Skip to main content
Log in

On the Serrin-Type Condition on One Velocity Component for the Navier–Stokes Equations

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

Abstract

In this paper we consider the regularity problem of the Navier–Stokes equations in \( {\mathbb {R}}^{3} \). We show that the Serrin-type condition imposed on one component of the velocity \( u_3\in L^p(0,T; L^q({\mathbb {R}}^{3} ))\) with \( \frac{2}{p}+ \frac{3}{q} <1\), \( 3<q \le +\infty \) implies the regularity of the weak Leray solution \( u: {\mathbb {R}}^{3} \times (0,T) \rightarrow {\mathbb {R}}^{3} \), with the initial data belonging to \( L^2({\mathbb {R}}^3) \cap L^3({\mathbb {R}}^{3})\). The result is an immediate consequence of a new local regularity criterion in terms of one velocity component for suitable weak solutions.

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. Bae, H.-O., Choe, H.J.: A regularity criterion for the Navier–Stokes equations. Commun. Partial Differ. Equ. 32(7–9), 1173–1187, 2007

    Article  MathSciNet  Google Scholar 

  2. Bae, H.-O., Wolf, J.: A local regularity condition involving two velocity components of Serrin-type for the Navier–Stokes equations. C. R. Math. Acad. Sci. Paris 354(2), 167–174, 2016

    Article  MathSciNet  Google Scholar 

  3. da Veiga, H.: Beirao: a sufficient condition on the pressure for the regularity of weak solutions to the Navier–Stokes equations. J. Math. Fluid Mech. 2(2), 99–106, 2000

    Article  MathSciNet  Google Scholar 

  4. da Veiga, H.: Beirao: a new regularity class for the Navier–Stokes equations. Chin. Ann. Math. Ser. B 16(4), 407–412, 1995

    MATH  Google Scholar 

  5. Berselli, L.C.: Some criteria concerning the vorticity and the problem of global regularity for the 3D Navier–Stokes equations. Ann. Univ. Ferrara Sez. VII Sci. Mat. 55(2), 209–224, 2009

    Article  MathSciNet  Google Scholar 

  6. Berselli, L.C., Galdi, G.P.: Regularity criteria involving the pressure for the weak solutions to the Navier–Stokes equations. Proc. Am. Math. Soc. 130(12), 3585–3595, 2002

    Article  MathSciNet  Google Scholar 

  7. Caffarelli, L., Kohn, R., Nirenberg, L.: Partial regularity of suitable weak solutions of the Navier–Stokes equations. Commun. Pure Appl. Math. 35, 771–831, 1982

    Article  ADS  MathSciNet  Google Scholar 

  8. Cao, C., Titi, E.S.: Global regularity criterion for the 3D Navier–Stokes equations involving one entry of the velocity gradient tensor. Arch. Ration. Mech. Anal. 202(3), 919–932, 2011

    Article  MathSciNet  Google Scholar 

  9. Cao, C., Titi, E.S.: Regularity criteria for the three-dimensional Navier–Stokes equations. Indiana Univ. Math. J. 57(6), 2643–2661, 2008

    Article  MathSciNet  Google Scholar 

  10. Chae, D., Choe, H.J.: Regularity of solutions to the Navier–Stokes equation. Electron. J. Differ. Equ. 05, 7, 1999

    MathSciNet  MATH  Google Scholar 

  11. Chae, D., Lee, J.: Regularity criterion in terms of pressure for the Navier–Stokes equations. Nonlinear Anal. 46(5), 727–735, 2001

    Article  MathSciNet  Google Scholar 

  12. Chae, D., Wolf, J.: On the Liouville type theorems for self-similar solutions to the Navier–Stokes equations. Arch. Ration. Mech. Anal. 225, 549–572, 2017

    Article  MathSciNet  Google Scholar 

  13. Chemin, J.-Y., Zhang, P.: On the critical one component regularity for 3-D Navier–Stokes equations. Ann. Sci. Norm. Supr. (4) 49(1), 131–167, 2016

    Article  Google Scholar 

  14. Chemin, J.-Y., Zhang, P., Zhang, Z.: On the critical one component regularity for 3-D Navier–Stokes system: general case. Arch. Ration. Mech. Anal. 224(3), 871–905, 2017

