Skip to main content
Log in

Smooth Controllability of the Navier–Stokes Equation with Navier Conditions: Application to Lagrangian Controllability

  • Published:
Archive for Rational Mechanics and Analysis Aims and scope Submit manuscript

Abstract

We deal with the 3D Navier–Stokes equation in a smooth simply connected bounded domain, with controls on a non-empty open part of the boundary and a Navier slip-with-friction boundary condition on the remaining, uncontrolled, part of the boundary. We extend the small-time global exact null controllability result in Coron et al. (J Eur Math Soc 22:1625–1673, 2020) from Leray weak solutions to the case of smooth solutions. Our strategy relies on a refinement of the method of well-prepared dissipation of the viscous boundary layers which appear near the uncontrolled part of the boundary, which allows to handle the multi-scale features in a finer topology. As a byproduct of our analysis we also obtain a small-time global approximate Lagrangian controllability result, extending to the case of the Navier–Stokes equations the recent results (Glass and Horsin in J Math Pures Appl (9) 93:61–90, 2010; Glass and Horsin in SIAM J Control Optim 50: 2726–2742, 2012; Horsin and Kavian in ESAIM Control Optim Calc Var 23:1179–1200, 2017) in the case of the Euler equations and the result (Glass and Horsin in ESAIM Control Optim Calc Var 22:1040–1053, 2016) in the case of the steady Stokes equations.

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

References

  1. Beirão de Veiga, H., Crispo, F.: Sharp inviscid limit results under Navier type boundary conditions. An \(L^p\) theory. J. Math. Fluid Mech. 12, 397–411, 2010

    Article  MathSciNet  Google Scholar 

  2. Buckmaster, T., Colombo, M., Vicol, V.: Wild solutions of the Navier-Stokes equations whose singular sets in time have Hausdorff dimension strictly less than 1. arXiv preprint arXiv:1809.00600 (2018)

  3. Buckmaster, T., Vicol, V.: Convex integration and phenomenologies in turbulence. EMS Surv. Math. Sci. 6(1), 173–263, 2020

    Article  MathSciNet  Google Scholar 

  4. Buckmaster, T., Vicol, V.: Nonuniqueness of weak solutions to the Navier–Stokes equation. Ann. Math. 189, 101–144, 2019

    Article  MathSciNet  Google Scholar 

  5. Coron, J.M., Marbach, F., Sueur, F.: Small-time global exact controllability of the Navier–Stokes equation with Navier slip-with-friction boundary conditions. J. Eur. Math. Soc. 22, 1625–1673, 2020. https://doi.org/10.4171/JEMS/952.

    Article  MathSciNet  MATH  Google Scholar 

  6. Coron, J.M., Marbach, F., Sueur, F.: On the controllability of the Navier–Stokes equation in spite of boundary layers. RIMS Kâyûroku 2058, 162–180, 2017

    Google Scholar 

  7. Coron, J.M., Marbach, F., Sueur, F., Zhang, P.: Controllability of the Navier–Stokes equation in a rectangle with a little help of a distributed phantom force. Ann. PDE 5(2), 49, 2019

    Article  MathSciNet  Google Scholar 

  8. Coron, J.M.: Contrôlabilité exacte frontière de l’équation d’Euler des fluides parfaits incompressibles bidimensionnels. C. R. Acad. Sci. Paris Sér. I Math. 317, 271–276, 1993

    MathSciNet  MATH  Google Scholar 

  9. Coron, J.M.: On the controllability of \(2\)-D incompressible perfect fluids. J. Math. Pures Appl. (9) 75, 155–188, 1996

    MathSciNet  MATH  Google Scholar 

  10. Dautray, R., Lions, J.L.: Mathematical Analysis and Numerical Methods for Sciences and Technology, vol. 3. Springer-Verlag, Berlin (1990)

    MATH  Google Scholar 

  11. Gagnon, L.: Lagrangian controllability of the 1-D Korteweg-de Vries equation. SIAM J. Control. Optim. 54, 3152–3173, 2016

    Article  MathSciNet  Google Scholar 

  12. Gie, G.M., Kelliher, J.: Boundary layer analysis of the Navier–Stokes equations with generalized Navier boundary conditions. J. Differ. Equ. 253, 1862–1892, 2012

    Article  ADS  MathSciNet  Google Scholar 

  13. Glass, O.: Exact boundary controllability of 3-D Euler equation. ESAIM Control Optim. Calc. Var. 5, 1–44, 2000

    Article  MathSciNet  Google Scholar 

  14. Glass, O.: Some questions of control in fluid mechanics. In: Control of Partial Differential Equations, pp. 131–206. Springer, Berlin (2012)

  15. Glass, O.: Contrôlabilité exacte frontière de l’équation d’Euler des fluides parfaits incompressibles en dimension 3. C. R. Acad. Sci. Paris Sér. I Math. 325, 987–992, 1997

    Article  ADS  MathSciNet  Google Scholar 

  16. Glass, O.: An addendum to a J. M. Coron theorem concerning the controllability of the Euler system for 2D incompressible inviscid fluids. “On the controllability of 2-D incompressible perfect fluids” J. Math. Pures Appl. (9) 75(2) (1996), 155–188; MR1380673 (97b:93010)

  17. Glass, O.: An addendum to a J. M. Coron theorem concerning the controllability of the Euler system for 2D incompressible inviscid fluids. “On the controllability of 2-D incompressible perfect fluids” J. Math. Pures Appl. (9), 80 (2001).

