definition 6.1

Let \(s\in {\mathbb {R}}\), \((r,\lambda ,p)\in [1,\,+\infty ]^3\) and \(T\in (0,\,+\infty ]\). We define \(\widetilde{L}^{\lambda }_T( B^s_{p\,r}({\mathbb {R}}^d))\) as the completion of \(C([0,T],{\mathcal {S}}({\mathbb {R}}^d))\) by the norm \( \Vert f\Vert _{\widetilde{L}^{\lambda }_T( B^s_{p,r})} \mathop {=}\limits ^{\mathrm{def}}\Big (\sum _{q\in {{\mathbb {Z}}}}2^{qrs} \Big (\int _0^T\Vert \Delta _q\,f(t) \Vert _{L^p}^{\lambda }\, dt\Big )^{\frac{r}{\lambda }}\Big )^{\frac{1}{r}} <\infty \) with the usual change if \(r=\infty .\) For short, we just denote this space by \(\widetilde{L}^{\lambda }_T( B^s_{p,r}).\) In the particular case when \(p=r=2,\) we denote this space by \(\widetilde{L}^\lambda _T({H}^s).\) We also denote the space \({\mathcal {C}}([0, T]; B^s_{p,r}) \cap \widetilde{L}^{\infty }_T(B^s_{p,r})\) by \(\widetilde{{\mathcal {C}}}([0, T]; B^s_{p,r})\).

It is easy to observe that for \(\theta \in [0,1],\) we have \( \Vert u\Vert _{\widetilde{L}^{\lambda }_T( B^s_{p,r})} \le \Vert u\Vert _{\widetilde{L}^{\lambda _1}_T( B^{s_1}_{p,r})}^{\theta } \Vert u\Vert _{\widetilde{L}^{\lambda _2}_T(B^{s_2}_{p,r})}^{1-\theta } \) with \(\frac{1}{\lambda }=\frac{\theta }{\lambda _1} +\frac{1-\theta }{\lambda _2}\) and \(s=\theta s_1+(1-\theta )s_2.\) Moreover, Minkowski inequality implies that

$$\begin{aligned} \Vert u\Vert _{\widetilde{L}^{\lambda }_T(B^s_{p,r})} \le \Vert u\Vert _{L^{\lambda }_T(B^s_{p,r})} \quad \text{ if }\quad \lambda \le r \quad \hbox {and}\quad \Vert u\Vert _{L^{\lambda }_T( B^s_{p,r})} \le \Vert u\Vert _{\widetilde{L}^{\lambda }_T(B^s_{p,r})} \quad \text{ if }\quad r\le \lambda . \end{aligned}$$

Similar definitions and the above properties may be easily understood for the homogeneous version of \(\widetilde{L}^{\lambda }_T(\dot{B}^s_{p,r}({\mathbb {R}}^d))\).

In what follows, we shall frequently use Bony’s decomposition [6] in the both homogeneous and inhomogeneous context:

$$\begin{aligned} \begin{aligned}&uv=\dot{T}_u v+\dot{R}(u,v)=\dot{T}_u v+\dot{T}_v u+\dot{{\mathcal {R}}}(u,v)\qquad \text{ and } \\&uv=T_u v+R(u,v)=T_u v+T_v u+{\mathcal {R}}(u,v), \end{aligned} \end{aligned}$$
(6.4)

where

$$\begin{aligned} \begin{aligned}&\dot{T}_u v\mathop {=}\limits ^{\mathrm{def}}\sum _{q \in {\mathbb {Z}}}\dot{S}_{q-1}u\dot{\Delta }_q v,\qquad \dot{R}(u,v)\mathop {=}\limits ^{\mathrm{def}}\sum _{q\in {{\mathbb {Z}}}}\dot{\Delta }_q u \dot{S}_{q+2}v,\\&\dot{{\mathcal {R}}}(u,v)\mathop {=}\limits ^{\mathrm{def}}\sum _{q\in {{\mathbb {Z}}}}\dot{\Delta }_q u \widetilde{\dot{\Delta }}_{q}v\quad \text{ and }\quad \widetilde{\dot{\Delta }}_{q}v\mathop {=}\limits ^{\mathrm{def}}\sum _{|q'-q|\le 1}\dot{\Delta }_{q'}v, \end{aligned} \end{aligned}$$

and similar definitions for the inhomogeneous version of \(T_uv\), R(uv), and \({\mathcal {R}}(u,v).\)

From this, we readily get the following product laws in Besov spaces.