    Article  MathSciNet  Google Scholar 

  15. Escauriaza, L., Sergin, G., Šverák, V.: Backward uniqueness for parabolic equations. Arch. Ration. Mech. Anal. 169, 147–157, 2003

    Article  MathSciNet  Google Scholar 

  16. Giaquinta, M.: Multiple Integrals in the Calculus of Variations and Nonlinear Elliptic systems. Princeton University Press, Princeton 1983

    MATH  Google Scholar 

  17. Hopf, E.: Über die Anfangswertaufgabe für die hydrodynamischen Grundgleichungen. Math. Nachr. 4, 213–231, 1951

  18. Kato, T.: \( L^p-\)solutions of the Navier–Stokes equation in \({\mathbb{R}}^m\) with applications to weak solutions. Math. Z. 187(4), 471–480, 1984

    Article  MathSciNet  Google Scholar 

  19. Kukavica, I., Ziane, M.: One component regularity for the Navier–Stokes equations. Nonlinearity 19(2), 453–469, 2006

    Article  ADS  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  21. Neustupa, J.: A contribution to the theory of regularity of a weak solution to the Navier–Stokes equations via one component of velocity and other related quantities. J. Math. Fluid Mech. 20, 1249–1267, 2018

    Article  ADS  MathSciNet  Google Scholar 

  22. Neustupa, J., Novotny, A., Penel, P.: An interior regularity of a weak solution to the Navier–Stokes equations in dependence on one component of velocity. In: Topics in Mathematical Fluid Mechanics, pp. 163–183, Quad. Mat., 10, Department of Mathematics, Seconda Universita’ degli Studi di Napoli, Caserta, 2002

  23. Neustupa, J., Penel, P.: Regularity of a Suitable Weak Solution to the Navier–Stokes Equations as a Consequence of Regularity of One Velocity Component, Applied Nonlinear Analysis, pp. 391–402. Kluwer/Plenum, New York 1999

    MATH  Google Scholar 

  24. Prodi, G.: Un teorema di unicita per le equazioni di Navier–Stokes. Ann. Mat. Pura Appl. 48, 173–182, 1959

    Article  MathSciNet  Google Scholar 

  25. Pokorný, M., Zhou, Y.: On the regularity of the solutions of the Navier–Stokes equations via one velocity component. Nonlinearity 23(5), 1097–1107, 2010

    Article  ADS  MathSciNet  Google Scholar 

  26. Seregin, G., Šverák, V.: Navier–Stokes equations with lower bounds on the pressure. Arch. Ration. Mech. Anal. 163(1), 65–86, 2002

    Article  MathSciNet  Google Scholar 

  27. Scheffer, V.: Partial regularity of solutions to the Navier–Stokes equations. Pac. J. Math. 66, 535–552, 1976

    Article  MathSciNet  Google Scholar 

  28. Scheffer, V.: Hausdorff measure and the Navier–Stokes equations. Commun. Math. Phys. 55, 97–102, 1977

    Article  ADS  MathSciNet  Google Scholar 

  29. Serrin, J.: On the interior regularity of weak solutions of the Navier–Stokes equations. Arch. Ration. Mech. Anal. 9, 187–191, 1962

    Article  MathSciNet  Google Scholar 

  30. Sohr, H.: The Navier–Stokes Equations. An Elementary Functional Analytic Approach. Birkhäuser, Basel 2001

    MATH  Google Scholar 

  31. Wolf, J.: A regularity criterion of Serrin-type for the Navier–Stokes equations involving the gradient of one velocity component. Analysis (Berlin) 35(4), 259–292, 2015

    MathSciNet  MATH  Google Scholar 

  32. Wolf, J.: On the local regularity of suitable weak solutions to the generalized Navier–Stokes equations. Ann. Univ. Ferrara 61, 149–171, 2015

    Article  MathSciNet  Google Scholar 

  33. Zhou, Y.: On a regularity criterion in terms of the gradient of pressure for the Navier–Stokes equations. Z. Angew. Math. Phys. 57(3), 384–392, 2006

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

Chae was partially supported by NRF Grant 2016R1A2B3011647 and 2021R1A2C1003234, while Wolf has been supported supported by NRF Grant 2017R1E1A1A01074536. The authors declare that they have no conflict of interest.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to D. Chae.