  18. Glass, O., Horsin, T.: Approximate Lagrangian controllability for the 2-D Euler equation. Application to the control of the shape of vortex patches. J. Math. Pures Appl. (9) 93, 61–90, 2010

    Article  MathSciNet  Google Scholar 

  19. Glass, O., Horsin, T.: Prescribing the motion of a set of particles in a three-dimensional perfect fluid. SIAM J. Control. Optim. 50, 2726–2742, 2012

    Article  MathSciNet  Google Scholar 

  20. Glass, O., Horsin, T.: Lagrangian controllability at low Reynolds number. ESAIM Control Optim. Calc. Var. 22, 1040–1053, 2016

    Article  MathSciNet  Google Scholar 

  21. Guerrero, S.: Local exact controllability to the trajectories of the Navier–Stokes system with nonlinear Navier-slip boundary conditions. ESAIM Control Optim. Calc. Var. 12, 484–544, 2006

    Article  MathSciNet  Google Scholar 

  22. Guès, O.: Problème mixte hyperbolique quasi-linéaire caractèristique. Commun. Partial Differ. Equ. 15, 595–645, 1990

    Article  Google Scholar 

  23. Hopf, E.: Über die Anfangswertaufgabe für die hydrodynamischen Grundgleichungen, (German) Math. Nachr. 4, 213–231, 1951

    Article  MathSciNet  Google Scholar 

  24. Horsin, T., Kavian, O.: Lagrangian controllability of inviscid incompressible fluids: a constructive approach. ESAIM Control Optim. Calc. Var. 23, 1179–1200, 2017

    Article  MathSciNet  Google Scholar 

  25. Iftimie, D., Sueur, F.: Viscous boundary layers for the Navier–Stokes equations with the Navier slip conditions. Arch. Ration. Mech. Anal. 199, 145–175, 2011

    Article  MathSciNet  Google Scholar 

  26. Koch, H.: Transport and instability for perfect fluids. Math. Ann. 323(3), 491–523, 2002

    Article  MathSciNet  Google Scholar 

  27. Krygin, A.B.: Extension of diffeomorphisms that preserve volume. Funktsional. Anal. i Prilozhen. 5, 72–76, 1971

    Article  MathSciNet  Google Scholar 

  28. Leray, J.: Sur le mouvement d’un liquide visqueux emplissant l’espace. Acta Math. 63, 193–248, 1934

    Article  MathSciNet  Google Scholar 

  29. Marbach, F.: Small time global null controllability for a viscous Burgers’ equation despite the presence of a boundary layer. J. Math. Pures Appl. (9) 102, 364–384, 2014

    Article  MathSciNet  Google Scholar 

  30. Masmoudi, N., Rousset, F.: Uniform regularity for the Navier–Stokes equation with Navier boundary condition. Arch. Ration. Mech. Anal. 203, 529–575, 2012

    Article  MathSciNet  Google Scholar 

  31. Navier, C.-L.: Mémoire sur les lois du mouvement des fluides. Mémoires de l’Académie Royale des Sciences de l’Institut de France 6, 389–440, 1823

    Google Scholar 

  32. Seregin, G.: Lecture notes on Regularity Theory for the Navier–Stokes Equations. World Scientific Publishing Co. Pte. Ltd., Hackensack (2015)

  33. Sueur, F.: Couches limites semilinéaires. Ann. Fac. Sci. Toulouse Math. (6) 15, 323–380, 2006

    Article  MathSciNet  Google Scholar 

  34. Sueur, F.: Approche visqueuse de solutions discontinues de systèmes hyperboliques semilinéaires. Ann. Inst. Fourier (Grenoble) 56, 183–245, 2006

    Article  MathSciNet  Google Scholar 

  35. Sueur, F.: Viscous profile of vortex patches. J. Inst. Math. Jussieu 14, 1–68, 2013

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

All the authors are supported by K. C. Wong Education Foundation. F. Sueur is partially supported by the Agence Nationale de la Recherche, Project IFSMACS, grant ANR-15-CE40-0010, Project SINGFLOWS, grant ANR-18-CE40-0027-01, and Project BORDS, grant ANR-16-CE40-0027-01; and by the H2020-MSCA-ITN-2017 program, Project ConFlex, Grant ETN-765579. P. Zhang is partially supported by NSF of China under Grants 11731007, 12031006 and 11688101. F. Sueur warmly thanks Morningside center of Mathematics, CAS, for its kind hospitality during his stays in May 2018 and October 2019.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Franck Sueur.

Additional information

Communicated by J. Bedrossian.

Publisher's Note

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

Appendix A. On the Regularization of the Uncontrolled Strong Solutions to the Navier–Stokes Equations with Navier Boundary Conditions

Appendix A. On the Regularization of the Uncontrolled Strong Solutions to the Navier–Stokes Equations with Navier Boundary Conditions

In this appendix we prove a regularization result of the uncontrolled strong solutions to the Navier–Stokes equations with Navier boundary conditions on the whole boundary \(\partial \Omega \), that is, to the following system:

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _t u+u\cdot \nabla u- \Delta u +\nabla p=0, \quad \text{ and } \quad \mathrm {div} \,u=0 \quad \text{ in } \Omega ,\\ u\cdot {\mathbf {n}}=0 \quad \text{ and } \quad {\mathcal {N}}(u)=0\quad \text{ on } \partial \Omega , \\ u=u_0\quad \text{ at } t=0 . \end{array}\right. } \end{aligned}$$
(A.1)

