Skip to main content
Log in

Sharp nonuniqueness for the Navier–Stokes equations

  • Published:
Inventiones mathematicae Aims and scope

Abstract

In this paper, we prove a sharp nonuniqueness result for the incompressible Navier–Stokes equations in the periodic setting. In any dimension \(d \ge 2\) and given any \( p<2\), we show the nonuniqueness of weak solutions in the class \(L^{p}_t L^\infty \), which is sharp in view of the classical Ladyzhenskaya–Prodi–Serrin criteria. The proof is based on the construction of a class of non-Leray–Hopf weak solutions. More specifically, for any \( p<2\), \(q<\infty \), and \(\varepsilon >0\), we construct non-Leray–Hopf weak solutions \( u \in L^{p}_t L^\infty \cap L^1_t W^{1,q}\) that are smooth outside a set of singular times with Hausdorff dimension less than \(\varepsilon \). As a byproduct, examples of anomalous dissipation in the class \(L^{ {3}/{2} - \varepsilon }_t C^{ {1}/{3}} \) are given in both the viscous and inviscid case.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. It is possible to consider more general initial data, such as \(L^p \) for some \( 1 \le p \le \infty \) as in [32].

  2. Our setting is on the d-dimensional torus and the initial data is always \(L^2\).

  3. This interpretation was made explicit in [68].

  4. Here by nontrivial we mean that the singular set is not empty or full since smooth solutions have no singularity while the singular set of the solutions in [11] is the whole space-time domain.

  5. In fact, \(\tau _n\) can be taken to be 1/2 if we use Proposition 2.2 instead of Theorem 1.8.

References

  1. Albritton, D., Brué, E., Colombo, M.: Non-uniqueness of Leray solutions of the forced Navier–Stokes equations. Ann. Math., to appear (2022)

  2. Amann, H.: On the strong solvability of the Navier–Stokes equations. J. Math. Fluid Mech. 2(1), 16–98 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  3. Beale, J.T., Kato, T., Majda, A.: Remarks on the breakdown of smooth solutions for the \(3\)-D Euler equations. Commun. Math. Phys. 94(1), 61–66 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  4. 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:1809.00600 (2018)

  5. Buckmaster, T., De Lellis, C., Székelyhidi Jr., L., Vicol, V.: Onsager’s conjecture for admissible weak solutions. Commun. Pure Appl. Math., to appear (2018)

  6. Buckmaster, T., De Lellis, C., Isett, P., Székelyhidi, Jr., L.: Anomalous dissipation for \(1/5\)-Hölder Euler flows. Ann. Math. (2) 182(1), 127–172 (2015)

  7. Buckmaster, T., De Lellis, C., Székelyhidi, Jr, László: Dissipative Euler flows with Onsager-critical spatial regularity. Commun. Pure Appl. Math. 69(9), 1613–1670 (2016)

  8. Buckmaster, T., Masmoudi, N., Novack, M., Vicol, V.: Non-conservative \(H^{\frac{1}{2}}\)-weak solutions of the incompressible 3D euler equations. arXiv:2101.09278 (2021)

  9. Buckmaster, T., Vicol, V.: Convex integration constructions in hydrodynamics. Bull. Am. Math. Soc., To appear (2021)

  10. Buckmaster, T.: Onsager’s conjecture almost everywhere in time. Commun. Math. Phys. 333(3), 1175–1198 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  11. Buckmaster, T., Vicol, V.: Nonuniqueness of weak solutions to the Navier–Stokes equation. Ann. Math. (2) 189(1), 101–144 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  12. Buckmaster, T., Shkoller, S., Vicol, V.: Nonuniqueness of weak solutions to the SQG equation. Commun. Pure Appl. Math. 72(9), 1809–1874 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  13. Caffarelli, L., Kohn, R., Nirenberg, L.: Partial regularity of suitable weak solutions of the Navier–Stokes equations. Commun. Pure Appl. Math. 35(6), 771–831 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  14. Chemin, J.-Y.: About weak-strong uniqueness for the 3D incompressible Navier–Stokes system. Commun. Pure Appl. Math. 64(12), 1587–1598 (2011)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  16. Cheskidov, A., Luo, X.: Nonuniqueness of weak solutions for the transport equation at critical space regularity. Ann. PDE, 7(1):Paper No. 2, 45 (2021)