Additional information

Communicated by V. Šverák

Publisher's Note

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

A Appendix

A Appendix

The aim of this appendix is to provide an estimate which will be used various times for the estimation of integrals involving the pressure during the proof of our main result. We start our discussion by defining a singular integral operator, which is necessary for the decomposition of the pressure. Throughout this appendix, let \( 0< R \le 1\) be fixed. Let \( Q_0(R) = U_0(R) \times (-1,0)\), where \( U_0(R) = B'(R)\times (-1,1)\). Given \( f_{ ij}\in L^p(Q_0(R)), 1<p< +\infty , i,j=1,2,3,\) we define

$$\begin{aligned} {\mathscr {J}}(f)(x,t) = P.V. \int \nolimits _{ {\mathbb {R}}^{3} } K(x-y) :f(y,t) \chi _{ U_0(R)}(y)\mathrm{d}y,\quad (x,t)\in {\mathbb {R}}^{3}\times (-1,0), \end{aligned}$$

with the Calderón-Zygmund kernel \( K_{ ij}= \partial _{ i}\partial _j N, i,j=1, 2,3,\) where

$$\begin{aligned} N(x) = \frac{1}{4\pi |x|},\quad x\in {\mathbb {R}}^{3} \setminus \{0\}. \end{aligned}$$

Clearly, by virtue of Calderón-Zygmund’s inequality, \( {\mathscr {J}}: L^p(Q_0(R)) \rightarrow L^p({\mathbb {R}}^{3} \times (-1,0))\) defines a bounded linear operator. In particular,

$$\begin{aligned} \Vert {\mathscr {J}}(f)\Vert _{ L^p({\mathbb {R}}^{3} )} \le c \Vert f\Vert _{ L^p(U_0(R))}. \end{aligned}$$
(A.1)

Furthermore, setting \( \pi _0= {\mathscr {J}}(f)\), it holds that

$$\begin{aligned} -\Delta \pi _0 = \nabla \cdot \nabla \cdot f\quad \text {in}\quad Q_0(R), \end{aligned}$$
(A.2)

in the sense of distributions. As in Section 2 we use the following notation:

$$\begin{aligned} r_j= 2^{ -j},\quad U_j(R) = B'(1)\times (-r_j, r_j),\quad Q_j(R) = U_j(R) \times (-r_j^2, 0)\quad j\in {\mathbb {N}}_0. \end{aligned}$$

Lemma A.1

Let \( n\in {\mathbb {N}}\). Let \( \Psi \in C^\infty ({\mathbb {R}}^3\times (-\infty , 0))\) such that for constants \( \alpha>0, c>0 \) and \( C>0\) it holds that

$$\begin{aligned} {\left\{ \begin{array}{ll} c r_j^{ -\alpha } \le \Psi (x, t) \le Cr_j^{ -\alpha } \quad \forall (x, t)\in A_j =Q_j(R) \setminus Q_{ j+1}(R),\quad \forall j=0, \ldots , n-1 \\ c r_n^{ -\alpha } \le \Psi (x, t) \le Cr_j^{ -\alpha } \quad \forall (x, t)\in Q_n(R). \end{array}\right. } \end{aligned}$$
(A.3)

Let \( 1< p,q<+\infty , 1 < l \le q'\). Let \( v\in L^{ p}(-1, 0; L^q(U_0(R)))\), and \( f\in L^{ p'}(-1, 0; L^{ q'}(U_0(R)))\), \( 1< m, l < +\infty \). Then, setting \( \pi _0 = {\mathscr {J}}(f)\) it holds that