Theorem A.1

Let \(T>0\), p in \({\mathbb {N}}^*\) and \(R>0\). Then there exists a continuous function \(C_{T,p,R}\) from \( [0,+\infty ) \) to \([0,+\infty )\) with \(C_{T,p,R}(0)=0\), such that there exists \(T_1\) in (0, T) and for any \(u_0\) in \(H^1(\Omega )\), with \( \Vert u_0\Vert _{H^1(\Omega )} \le R\), divergence free and tangent to \(\partial \Omega \), the unique strong solution u in \( C([0,T_1];H^1(\Omega ))\cap L^2([0,T_1];H^2(\Omega ))\) to (A.1) satisfies

$$\begin{aligned}&\sum _{0\le j \le \frac{p}{2}}\bigl \Vert t^{\frac{p-1}{2}}\partial _t^ju\bigr \Vert _{L^\infty _{T_1}(H^{p-2j}(\Omega ))} +\sum _{0\le j \le \frac{p+1}{2}}\bigl \Vert t^{\frac{p-1}{2}}\partial _t^ju\bigr \Vert _{L^2_{T_1}(H^{p+1-2j}(\Omega ))}\nonumber \\&\quad \le C_{p,T_1,R}(\Vert u_0\Vert _{H^1(\Omega )}). \end{aligned}$$
(A.2)

As recalled in Sect. 2.1 The goal of this section is to present the proof of Theorem 2.1. The local-in-time existence and uniqueness of strong solutions with \(H^1\) initial data is classical. The interest of Theorem A.1 is to detail the regularization in time of this strong solution near the time zero. In particular it implies the part of Theorem 2.1 regarding the regularization.

Proof

We will proceed by induction on p. We start with recalling how to prove the case \(p=1\), by proving first a \(L^2(\Omega )\) energy estimate and then a \(H^1(\Omega )\) energy estimate.

\(\bullet \) \({\underline{L^{2}(\Omega )\,\, \hbox {energy}\,\, \hbox {estimate}}}\)

Indeed, we first get, by taking \(L^2(\Omega )\) inner product of the u equation in (A.1) with u,  that

$$\begin{aligned} \frac{1}{2}\frac{d}{dt}\Vert u(t)\Vert _{L^2(\Omega )}^2+\left( u\cdot \nabla u | u\right) _{L^2(\Omega )}-\left( \Delta u | u\right) _{L^2(\Omega )}+\left( \nabla p | u\right) _{L^2(\Omega )}=0. \end{aligned}$$
(A.3)

Here and in all that follows, we always denote \((f | g)_{L^2(\Omega )}:=\int _{\Omega } f g\,dx.\)

Due to \(\mathrm {div} \,u=0\) and \(u\cdot \mathbf{n}|_{\partial \Omega }=0,\) we have

$$\begin{aligned} \left( u\cdot \nabla u | u\right) _{L^2(\Omega )}=0=\left( \nabla p | u\right) _{L^2(\Omega )}. \end{aligned}$$

Moreover it follows from Stokes formula that

$$\begin{aligned} \begin{aligned} -\left( \Delta u | u\right) _{L^2(\Omega )} =&\int _{\partial \Omega }\left[ (\nabla \times u)\times u\right] \cdot \mathbf{n}\,dS+\int _{\Omega }|\nabla \times u|^2\,dx. \end{aligned} \end{aligned}$$

By inserting the above equalities into (A.3), we obtain

$$\begin{aligned} \frac{1}{2}\frac{d}{dt}\Vert u(t)\Vert _{L^2(\Omega )}^2+\Vert \nabla \times u\Vert _{L^2(\Omega )}^2=\int _{\partial \Omega }\left[ u\times (\nabla \times u)\right] \cdot \mathbf{n}\,dS. \end{aligned}$$
(A.4)

Let us denote by \(M_{\mathrm{w}}\) the shape operator associated with \(\Omega \). Recall that, since \(\Omega \) is smooth, the shape operator \(M_{\mathrm{w}}\) is smooth and for any x in \(\partial \Omega ,\) it defines a self-adjoint operator with values in the tangent space \(T_x\). Then we have the following result, see [1, 12].

Lemma A.2

For any smooth divergence free vector field u satisfying \(u\cdot \mathbf{n}=0\) on \(\partial \Omega ,\) we have

$$\begin{aligned} \left[ D(u)\mathbf{n}+M_{\mathrm{w}}u\right] _{\text{ tan }}=\frac{1}{2}(\nabla \times u)\times \mathbf{n}. \end{aligned}$$
(A.5)

However, due to \({\mathcal {N}}(u)|_{\partial \Omega }=0,\) we deduce from Lemma A.2 that

$$\begin{aligned} \begin{aligned} \left[ u \times (\nabla \times u)\right] \cdot \mathbf{n}\bigr |_{\partial \Omega }=&u\cdot \left[ (\nabla \times u)\times \mathbf{n}\right] \bigr |_{\partial \Omega }\\ =&2\left[ (M_{\mathrm{w}}-M)u\right] _{\text{ tan }}\cdot u\bigr |_{\partial \Omega }\\ =&2\left[ (M_{\mathrm{w}}-M)u\right] \cdot u\bigr |_{\partial \Omega }, \end{aligned} \end{aligned}$$
(A.6)

where we used \(u\cdot \mathbf{n}\bigr |_{\partial \Omega }=0\) in the last step. Then by applying Stokes formula and Young’s inequality, we find that for any \(\lambda >0,\) there exists \(C_\lambda \) so that

$$\begin{aligned} \begin{aligned} \bigl |\int _{\partial \Omega }\left[ (\nabla \times u)\times u\right] \cdot \mathbf{n}\,dS\bigr |=&2\bigl |\int _{\Omega }\mathrm {div} \,\left[ \bigl ((M_{\mathrm{w}}-M)u\cdot u\bigr )\mathbf{n}\right] \,dx\bigr |\\ \le&\lambda \Vert \nabla u\Vert _{L^2(\Omega )}^2+C_\lambda \Vert u\Vert _{L^2(\Omega )}^2, \end{aligned} \end{aligned}$$
(A.7)