  17. Cheskidov, A., Constantin, P., Friedlander, S., Shvydkoy, R.: Energy conservation and Onsager’s conjecture for the Euler equations. Nonlinearity 21(6), 1233–1252 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  18. Choi, K., Vasseur, A.F.: Estimates on fractional higher derivatives of weak solutions for the Navier–Stokes equations. Ann. Inst. H. Poincaré Anal. Non Linéaire 31(5), 899–945 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  19. Colombo, M., De Rosa, L., Sorella, M.: Typicality results for weak solutions of the incompressible Navier–Stokes equations. arXiv:2102.03244 (2021)

  20. Constantin, P., E, W., Titi, E.S.: Onsager’s conjecture on the energy conservation for solutions of Euler’s equation. Commun. Math. Phys 165(1), 207–209 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  21. Daneri, S., Runa, E., Székelyhidi, L.: Non-uniqueness for the Euler equations up to Onsager’s critical exponent. Ann. PDE, 7(1):Paper No. 8, 44 (2021)

  22. Daneri, S., Székelyhidi, Jr, L.: Non-uniqueness and h-principle for Hölder-continuous weak solutions of the Euler equations. Arch. Ration. Mech. Anal 224(2), 471–514 (2017)

  23. Daneri, S.: Cauchy problem for dissipative Hölder solutions to the incompressible Euler equations. Commun. Math. Phys. 329(2), 745–786 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  24. De Lellis, C., Székelyhidi, L., Jr.: The Euler equations as a differential inclusion. Ann. Math. (2) 170(3), 1417–1436 (2009)

  25. De Rosa, L., Haffter, S.: Dimension of the singular set of wild hölder solutions of the incompressible euler equations. arXiv:2102.06085 (2021)

  26. De Lellis, C., Székelyhidi, Jr., L.: Dissipative continuous Euler flows. Invent. Math 193(2), 377–407 (2013)

  27. De Lellis, C., Székelyhidi, Jr., L.: Dissipative Euler flows and Onsager’s conjecture. J. Eur. Math. Soc. 16(7), 1467–1505 (2014)

  28. Dong, H., Dapeng, D.: Partial regularity of solutions to the four-dimensional Navier–Stokes equations at the first blow-up time. Commun. Math. Phys. 273(3), 785–801 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  29. Duchon, J., Robert, R.: Inertial energy dissipation for weak solutions of incompressible Euler and Navier–Stokes equations. Nonlinearity 13(1), 249–255 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  30. Escauriaza, L., Seregin, G.A., Šverák, V.: \(L_{3,\infty }\)-solutions of Navier–Stokes equations and backward uniqueness. Uspekhi Mat. Nauk 58(2(350)), 3–44 (2003)

    MathSciNet  Google Scholar 

  31. Eyink, G.L.: Energy dissipation without viscosity in ideal hydrodynamics. I. Fourier analysis and local energy transfer. Phys. D 78(3–4), 222–240 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  32. Fabes, E.B., Jones, B.F., Rivière, N.M.: The initial value problem for the Navier–Stokes equations with data in \(L^{p}\). Arch. Rational Mech. Anal. 45, 222–240 (1972)

    Article  MathSciNet  MATH  Google Scholar 

  33. Frisch, U.: Turbulence. Cambridge University Press, Cambridge (1995). The legacy of A. N. Kolmogorov

  34. Fujita, H., Kato, T.: On the Navier–Stokes initial value problem, I. Arch. Rational Mech. Anal. 16, 269–315 (1964)

    Article  MathSciNet  MATH  Google Scholar 

  35. Furioli, G., Lemarié-Rieusset, P.G., Terraneo, E.: Unicité dans \(L^3({\mathbb{R}}^3)\) et d’autres espaces fonctionnels limites pour Navier–Stokes. Rev. Mat. Iberoamericana 16(3), 605–667 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  36. Galdi, G.P.: An Introduction to the Mathematical Theory of the Navier–Stokes Equations. Vol. II, vol. 39 of Springer Tracts in Natural Philosophy. Springer, New York (1994). Nonlinear steady problems