$$\begin{aligned}&\int \nolimits _{-1}^{ t}\int \nolimits _{U_0(R)} \pi _0 v \Psi \eta \mathrm{d}x\mathrm{d}s \nonumber \\&\quad \le c \sup |\eta | \sum _{k=0}^{n} r_k^{ -\alpha } \Vert f\Vert _{ L^{ p'}(-r_k^2, 0; L^{ q'}(U_k(R)))} \Vert v\Vert _{ L^{p}(-r_k^2, 0; L^{ q}(U_k(R)))} \nonumber \\&\qquad + c\sup |\eta |\sum _{j=0}^{n} \sum _{k=j}^{n} r_k^{ \frac{1}{q'}-\alpha } r_{ j}^{ \frac{2}{q'}- \frac{3}{l}} \Vert f\Vert _{L^{ p'}(-r_k^2, 0; L^{ l}(U_j(R)))} \Vert v\Vert _{ L^{p}(-r_k^2, 0; L^{ q}(U_k(R)))}, \end{aligned}$$
(A.4)

where \( \eta \in C^\infty _c(U_0(R)\times (-1, 0])\) stands for a cut off function. The constant in (A.4) depends only on pq and l.

Proof

Let \( f\in L^p(Q_0(R))\). Set \( \pi _0 = {\mathscr {J}}(f)\). For \( j\in {\mathbb {N}}_0\) let \( \chi _j\in C^\infty _c(U_j(R) \times (-r_j^2, 0] )\) with \( \chi _j=1\) on \( Q_{ j+1}(R)\) such that \( 0 \le \chi _j \le 1\), and \( |\partial _3\chi _j| \le c r_j^{ -1}\), and \( |\nabla ' \chi _j| \le c R^{ -1}\). We set

$$\begin{aligned} \phi _j={\left\{ \begin{array}{ll} 1- \chi _0\quad &{}\text {if}\quad j=0, \\ \chi _{ j}- \chi _{ j+1} &{}\text {if}\quad j=1, \ldots , n-1, \\ \chi _n\quad &{}\text {if}\quad j=n. \end{array}\right. } \end{aligned}$$

We have \( \displaystyle \sum _{j=0}^n\phi _j= 1- \chi _0 + \chi _0- \chi _1 + \cdots +\chi _{ n-1}- \chi _n + \chi _n=1\). Accordingly, \( f = \displaystyle \sum _{j=0}^{n} f \phi _j\), and therefore it holds that

$$\begin{aligned} \pi _0 = {\mathscr {J}}(f) = \sum _{j=0}^{n} {\mathscr {J}}(\phi _j f) = \sum _{j=0}^{n} \pi _{ 0,j}. \end{aligned}$$

This yields

$$\begin{aligned}&\int \nolimits _{-0}^{ t}\int \nolimits _{U_0(R)} \pi _0 v\Psi \eta \mathrm{d}x\mathrm{d}s = \sum _{k=0}^{n}\int \nolimits _{-1}^{ t}\int \nolimits _{U_0(R)} \pi _0 v \Psi \phi _k\eta \mathrm{d}x\mathrm{d}s \\&\quad = \sum _{j=0}^{n}\sum _{k=0}^{n} \int \nolimits _{-1}^{ t}\int \nolimits _{U_0(R)} \pi _{ 0,j} v \Psi \phi _k \eta \mathrm{d}x\mathrm{d}s = \sum _{k=0}^{n}\sum _{j=k}^{n}\int \nolimits _{-1}^{ t}\int \nolimits _{U_0(R)} \pi _{ 0,j} v\Psi \phi _k \eta \mathrm{d}x\mathrm{d}s \\&\quad \qquad + \sum _{j=0}^{n} \sum _{k=j+1}^{n}\int \nolimits _{-1}^{ t}\int \nolimits _{U_0(R)} \pi _{ 0,j} v \Psi \phi _k \eta \mathrm{d}x\mathrm{d}s = I+II. \end{aligned}$$

First, we calculate

$$\begin{aligned} I = \sum _{k=0}^{n}\int \nolimits _{-1}^{ t}\int \nolimits _{U_0(R)} \Pi _{ 0,k} v\Psi \phi _k\eta \mathrm{d}x\mathrm{d}s, \end{aligned}$$

where

$$\begin{aligned} \Pi _{ 0,k}= {\left\{ \begin{array}{ll} \pi _0\quad &{} \text {if}\quad k=0, \\ {\mathscr {J}} (\phi _k f)\quad &{}\text {if}\quad k=1, \ldots , n. \end{array}\right. } \end{aligned}$$