On the other hand, due to \(\mathrm {div} \,u=0\) in \(\Omega \) and \(u\cdot \mathbf{n}|_{\partial \Omega }=0,\) we deduce from Korn’s type inequality (see [10] for instance) that there exists a positive constant \(C_\Omega \) so that

$$\begin{aligned} \Vert \nabla \times u\Vert _{L^2(\Omega )}^2\ge \frac{1}{C_\Omega } \Vert u\Vert _{H^1(\Omega )}^2-\Vert u\Vert _{L^2(\Omega )}^2. \end{aligned}$$
(A.8)

By inserting the estimates, (A.7) and (A.8), into (A.4) and taking \(\lambda =\frac{1}{2C_\Omega }\) in the resulting inequality, we achieve

$$\begin{aligned} \frac{d}{dt}\Vert u(t)\Vert _{L^2(\Omega )}^2+\frac{1}{C_\Omega } \Vert u\Vert _{H^1(\Omega )}^2\le C \Vert u\Vert _{L^2(\Omega )}^2. \end{aligned}$$
(A.9)

Applying Gronwall’s inequality gives rise to

$$\begin{aligned} \Vert u\Vert _{L^\infty _t(L^2(\Omega ))}^2+\frac{1}{C_\Omega } \Vert u\Vert _{L^2_t(H^1(\Omega ))}^2 \le \Vert u_0\Vert _{L^2(\Omega )}^2 e^{Ct}. \end{aligned}$$
(A.10)

\(\bullet \) \({\underline{H^{1}(\Omega )\,\, \hbox {energy}\,\, \hbox {estimate}}}\)

By taking \(L^2(\Omega )\) inner product of the u equation of (A.1) with \(\partial _tu,\) we get

$$\begin{aligned} \Vert \partial _tu\Vert _{L^2(\Omega )}^2-\left( \Delta u | \partial _t u\right) _{L^2(\Omega )}+\left( \nabla p | \partial _t u\right) _{L^2(\Omega )}=-\left( u\cdot \nabla u | \partial _t u\right) _{L^2(\Omega )}. \end{aligned}$$
(A.11)

Notice that \(\partial _tu\cdot \mathbf{n}|_{\partial \Omega }=0,\) by applying Stokes formula and along the same line to the proof of (A.6), we obtain

$$\begin{aligned} \begin{aligned} -\left( \Delta u | \partial _t u\right) _{L^2(\Omega )}=&\int _{\partial \Omega }[(\nabla \times u)\times \partial _tu]\cdot \mathbf{n}\,dS+\int _{\Omega }(\nabla \times u) \cdot (\nabla \times \partial _tu)\,dx\\ =&2\int _{\partial \Omega } \partial _tu(M-M_\mathrm{w})u\,dS+\frac{1}{2}\frac{d}{dt}\int _{\Omega }|\nabla \times u|^2\,dx, \end{aligned} \end{aligned}$$

which together with the facts: M is a symmetric matrix and \(M_{\mathrm{w}}\) is a self-adjoint operator on \(T_x,\) ensures that

$$\begin{aligned} -\left( \Delta u | \partial _t u\right) _{L^2(\Omega )}=\frac{d}{dt}\Bigl (\int _{\partial \Omega } u(M-M_{\mathrm{w}})u\,dS+\frac{1}{2}\int _{\Omega }|\nabla \times u|^2\,dx\Bigr ). \end{aligned}$$

Again due to \(\partial _tu\cdot \mathbf{n}|_{\partial \Omega }=0,\) one has

$$\begin{aligned} \left( \nabla p | \partial _t u\right) _{L^2(\Omega )}=0. \end{aligned}$$

By inserting the above equalities into (A.11), we achieve

$$\begin{aligned} \begin{aligned}&\frac{d}{dt}\Bigl (\int _{\partial \Omega } u(M-M_\mathrm{w})u\,dS+\frac{1}{2}\int _{\Omega }|\nabla \times u|^2\,dx\Bigr )+\Vert \partial _tu\Vert _{L^2(\Omega )}^2 =-\left( u\cdot \nabla u | \partial _t u\right) _{L^2(\Omega )}\\&\quad \le \Vert u\Vert _{L^6(\Omega )}\Vert \nabla u\Vert _{L^3(\Omega )}\Vert \partial _tu\Vert _{L^2(\Omega )}\\&\quad \le C\Vert u\Vert _{H^1(\Omega )}\Vert \nabla u\Vert _{L^2(\Omega )}^{\frac{1}{2}}\Vert \nabla u\Vert _{H^1(\Omega )}^{\frac{1}{2}}\Vert \partial _tu\Vert _{L^2(\Omega )}. \end{aligned} \end{aligned}$$

Applying Young’s inequality yields

$$\begin{aligned} \begin{aligned}&\frac{d}{dt}\Bigl (\int _{\partial \Omega } u(M-M_{\mathrm{w}})u\,dS+\frac{1}{2}\int _{\Omega }|\nabla \times u|^2\,dx\Bigr )+\frac{3}{4}\Vert \partial _tu\Vert _{L^2(\Omega )}^2\\&\quad \le C_\lambda \bigl (1+\Vert u\Vert _{H^1(\Omega )}^4\bigr )\Vert \nabla u\Vert _{L^2(\Omega )}^2+\lambda \Vert \nabla ^2 u\Vert _{L^2(\Omega )}^2. \end{aligned} \end{aligned}$$
(A.12)

Moreover in view of (A.1), we write

$$\begin{aligned} {\left\{ \begin{array}{ll} - \Delta u +\nabla p=-\partial _t u-u\cdot \nabla u \\ \mathrm {div} \,u=0 \quad \text{ in } \Omega ,\\ u\cdot {\mathbf {n}}=0 \quad \text{ and } \quad {\mathcal {N}}(u)=0\quad \text{ on } \partial \Omega . \end{array}\right. } \end{aligned}$$
(A.13)