Lemma 6.1

Assume that \(1 \le p, \, r \le +\infty ,\) the following estimates hold:

  1. (i).

    for \(s>0\), \(\Vert fg\Vert _{B^{s}_{p, r}}\le C (\Vert f\Vert _{B^{s}_{p, r}}\Vert g\Vert _{L^{\infty }}+ \Vert g\Vert _{B^{s}_{p, r}}\Vert f\Vert _{L^{\infty }});\)

  2. (ii).

    for \(|s| < \frac{d}{2}\), \( \Vert fg\Vert _{H^{s}({\mathbb {R}}^d)}\le C \Vert f\Vert _{B^{\frac{d}{2}}_{2,1}({\mathbb {R}}^d)}\Vert g\Vert _{H^{s}({\mathbb {R}}^d)}; \)

  3. (iii).

    for \(s_1 \le \frac{d}{p},\, s_2>\frac{d}{p}\) (\( s_2\ge \frac{d}{p}\) if \(r=1\)) and \(s_1+s_2>0,\)

    $$\begin{aligned} \Vert fg\Vert _{B^{s_1}_{p, r}({\mathbb {R}}^d)}\le C \Vert f\Vert _{B^{s_1}_{p, r}({\mathbb {R}}^d)}\Vert g\Vert _{B^{s_2}_{p, r}({\mathbb {R}}^d)}; \end{aligned}$$
  4. (iv).

    for any \((s_1, s_2) \in (-d/2, d/2)\) and \(s_1+s_2>0\),

    $$\begin{aligned} \Vert fg\Vert _{\dot{B}^{s_1+s_2-\frac{d}{2}}_{2, 1}({\mathbb {R}}^d)}\le C \Vert f\Vert _{\dot{H}^{s_1}({\mathbb {R}}^d)}\Vert g\Vert _{\dot{H}^{s_2}({\mathbb {R}}^d)}, \end{aligned}$$

where C’s are constants independent of f and g.

The following basic lemma will be of constant use in this paper.

Lemma 6.2

(Lemma 2.97 in [2]) (Commutator estimates) Let \((p, q, r) \in [1, \infty ]^3\), \(\theta \) be a \(C^1\) function on \({\mathbb {R}}^{d}\) such that \((1+|\cdot |) \hat{\theta } \in L^1\). There exists a constant C such that for any Lipschitz function a with gradient in \(L^p\) and any function b in \(L^q\), we have, for any positive \(\lambda \),

$$\begin{aligned} \Vert [\theta (\lambda ^{-1} D), a]b\Vert _{L^r} \le C \lambda ^{-1}\Vert \nabla a\Vert _{L^p}\Vert b\Vert _{L^q} \quad \text{ with } \quad \frac{1}{p}+ \frac{1}{q}= \frac{1}{r}. \end{aligned}$$
(6.5)

From this, we may get the following important corollary (see Lemma 2.100 in [2] up to a slight modification).

Proposition 6.2

(Lemma 2.100 in [2]) Let \(\sigma \in {\mathbb {R}}\), \(1 \le r\le \infty \), and \(1 \le p \le p_1\le \infty \), \(1 \le p_2 \le \infty \). Let v be a vector field over \({\mathbb {R}}^{d}\). Assume that

$$\begin{aligned} \sigma >-d\min \Bigl \{\frac{1}{p_1},\frac{1}{p'}\Bigr \} \quad \text{ or }\quad \sigma >-1-d\min \Bigl \{\frac{1}{p_1},\frac{1}{p'}\Bigr \}\quad if\quad \text{ div }\, v=0. \end{aligned}$$
(6.6)

For \(j\in {\mathbb {Z}},\) denote \(R_j:=[v\cdot \nabla , \dot{\Delta }_j]f\) (or \(R_j(u,v):=\text{ div }\,[v,\dot{\Delta }_j]f\) if \(\text{ div }\, v=0\)). There exists a constant C, depending continuously on p, \(p_1\), \(\sigma \), and d, such that

$$\begin{aligned} \bigg \Vert \bigg (2^{j\sigma }\Vert R_j\Vert _{L^{p}}\bigg )_j\bigg \Vert _{\ell ^r}\lesssim {\left\{ \begin{array}{ll} \Vert \nabla v\Vert _{\dot{B}^{\frac{d}{p_1}}_{p_1,\infty }\cap L^{\infty }}\Vert f\Vert _{\dot{B}^{\sigma }_{p,r}} \quad \text{ if }\quad \sigma < 1+\frac{d}{p_1},\\ \Vert \nabla v\Vert _{\dot{B}^{\frac{d}{p_1}}_{p_1,1}}\Vert f\Vert _{\dot{B}^{\sigma }_{p,r}} \quad {if}\quad \sigma = 1+\frac{d}{p_1} \quad {and}\quad r=1. \end{array}\right. } \end{aligned}$$
(6.7)

Further, if \(f=v\), \(\sigma >0\) (or \(\sigma >-1\) if \(\text{ div }\, v=0\)), then

$$\begin{aligned} \bigg \Vert \bigg (2^{j\sigma }\Vert R_j\Vert _{L^{p}}\bigg )_j\bigg \Vert _{\ell ^r}\lesssim \Vert \nabla v\Vert _{ L^{\infty }}\Vert f\Vert _{\dot{B}^{\sigma }_{p,r}}. \end{aligned}$$
(6.8)