Observing (A.3), applying Hölder’s inequality along with (A.1), we get

$$\begin{aligned} I&\le c \sup |\eta | \sum _{k=0}^{n} r_k^{ -\alpha } \Vert \Pi _{ 0,k}\Vert _{ L^{ p'}(-r_k^2, 0; L^{ q'}(U_k(R)))} \Vert v\Vert _{ L^{ p}(-r_k^2, 0; L^{ q}(U_k(R)))} \\&\le c \sup |\eta |\sum _{k=0}^{n} r_k^{ -\alpha } \Vert f\Vert _{ L^{ p'}(-r_k^2, 0; L^{ q'}(U_k(R)))} \Vert v\Vert _{ L^{p}(-r_k^2, 0; L^{q}(U_k(R)))}. \end{aligned}$$

For the second integral we find that

$$\begin{aligned} II&= \sum _{j=n-2}^{n} \sum _{k=j}^{n} \int \nolimits _{-1}^{ t}\int \nolimits _{U_0(R)} \Pi _{ 0,j} v\Psi \phi _{ k} \eta \mathrm{d}x\mathrm{d}s \\&\qquad +\sum _{j=0}^{n-3} \sum _{k=j}^{j+3} \int \nolimits _{-1}^{ t}\int \nolimits _{U_0(R)} \Pi _{ 0,j} v\Psi \phi _{k} \eta \mathrm{d}x\mathrm{d}s \\&\qquad + \sum _{j=0}^{n-3} \sum _{k=j+4}^{n}\int \nolimits _{-1}^{ t}\int \nolimits _{U_0(R)} \Pi _{ 0,j} v\Psi \phi _k \eta \mathrm{d}x\mathrm{d}s = II_1+ II_2+ II_3. \end{aligned}$$

Arguing as above, observing (A.3) and applying Hölder’s inequality and (A.1), we see that

$$\begin{aligned} II_1+II_2 \le c \sup |\eta | \sum _{k=1}^{n} r_k^{ -\alpha } \Vert f\Vert _{ L^{ p'}(-r_k^2, 0; L^{ q'}(U_k(R)))} \Vert v\Vert _{ L^{ p}(-r_k^2, 0; L^{ q}(U_k(R)))}. \end{aligned}$$

It remains to estimate \( II_3\). We calculate

$$\begin{aligned} II_3 = \sum _{j=0}^{n-2} \sum _{k=j+4}^{n}\int \nolimits _{-1}^{ t}\int \nolimits _{U_0(R)} \Pi _{ 0,j} v \Psi \phi _k \mathrm{d}x\mathrm{d}s = \sum _{j=0}^{n-2} \sum _{k=j+4}^{n} J_{ jk} . \end{aligned}$$
(A.5)

Let \( j+4 \le k \le n\) be fixed. Applying Hölder’s inequality, together with (A.3), we find that

$$\begin{aligned} J_{ jk}&\le c\sup |\eta | r_k^{ -\alpha } \Vert \Pi _{ 0, j}\Vert _{ L^{ p'}(-r_k^2, 0; L^{ q'}(U_k(R)))} \Vert v\Vert _{ L^{p}(-r_k^2, 0; L^{ q}(U_k(R)))}. \end{aligned}$$

From the definition of \( \Pi _{ 0, j}\) it follows that \( \Delta \pi _{ 0, j} = \nabla \cdot \nabla \cdot (f\phi _j)\) in the sense of distributions. Since \( {\text {supp}}(\phi _j) \subset Q_{ j}(R) \setminus Q_{ j+2}(R)\) the function \( \pi _{ 0, j}\) is harmonic in \( {\mathbb {R}}^{2}\times (-r_{ j+2}, r_{ j+2})\times (-r_{ j+2}^2, 0)\). Applying Lemma A.2 below for \( h=\pi _{ 0, j}, r=r_{ j+2} \) and \( \rho = r_k\), we get, for almost all \( s\in (-r_k^2, 0)\),

$$\begin{aligned} \Vert \Pi _{ 0, j}(s)\Vert _{ L^{ q'}(U_k(R))}&=\Vert \pi _{ 0, j}(s)\Vert _{ L^{ q'}(B'(R)\times (-r_k, r_k))} \\&\le c\sup |\eta | r_k^{ \frac{1}{q'}} r_{ j+2}^{ \frac{2}{q'}- \frac{3}{l}} \Vert \Pi _{ 0, j}(s)\Vert _{ L^{ l}({\mathbb {R}}^{3} )}. \end{aligned}$$