The following type of Cattabriga-Solonnikov estimate can be proved along the same line to that of Theorem 2.2 in [32]:

Lemma A.3

Let k be a non-negative integer and \(\Omega \) be a bounded domain with sufficiently smooth boundary. Let f in \( H^k(\Omega )\) and g in \( H^{k+1}(\Omega )\) with \(\int _\Omega g\,dx=0.\) Then the non-homogeneous Stokes problem

$$\begin{aligned} {\left\{ \begin{array}{ll} - \Delta u +\nabla p=f \\ \mathrm {div} \,u=g \quad \text{ in } \Omega ,\\ u\cdot {\mathbf {n}}=0 \quad \text{ and } \quad {\mathcal {N}}(u)=0\quad \text{ on } \partial \Omega \end{array}\right. } \end{aligned}$$

has a unique solution (up) so that

$$\begin{aligned} \Vert \nabla ^2u\Vert _{H^k(\Omega )}+\Vert \nabla p\Vert _{H^k(\Omega )}\le C\bigl (\Vert f\Vert _{H^k(\Omega )}+\Vert \nabla g\Vert _{H^k(\Omega )}\bigr ). \end{aligned}$$
(A.14)

Then it follows from Lemma A.3 and (A.13) that

$$\begin{aligned} \begin{aligned} \Vert \nabla ^2 u\Vert _{L^2(\Omega )}\le&C\bigl (\Vert \partial _tu\Vert _{L^2(\Omega )}+\Vert u\cdot \nabla u\Vert _{L^2(\Omega )}\bigr )\\ \le&C\bigl (\Vert \partial _tu\Vert _{L^2(\Omega )}+\Vert u\Vert _{H^1(\Omega )}\Vert \nabla u\Vert _{L^2(\Omega )}^{\frac{1}{2}}\Vert \nabla u\Vert _{H^1(\Omega )}^{\frac{1}{2}}\bigr ), \end{aligned} \end{aligned}$$

from which, we infer

$$\begin{aligned} \Vert \nabla u\Vert _{H^1(\Omega )}\le C\bigl (\Vert \partial _tu\Vert _{L^2(\Omega )}+(1+\Vert u\Vert _{H^1(\Omega )}^2)\Vert \nabla u\Vert _{L^2(\Omega )}\bigr ). \end{aligned}$$
(A.15)

By substituting (A.15) into (A.12) and then taking \(\lambda =\frac{1}{4C},\) we achieve

$$\begin{aligned} \begin{aligned} \frac{d}{dt}\Bigl (\int _{\partial \Omega } u(M-M_{\mathrm{w}})u\,dS+\frac{1}{2}\int _{\Omega }|\nabla \times u|^2\,dx\Bigr )+&\frac{1}{2}\Vert \partial _tu\Vert _{L^2(\Omega )}^2\\ \le&C\bigl (1+\Vert u\Vert _{H^1(\Omega )}^4\bigr )\Vert \nabla u\Vert _{L^2(\Omega )}^2. \end{aligned} \end{aligned}$$
(A.16)

On the other hand, it follows from trace inequality (5.25) that

$$\begin{aligned} \begin{aligned} \bigl |\int _{\partial \Omega } u(M-M_{\mathrm{w}})u\,dS\bigr |\le C\Vert u\Vert _{L^2(\partial \Omega )}^2\le&C\bigl (\Vert u\Vert _{L^2(\Omega )}^2+\Vert u\Vert _{L^2(\Omega )}\Vert \nabla u\Vert _{L^2(\Omega )}\bigr )\\ \le&\frac{1}{4C_\Omega }\Vert u\Vert _{H^1(\Omega )}^2+C\Vert u\Vert _{L^2(\Omega )}^{2}, \end{aligned} \end{aligned}$$

so that in view of (A.8), there exists a large enough constant K which satisfies

$$\begin{aligned} E_1(u):= K\Vert u\Vert _{L^2(\Omega )}^{2}+\int _{\partial \Omega } u(M-M_\mathrm{w})u\,dS+\frac{1}{2}\int _{\Omega }|\nabla \times u|^2\,dx\ge \frac{1}{4C_\Omega }\Vert u\Vert _{H^1(\Omega )}^2.\nonumber \\ \end{aligned}$$
(A.17)

Then we get, by summing up \(K\times \)(A.9) and (A.16), that

$$\begin{aligned} \frac{d}{dt}E_1(u)+\frac{1}{2}\Vert \partial _tu\Vert _{L^2(\Omega )}^2\le CE_1(u)\bigl (1+E_1^2(u)\bigr ), \end{aligned}$$
(A.18)

from which, we deduce by a comparison argument that for any \(T>0\) and \(R>0\), there exists a continuous function \(C_{T,p,R}\) from \( [0,+\infty ) \) to \([0,+\infty )\) with \(C_{T,1,R}(0)=0\), such that there exists \(T_1\) in (0, T) and such that for any \(u_0\) in \(H^1(\Omega )\), with \( \Vert u_0\Vert _{H^1(\Omega )} \le R\), divergence free and tangent to \(\partial \Omega \), the unique strong solution u in \( C([0,T_1];H^1(\Omega ))\cap L^2([0,T_1];H^2(\Omega ))\) to (A.1) satisfies (A.2) holds true for \(p=1.\)

\(\bullet \) \(\underline{\hbox {Higher}\,\, \hbox {energy}\,\, \hbox {estimates}}\)