  37. Galdi, G.P.: An introduction to the Navier–Stokes initial-boundary value problem. In: Fundamental Directions in Mathematical Fluid Mechanics, Adv. Math. Fluid Mech., pp. 1–70. Birkhäuser, Basel (2000)

  38. Gallagher, I., Planchon, F.: On global infinite energy solutions to the Navier–Stokes equations in two dimensions. Arch. Ration. Mech. Anal. 161(4), 307–337 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  39. Gallagher, I., Koch, G.S., Planchon, F.: Blow-up of critical Besov norms at a potential Navier–Stokes singularity. Commun. Math. Phys. 343(1), 39–82 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  40. Germain, P.: Multipliers, paramultipliers, and weak-strong uniqueness for the Navier–Stokes equations. J. Differ. Equ. 226(2), 373–428 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  41. Guillod, J., Šverák, V.: Numerical investigations of non-uniqueness for the Navier–Stokes initial value problem in borderline spaces. arXiv:1704.00560 (2017)

  42. Hopf, E.: Über die anfangswertaufgabe für die hydrodynamischen grundgleichungen. erhard schmidt zu seinem 75. geburtstag gewidmet. Math. Nachr. 4(1–6), 213–231 (1951)

    MathSciNet  Google Scholar 

  43. Isett, P.: On the Endpoint Regularity in Onsager’s Conjecture. arXiv:1706.01549 (2017)

  44. Isett, P.: A proof of Onsager’s conjecture. Ann. Math. (2) 188(3), 871–963 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  45. Jia, H., Šverák, V.: Are the incompressible 3D Navier–Stokes equations locally ill-posed in the natural energy space? J. Funct. Anal. 268(12), 3734–3766 (2015)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  47. Koch, H., Tataru, D.: Well-posedness for the Navier–Stokes equations. Adv. Math. 157(1), 22–35 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  48. Kozono, H., Sohr, H.: Remark on uniqueness of weak solutions to the Navier–Stokes equations. Analysis 16(3), 255–271 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  49. Ladyženskaja, O.A.: Uniqueness and smoothness of generalized solutions of Navier–Stokes equations. Zap. Naučn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. 5, 169–185 (1967)

    MathSciNet  Google Scholar 

  50. Ladyženskaja, O.A.: Example of nonuniqueness in the hopf class of weak solutions for the Navier–Stokes equations. Math. USSR-Izvestiya 3(1), 229–236 (1969)

    Article  Google Scholar 

  51. Lemarié-Rieusset, P.G.: The Navier–Stokes Problem in the 21st Century. CRC Press, Boca Raton (2016)

    Book  MATH  Google Scholar 

  52. Leray, J.: Sur le mouvement d’un liquide visqueux emplissant l’espace. Acta Math. 63(1), 193–248 (1934)

    Article  MathSciNet  MATH  Google Scholar 

  53. Lions, P.-L., Masmoudi, N.: Uniqueness of mild solutions of the Navier–Stokes system in \(L^N\). Commun. Partial Differ. Equ. 26(11–12), 2211–2226 (2001)

    Article  MATH  Google Scholar 

  54. Luo, T., Titi, E.S.: Non-uniqueness of weak solutions to hyperviscous Navier–Stokes equations: on sharpness of J.-L. Lions exponent. Calc. Var. Partial Differ. Equ., 59(3):Paper No. 92, 15 (2020)

  55. Luo, X.: Stationary solutions and nonuniqueness of weak solutions for the Navier–Stokes equations in high dimensions. Arch. Ration. Mech. Anal. 233(2), 701–747 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  56. Luo, T., Peng, Q.: Non-uniqueness of weak solutions to 2D hypoviscous Navier–Stokes equations. J. Differ. Equ. 269(4), 2896–2919 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  57. Mandelbrot, B.: Intermittent turbulence and fractal dimension: kurtosis and the spectral exponent \(5/3+B\). In: Turbulence and Navier–Stokes equations (Proceedings of Conference, Univ. Paris-Sud, Orsay, 1975), pp. 121–145. Lecture Notes in Math., Vol. 565. Springer, Berlin (1976)

  58. Masuda, K.: Weak solutions of Navier–Stokes equations. Tohoku Math. J. (2) 36(4), 623–646 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  59. Meyer, Y.: Wavelets, paraproducts, and Navier–Stokes equations. In: Current Developments in Mathematics. 1996 (Cambridge, MA), pp. 105–212. Int. Press, Boston, MA (1997)