Taking the \( L^{ p'}\) norm with respect to s, and employing (A.1), we find that

$$\begin{aligned} \Vert \Pi _{ 0, j}\Vert _{L^{ p'}(-r_k^2,0; L^{ q'}(U_k(R)))} \le c\sup |\eta | r_k^{ \frac{1}{q'}} r_{ j+2}^{ \frac{2}{q'}- \frac{3}{l}} \Vert f\Vert _{L^{ p'}(-r_k^2, 0; L^{ l}(U_j(R)))}. \end{aligned}$$

Accordingly,

$$\begin{aligned} J_{ jk}&\le c \sup |\eta | r_k^{ \frac{1}{q'}-\alpha } r_{ j+2}^{ \frac{2}{q'}- \frac{3}{l}} \Vert f\Vert _{L^{ p'}(-r_k^2, 0; L^{ l}(U_j(R)))} \Vert v\Vert _{ L^{p}(-r_k^2, 0; L^{ q}(U_k(R)))}. \end{aligned}$$

Inserting this inequality into (A.5), we arrive at

$$\begin{aligned} II_3&= c\sup |\eta |\sum _{j=1}^{n-2} \sum _{k=j+3}^{n} r_k^{ \frac{1}{q'}-\alpha } r_{ j+2}^{ \frac{2}{q'}- \frac{3}{l}} \Vert f\Vert _{L^{ p'}(-r_k^2, 0; L^{ l}(U_j(R)))} \Vert v\Vert _{ L^{p}(-r_k^2, 0; L^{ q}(U_k(R)))} \\&\le c\sup |\eta |\sum _{j=0}^{n} \sum _{k=j}^{n} r_k^{ \frac{1}{q'}-\alpha } r_{ j}^{ \frac{2}{q'}- \frac{3}{l}} \Vert f\Vert _{L^{ p'}(-r_k^2, 0; L^{ l}(U_j(R)))} \Vert v\Vert _{ L^{p}(-r_k^2, 0; L^{ q}(U_k(R)))}. \end{aligned}$$

Combining the above estimates, we get the claim. \(\quad \square \)

Lemma A.2

Let \( 0< r \le R <+\infty \). Let \( h: B'(2R)\times (-r, r) \rightarrow {\mathbb {R}}\) be harmonic. Then for all \( 0< \rho \le \frac{r}{4} \) and \( 1 \le l \le p \le +\infty \) we get

$$\begin{aligned} \Vert h\Vert _{ L^p(B'(R)\times (-\rho , \rho ))}^p \le c\rho r^{2 -3\frac{p}{l}} \Vert h\Vert _{ L^l(B'(2R)\times (-r, r))}^p, \end{aligned}$$
(A.6)

where c stands for a positive constant depending only on p and l.

Proof

Let \( k\in {\mathbb {N}}, k \ge 2\). Set \( \rho _k = 2^{ -k} r \). Since \( B'(2R)\times (- r, r)\) is a non isotropic cylinder, in order to apply the mean value property of harmonic functions we use a covering argument. We may choose a finite family of points \( \{x'_{\nu }\}\) in \( B'(R)\) such that \( \{B'(x_{\nu }', r/4)\}\) is a covering of \( \overline{B'(R)}\), and it holds that

$$\begin{aligned} \sum _{\nu } \chi _{ B'(x'_{ \nu }, r)} \le N,\quad |x_{\nu }- x_{ \mu }| \ge \frac{r}{4}\quad \forall \nu \ne \mu , \end{aligned}$$
(A.7)

where N stands for an absolute number. Setting \( x_{\nu }= (x'_{ \nu }, 0)\), we see that \( B'(x'_{\nu }, r/4 )\times (- r/4, r/4) \subset B(x_{\nu }, r/2)\). With this notation we have that