Inductively, we assume that (A.2) holds for \(p\le \ell -1,\) we are going to show that (A.2) holds for \(p=\ell .\) Without loss of generality, we may assume that \(\ell \) is an even integer. The odd integer case can be proved along the same line. Indeed we first get, by applying \(\partial _t^{\ell /2}\) to (A.1), that

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _t^{1+\frac{\ell }{2}} u+\partial _t^{\frac{\ell }{2}}(u\cdot \nabla u)- \Delta \partial _t^{\frac{\ell }{2}}u +\nabla \partial _t^{\frac{\ell }{2}}p=0, \\ \mathrm {div} \,\partial _t^{\frac{\ell }{2}}u=0 \quad \text{ in } (0,T_1)\times \Omega ,\\ \partial _t^{\frac{\ell }{2}}u\cdot {\mathbf {n}}=0 \quad \text{ and } \quad {\mathcal {N}}(\partial _t^{\frac{\ell }{2}}u)=0\quad \text{ on } \ (0,T_1)\times \partial \Omega , \end{array}\right. } \end{aligned}$$
(A.19)

from which, we get, by a similar derivation of (A.4) that

$$\begin{aligned} \begin{aligned}&\frac{1}{2}\frac{d}{dt}\bigl (t^{\ell -1}\Vert \partial _t^{\frac{\ell }{2}}u(t)\Vert _{L^2(\Omega )}^2\bigr )+t^{\ell -1}\Vert \nabla \times \partial _t^{\frac{\ell }{2}}u\Vert _{L^2(\Omega )}^2 =\frac{\ell -1}{2}t^{\ell -2}\Vert \partial _t^{\frac{\ell }{2}}u\Vert _{L^2(\Omega )}^2\\&\quad +t^{\ell -1}\int _{\partial \Omega }\bigl [\partial _t^{\frac{\ell }{2}}u\times (\nabla \times \partial _t^{\frac{\ell }{2}}u)\bigr ]\cdot \mathbf{n}\,dS -t^{\ell -1}\bigl (\partial _t^{\frac{\ell }{2}}(u\cdot \nabla u) | \partial _t^{\frac{\ell }{2}}u\bigr )_{L^2(\Omega )}. \end{aligned} \end{aligned}$$
(A.20)

Similarly to (A.7), we have

$$\begin{aligned} t^{\ell -1}\bigl |\int _{\partial \Omega }\bigl [\partial _t^{\frac{\ell }{2}}u\times (\nabla \times \partial _t^{\frac{\ell }{2}}u)\bigr ]\cdot \mathbf{n}\,dS\bigr | \le \lambda \bigl \Vert t^{\frac{\ell -1}{2}}\nabla \partial ^{\frac{\ell }{2}}_tu\bigr \Vert _{L^2(\Omega )}^2+C_\lambda \bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{\frac{\ell }{2}}u\bigr \Vert _{L^2(\Omega )}^2. \end{aligned}$$

On the other hand, due to \(u\cdot \mathbf{n}|_{\partial \Omega }=0\) and \(\mathrm {div} \,u=0,\) we get, by using integration by parts, that

$$\begin{aligned} \begin{aligned} \bigl (\partial _t^{\frac{\ell }{2}}(u\cdot \nabla u) | \partial _t^{\frac{\ell }{2}}u\bigr )_{L^2(\Omega )}=&\bigl (\partial _t^{\frac{\ell }{2}}( u\cdot \nabla u)-u\cdot \nabla \partial _t^{\frac{\ell }{2}}u | \partial ^{\frac{\ell }{2}}_tu \bigr )_{L^2(\Omega )}\\ =&-\sum _{\begin{array}{c} \ell _1+\ell _2=\frac{\ell }{2}\\ \ell _1\ge 1 \end{array}}C_{\frac{\ell }{2}}^{\ell _1}\bigl (\partial _t^{\ell _1}u\otimes \partial _t^{\ell _2}u | \nabla \partial ^{\frac{\ell }{2}}_tu \bigr )_{L^2(\Omega )}, \end{aligned} \end{aligned}$$

from which we infer

$$\begin{aligned} \begin{aligned}&t^{\ell -1}\bigl |\bigl (\partial _t^{\frac{\ell }{2}}(u\cdot \nabla u) | \partial _t^{\frac{\ell }{2}}u\bigr )_{L^2(\Omega )}\bigr | \\&\quad \lesssim \sum _{\begin{array}{c} \ell _1+\ell _2=\frac{\ell }{2}\\ \ell _1\ge 1 \end{array}} t^{\ell -1}\Vert \partial _t^{\ell _1}u\Vert _{L^3(\Omega )}\Vert \partial _t^{\ell _2}u\Vert _{L^6(\Omega )}\Vert \nabla \partial ^{\frac{\ell }{2}}_tu\Vert _{L^2(\Omega )}\\&\quad \lesssim \sum _{\begin{array}{c} \ell _1+\ell _2=\frac{\ell }{2}\\ \ell _1\ge 1 \end{array}} t^{\ell -1}\Vert \partial _t^{\ell _1}u\Vert _{L^2(\Omega )}^{\frac{1}{2}}\Vert \partial _t^{\ell _1}u\Vert _{H^1(\Omega )}^{\frac{1}{2}}\Vert \partial _t^{\ell _2}u\Vert _{H^1(\Omega )}\Vert \nabla \partial ^{\frac{\ell }{2}}_tu\Vert _{L^2(\Omega )}\\&\quad \le \lambda \bigl \Vert t^{\frac{\ell -1}{2}} \partial _t^{\frac{\ell }{2}}u\bigr \Vert _{H^1(\Omega )}^2+C_\lambda \Vert u\Vert _{H^1(\Omega )}^{4}\bigl \Vert t^{\frac{\ell -1}{2}} \partial ^{\frac{\ell }{2}}_tu\bigr \Vert _{L^2(\Omega )}^2\\&\qquad +C_\lambda \sum _{\begin{array}{c} \ell _1+\ell _2=\frac{\ell }{2}\\ 1\le \ell _1\le \frac{\ell }{2}-1 \end{array}} \bigl \Vert t^{\ell _1-\frac{1}{2}}\partial _t^{\ell _1}u\bigr \Vert _{H^1(\Omega )}^2\bigl \Vert |t^{\ell _2}\partial _t^{\ell _2}u\bigr \Vert _{H^1(\Omega )}^2. \end{aligned} \end{aligned}$$