In the limit case \(\sigma =-d\min \Bigl \{\frac{1}{p_1},\frac{1}{p'}\Bigr \}\) or \(\sigma =-1-d\min \Bigl \{\frac{1}{p_1},\frac{1}{p'}\Bigr \}\) if \(\text{ div }\, v=0\), we have The following limit cases also hold true: \( \sup _{j \ge -1} 2^{j\sigma }\Vert R_j\Vert _{L^{p}} \lesssim \Vert \nabla v\Vert _{\dot{B}^{\frac{d}{p_1}}_{p_1,1}}\Vert f\Vert _{\dot{B}^{\sigma }_{p,\infty }}. \)

We now consider the transport-diffusion equation

$$\begin{aligned} (TD_{\nu }) {\left\{ \begin{array}{ll} \partial _t f+v\cdot \nabla f+\nu \Lambda ^{\gamma } f=g, \quad (t, x) \in {\mathbb {R}}^{+}\times {\mathbb {R}}^{d},\\ f|_{t=0}=f_0, \end{array}\right. } \end{aligned}$$
(6.9)

where \(\nu \ge 0\), \(\gamma \in (0, 2)\), \(f_0\), g, and v stand for given initial data, external force, and vector field, respectively. We aim to state a priori estimates which apply for all possible values of \(f_0\) and Lipschitz vector fields v.

From Proposition 6.2, we may get the following lemma, which proof is similar to the one of Theorem 3.14 in [2] and we omit the details.

Lemma 6.3

Let \(2\le p_1 \le \infty \) and \(r \in \{1, \, 2\} \). Let \(\sigma \in {\mathbb {R}}\) satisfy

$$\begin{aligned} s > -d \min \left\{ \frac{1}{p_1}, \frac{1}{2}\right\} \quad \text{ or } \quad s > -1-d \min \left\{ \frac{1}{p_1}, \frac{1}{2}\right\} \quad \text{ if } \quad \text{ div }\, v=0. \end{aligned}$$
(6.10)

There exists a constant \(C>0\) depending only on \(d, \, s,\, r,\,\) and \(s- 1- \frac{d}{p_1}\), such that for any smooth solution f of \((TD_{\nu })\) with \(\nu > 0\), and \(\rho \in [\rho _1, \infty ]\), we have the following a priori estimates:

$$\begin{aligned}&\begin{aligned} \Vert f\Vert _{\widetilde{L}^{\infty }_t(\dot{B}^{s}_{2, r})}\le \Vert u_0\Vert _{\dot{B}^{s}_{2, r}}+C\Vert g\Vert _{\widetilde{L}^{1}_t(\dot{B}^{s}_{2, r})} + C\int _0^t V'_{p_1}(\tau )\Vert f\Vert _{\dot{B}^{s}_{2, r}}\,d\tau \quad \text{ and } \quad \end{aligned}\\&\begin{aligned} \nu \, \Vert f\Vert _{\widetilde{L}^{1}_t(\dot{B}^{s+\gamma }_{2, r})}\lesssim&\bigg \Vert 2^{qs}(1-e^{-c_1 \nu 2^{q\gamma }t} ) \Vert \dot{\Delta }_q f_0\Vert _{L^2}\bigg \Vert _{\ell ^{r}({\mathbb {Z}})}+ \Vert g\Vert _{\widetilde{L}^{1}_t(\dot{B}^{s}_{2, r})} + \int _0^t V'_{p_1}(\tau )\Vert f\Vert _{\dot{B}^{s}_{2, r}}\,d\tau , \end{aligned} \end{aligned}$$

with, if the inequality is strict in (6.10),

$$\begin{aligned} V_{p_1}(t) \mathop {=}\limits ^{\mathrm{def}}{\left\{ \begin{array}{ll} \int _0^{t} \Vert \nabla v(\tau )\Vert _{\dot{B}^{\frac{d}{p_1}}_{p_1, \infty }\cap L^{\infty }}\, d\tau \quad &{}\text{ if } \quad s < 1+ \frac{d}{p_1},\\ \int _0^{t} \Vert \nabla v(\tau )\Vert _{\dot{B}^{\frac{d}{p_1}}_{p_1, 1}}\, d\tau \quad &{}\text{ if } \quad s =1+ \frac{d}{p_1}\,\, \text{ and } \quad r=1,\\ \int _0^{t} \Vert \nabla v(\tau )\Vert _{L^{\infty }}\, d\tau \quad &{}\text{ if } \quad f=v \, \text{ and } \, s>0 (\text{ or } \,\, s>-1\,\, \text{ if } \quad \text{ div }\, v=0). \end{array}\right. } \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gui, G., Liu, Y. Global well-posedness and blow-up of solutions for the Camassa–Holm equations with fractional dissipation. Math. Z. 281, 993–1020 (2015). https://doi.org/10.1007/s00209-015-1517-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00209-015-1517-5

Keywords

Mathematics Subject Classification

Navigation