$$\begin{aligned} \Vert h\Vert _{ L^p(B'(R)\times (-\rho _k, \rho _k))}^p&\le \sum _{\nu }\Vert h\Vert _{ L^p(B'(x'_{ \nu }, r/4)\times (- \rho _k, \rho _k))}^p\nonumber \\&\le c r^2 \rho _k\sum _{\nu }\Vert h\Vert _{ L^\infty (B'(x'_{ \nu }, r/4)\times (- r/4, r/4))}^p \nonumber \\&\le c r^2 \rho _k\sum _{\nu }\Vert h\Vert _{ L^\infty (B(x_{ \nu }, r/2))}^p. \end{aligned}$$
(A.8)

Since h is harmonic, using the mean value property, we find that

$$\begin{aligned} \Vert h\Vert ^p_{ L^\infty (B(x_{\nu }, r/2 ))}&\le c r^{ -\frac{3p}{l}} \Vert h\Vert ^p_{ L^l(B(x_{\nu }, r ))} \\&\le c r^{ -\frac{3p}{l}} \Vert h\Vert ^{ p-l}_{ L^l( B'(2R)\times (-r,r))} \Vert h\Vert ^l_{ L^l(B'(x'_{\nu }, r )\times (-r,r))} \\&= c r^{ -\frac{3p}{l}} \Vert h\Vert ^{ p-l}_{ L^l( B'(2R)\times (-r,r))} \int \nolimits _{-r}^r \int \nolimits _{B'(2R)} |h|^l \chi _{ B'(x'_{\nu }, r )} \mathrm{d}x'\mathrm{d}x_3. \end{aligned}$$

Taking the sum over \( \nu \) and using (A.7), we obtain that

$$\begin{aligned} \sum _{\nu }\Vert h\Vert _{ L^\infty (B(x_{ \nu }, r/2))}^p \le cN r^{ -\frac{3p}{l}} \Vert h\Vert ^{ p}_{ L^l( B'(2R)\times (-r,r))}. \end{aligned}$$
(A.9)

Combining (A.8) and (A.9), we get

$$\begin{aligned} \Vert h\Vert _{ L^p(B'(R)\times (-\rho _k , \rho _k))}^p \le c\rho _k r^{2 -3\frac{p}{l}} \Vert h\Vert _{ L^l(B'(2R)\times (-r, r))}^p. \end{aligned}$$
(A.10)

Let \( 0< \rho \le \frac{r}{4}\). Then there exists a unique integer \( k\ge 2\) such that \( \rho _{ k+1} < \rho \le \rho _k\), Thus, (A.10) implies

$$\begin{aligned} \Vert h\Vert _{ L^p(B'(R)\times (-\rho , \rho ))}^p&\le c\rho _k r^{2 -3\frac{p}{l}} \Vert h\Vert _{ L^l(B'(2R)\times (-r, r))}^p \\&\le 2c\rho r^{2 -3\frac{p}{l}} \Vert h\Vert _{ L^l(B'(2R)\times (-r, r))}^p, \end{aligned}$$

whence we get (A.6). \(\quad \square \)

The following iteration lemma has been used in the sequel of the proofs above, and its proof can be found in [16, V. Lemma 3.1]:

Lemma A.3

Let f(t) be a nonnegative bounded function defined in \( [r_0, r_1], 0 \le r_0< r_1 <+\infty \). Suppose that for \( r_0 \le t < s \le r_1\) we have

$$\begin{aligned} f(t)\le [A(s-t)^{-\alpha } + B] + \theta f(s), \end{aligned}$$
(A.11)

where \( A,B, \alpha , \theta \) are nonnegative constants with \( 0 \le \theta <1\). Then, for all \( r_0 \le \rho <R \le r_1\), we have

$$\begin{aligned} f(\rho )\le c[A(R-\rho )^{-\alpha } + B], \end{aligned}$$
(A.12)

where c is a constant depending on \( \alpha \) and \( \theta \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chae, D., Wolf, J. On the Serrin-Type Condition on One Velocity Component for the Navier–Stokes Equations. Arch Rational Mech Anal 240, 1323–1347 (2021). https://doi.org/10.1007/s00205-021-01636-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00205-021-01636-5

Navigation