By substituting the above estimates into (A.20) and using Korn’s type inequality (A.8), we find

$$\begin{aligned} \begin{aligned}&\frac{1}{2}\frac{d}{dt}\bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{\frac{\ell }{2}}u(t)\bigr \Vert _{L^2(\Omega )}^2+\frac{1}{C_\Omega }\bigl \Vert t^{\frac{\ell -1}{2}} \partial _t^{\frac{\ell }{2}}u\bigr \Vert _{H^1(\Omega )}^2 \\&\quad \le \frac{\ell -1}{2}\bigl \Vert t^{\frac{\ell }{2}-1}\partial _t^{\frac{\ell }{2}}u\bigr \Vert _{L^2(\Omega )}^2+C_\lambda \bigl (1+ \Vert u\Vert _{H^1(\Omega )}^{4}\bigr )\bigl \Vert t^{\frac{\ell -1}{2}} \partial ^{\frac{\ell }{2}}_tu\bigr \Vert _{L^2(\Omega )}^2\\&\qquad +2\lambda \bigl \Vert t^{\frac{\ell -1}{2}} \partial _t^{\frac{\ell }{2}}u\bigr \Vert _{H^1(\Omega )}^2+ C_\lambda \sum _{\begin{array}{c} \ell _1+\ell _2=\frac{\ell }{2}\\ 1\le \ell _1\le \frac{\ell }{2}-1 \end{array}} \bigl \Vert t^{\ell _1-\frac{1}{2}}\partial _t^{\ell _1}u\bigr \Vert _{H^1(\Omega )}^2\bigl \Vert |t^{\ell _2}\partial _t^{\ell _2}u\bigr \Vert _{H^1(\Omega )}^2. \end{aligned} \end{aligned}$$

By taking \(\lambda =\frac{1}{4C_\Omega }\) in the above inequality and then applying Gronwall’s inequality to the resulting inequality, we achieve

$$\begin{aligned} \begin{aligned}&\bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{\frac{\ell }{2}}u\bigr \Vert _{L^\infty _t(L^2(\Omega ))}^2+\frac{1}{C_\Omega }\bigl \Vert t^{\frac{\ell -1}{2}} \partial _t^{\frac{\ell }{2}}u\bigr \Vert _{L^2_t(H^1(\Omega ))}^2\le C\exp \left( C\bigl (1+t\Vert u\Vert _{L^\infty _{t}(H^1(\Omega ))}^4\bigr )\right) \\&\quad \times \Bigl (\bigl \Vert t^{\frac{\ell }{2}-1}\partial _t^{\frac{\ell }{2}}u\bigr \Vert _{L^2_t(L^2(\Omega ))}^2+ \sum _{\begin{array}{c} \ell _1+\ell _2=\frac{\ell }{2}\\ 1\le \ell _1\le \frac{\ell }{2}-1 \end{array}} \bigl \Vert t^{\ell _1-\frac{1}{2}}\partial _t^{\ell _1}u\bigr \Vert _{L^2_t(H^1(\Omega ))}^2\bigl \Vert |t^{\ell _2}\partial _t^{\ell _2}u\bigr \Vert _{L^\infty _t(H^1(\Omega ))}^2\Bigr ), \end{aligned} \end{aligned}$$

from which, with the inductive assumption, we deduce that

$$\begin{aligned} \begin{aligned} \bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{\frac{\ell }{2}}u\bigr \Vert _{L^\infty _{T_1}(L^2(\Omega ))}^2+\frac{1}{C_\Omega }\bigl \Vert t^{\frac{\ell -1}{2}} \partial _t^{\frac{\ell }{2}}u\bigr \Vert _{L^2_{T_1}(H^1(\Omega ))}^2\le C_{\ell , T_1}(\Vert u_0\Vert _{H^1(\Omega )}). \end{aligned} \end{aligned}$$
(A.21)

On the other hand, for any non-negative integer \(j\le \frac{\ell }{2}-1,\) we infer from the inductive assumption that

$$\begin{aligned} \begin{aligned} \bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j}u\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2j}(\Omega ))} =&\bigl \Vert t^{\frac{\ell -1}{2}}\nabla ^2\partial _t^{j}u\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2-2j}(\Omega ))} +\bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j}u\bigr \Vert _{L^\infty _{T_1}(H^{\ell -1-2j}(\Omega ))}\\ \le&\bigl \Vert t^{\frac{\ell -1}{2}}\nabla ^2\partial _t^{j}u\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2-2j}(\Omega ))}+ C_{\ell , T_1}(\Vert u_0\Vert _{H^1(\Omega )}). \end{aligned} \end{aligned}$$

Moreover in view of (A.1), we write

$$\begin{aligned} - \Delta \partial _t^{j}u +\nabla \partial _t^{j}p=-\partial _t^{j+1} u-\partial _t^{j}(u\cdot \nabla u), \end{aligned}$$

from which, with Lemma A.3, we infer

$$\begin{aligned} \begin{aligned} \bigl \Vert t^{\frac{\ell -1}{2}}\nabla ^2\partial _t^{j}u\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2-2j}(\Omega ))}\lesssim&\bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j+1}u\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2-2j}(\Omega ))}\\&+ \bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j}(u\cdot \nabla u)\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2-2j}(\Omega ))}. \end{aligned} \end{aligned}$$

As a result, we get that

$$\begin{aligned} \begin{aligned} \bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j}u\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2j}(\Omega ))} \le&C_{\ell , T_1}(\Vert u_0\Vert _{H^1(\Omega )})+\bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j+1}u\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2-2j}(\Omega ))}\\&+ \bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j}(u\cdot \nabla u)\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2-2j}(\Omega ))}, \quad \forall \ j\le \frac{\ell }{2}-1. \end{aligned} \end{aligned}$$
(A.22)

However, it follows from Moser type inequality and the inductive assumption that

$$\begin{aligned} \begin{aligned} \bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j}\nabla (u\otimes u)\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2-2j}(\Omega ))} \lesssim&\sum _{j_1+j_2=j}\bigl \Vert t^{j_1+\frac{1}{2}}\partial _t^{j_1}u\bigr \Vert _{L^\infty _{T_1}(H^{2}(\Omega ))}\\&\times \bigl \Vert t^{\frac{\ell -2}{2}-j+j_2}\partial _t^{j_2}u\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2j-1}(\Omega ))}\\ \le&C_{\ell , T_1}(\Vert u_0\Vert _{H^1(\Omega )}). \end{aligned} \end{aligned}$$

Substituting the above estimates into (A.22) gives rise to

$$\begin{aligned} \begin{aligned} \bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j}u\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2j}(\Omega ))} \le&C_{\ell , T_1}(\Vert u_0\Vert _{H^1(\Omega )})+\bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j+1}u\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2-2j}(\Omega ))}. \end{aligned} \end{aligned}$$

We deduce from this inequality and from (A.21), by an iterative argument, that

$$\begin{aligned} \sum _{0\le j \le \frac{\ell }{2}}\bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^ju\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2j}(\Omega ))} \le C_{\ell ,T_1}(\Vert u_0\Vert _{H^1(\Omega )}). \end{aligned}$$
(A.23)

Exactly along the same line to the proof of (A.23), for any non-negative integer \(j\le \frac{\ell }{2}-1,\) we infer from the inductive assumption that

$$\begin{aligned} \begin{aligned} \bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j}u\bigr \Vert _{L^2_{T_1}(H^{\ell +1-2j}(\Omega ))} \le&\bigl \Vert t^{\frac{\ell -1}{2}}\nabla ^2\partial _t^{j}u\bigr \Vert _{L^2_{T_1}(H^{\ell -1-2j}(\Omega ))}+ C_{\ell , T_1}(\Vert u_0\Vert _{H^1(\Omega )}). \end{aligned} \end{aligned}$$

On the other hand it follows from Lemma A.3 that

$$\begin{aligned} \begin{aligned} \bigl \Vert t^{\frac{\ell -1}{2}}\nabla ^2\partial _t^{j}u\bigr \Vert _{L^2_{T_1}(H^{\ell -1-2j}(\Omega ))}\lesssim&\bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j+1}u\bigr \Vert _{L^2_{T_1}(H^{\ell -1-2j}(\Omega ))}\\&+ \bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j}(u\otimes u)\bigr \Vert _{L^\infty _{T_1}(H^{\ell -2j}(\Omega ))}. \end{aligned} \end{aligned}$$

For, any \(j\le \frac{\ell }{2}-1,\) it follows from Moser type inequality and the inductive assumption that

$$\begin{aligned} \begin{aligned} \bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j}(u\otimes u)\bigr \Vert _{L^2_{T_1}(H^{\ell -2j}(\Omega ))} \lesssim&\sum _{j_1+j_2=j}\bigl \Vert t^{j_1+\frac{1}{2}}\partial _t^{j_1}u\bigr \Vert _{L^\infty _{T_1}(H^{2}(\Omega ))}\\&\times \bigl \Vert t^{\frac{\ell -2}{2}-j+j_2}\partial _t^{j_2}u\bigr \Vert _{L^2_{T_1}(H^{\ell -2j}(\Omega ))}\\ \le&C_{\ell , T_1}(\Vert u_0\Vert _{H^1(\Omega )}). \end{aligned} \end{aligned}$$

As a result, for any \(j\le \frac{\ell -1}{2},\) we arrive at

$$\begin{aligned} \begin{aligned} \bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j}u\bigr \Vert _{L^2_{T_1}(H^{\ell +1-2j}(\Omega ))} \le&C_{\ell , T_1}(\Vert u_0\Vert _{H^1(\Omega )})+\bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^{j+1}u\bigr \Vert _{L^2_{T_1}(H^{\ell -1-2j}(\Omega ))}, \end{aligned} \end{aligned}$$

from which, with (A.21), we deduce by an iterative argument that

$$\begin{aligned} \sum _{0\le j \le \frac{\ell }{2}}\bigl \Vert t^{\frac{\ell -1}{2}}\partial _t^ju\bigr \Vert _{L^2_{T_1}(H^{\ell +1-2j}(\Omega ))} \le C_{\ell ,T_1}(\Vert u_0\Vert _{H^1(\Omega )}). \end{aligned}$$
(A.24)

By combining (A.23) and (A.24), we obtain that (A.2) holds for \(p=\ell .\) This finishes the proof of (A.2) and therefore the proof of Theorem 2.1. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liao, J., Sueur, F. & Zhang, P. Smooth Controllability of the Navier–Stokes Equation with Navier Conditions: Application to Lagrangian Controllability. Arch Rational Mech Anal 243, 869–941 (2022). https://doi.org/10.1007/s00205-021-01744-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00205-021-01744-2

Navigation