  60. Modena, S., Székelyhidi, Jr, L.: Non-uniqueness for the transport equation with Sobolev vector fields. Ann. PDE, 4(2):Art. 18, 38 (2018)

  61. Monniaux, S.: Uniqueness of mild solutions of the Navier–Stokes equation and maximal \(L^p\)-regularity. C. R. Acad. Sci. Paris Sér. I Math. 328(8), 663–668 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  62. Nash, J.: \(C^1\) isometric imbeddings. Ann. Math. 2(60), 383–396 (1954)

    Article  MATH  Google Scholar 

  63. Novack, M.: Non-uniqueness of weak solutions to the 3D quasi-geostrophic equations. SIAM J. Math. Anal., To appear (2020)

  64. Onsager, L.: Statistical hydrodynamics. Nuovo Cimento (9), 6 (Suppl, Supplemento, no. 2 (Convegno Internazionale di Meccanica Statistica)):279–287 (1949)

  65. Ożański, W.S.: Weak solutions to the Navier–Stokes inequality with arbitrary energy profiles. Commun. Math. Phys. 374(1), 33–62 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  66. Prodi, G.: Un teorema di unicità per le equazioni di Navier–Stokes. Ann. Mat. Pura ed Appl. 48(1), 173–182 (1959)

    Article  MathSciNet  MATH  Google Scholar 

  67. Robinson, J. C., Rodrigo, J. L., Sadowski, W.: The three-dimensional Navier–Stokes equations, vol. 157 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge (2016). Classical theory

  68. Scheffer, V.: Turbulence and Hausdorff dimension. In Turbulence and Navier–Stokes equations (Proc. Conf., Univ. Paris-Sud, Orsay, 1975), pp. 174–183. Lecture Notes in Math. Vol. 565 (1976)

  69. Scheffer, V.: Partial regularity of solutions to the Navier–Stokes equations. Pacific J. Math. 66(2), 535–552 (1976)

    Article  MathSciNet  MATH  Google Scholar 

  70. Scheffer, V.: Hausdorff measure and the Navier–Stokes equations. Commun. Math. Phys. 55(2), 97–112 (1977)

    Article  MathSciNet  MATH  Google Scholar 

  71. Scheffer, V.: The Navier–Stokes equations on a bounded domain. Commun. Math. Phys. 73(1), 1–42 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  72. Scheffer, V.: A solution to the Navier–Stokes inequality with an internal singularity. Commun. Math. Phys. 101(1), 47–85 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  73. Scheffer, V.: Nearly one-dimensional singularities of solutions to the Navier–Stokes inequality. Commmun. Math. Phys. 110(4), 525–551 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  74. Scheffer, V.: An inviscid flow with compact support in space-time. J. Geom. Anal. 3(4), 343–401 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  75. Serrin, J.: On the interior regularity of weak solutions of the Navier–Stokes equations. Arch. Ration. Mech. Anal. 9, 187–195 (1962)

    Article  MathSciNet  MATH  Google Scholar 

  76. Shnirelman, A.: On the nonuniqueness of weak solution of the Euler equation. Commun. Pure Appl. Math. 50(12), 1261–1286 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  77. Sohr, H., von Wahl, W.: On the singular set and the uniqueness of weak solutions of the Navier–Stokes equations. Manuscripta Math. 49(1), 27–59 (1984)

  78. Struwe, M.: On partial regularity results for the Navier–Stokes equations. Commun. Pure Appl. Math. 41(4), 437–458 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  79. Szekelyhidi, Jr, L.: From isometric embeddings to turbulence. In: HCDTE Lecture Notes. Part II. Nonlinear Hyperbolic PDEs, Dispersive and Transport Equations, vol. 7 of AIMS Ser. Appl. Math., page 63. Am. Inst. Math. Sci. (AIMS), Springfield, MO (2013)

  80. Tao, T.: 255B, Notes 2: Onsager’s conjecture. https://terrytao.wordpress.com/2019/01/08/255b-notes-2-onsagers-conjecture (2019)

  81. Vishik, M.: Instability and non-uniqueness in the Cauchy problem for the Euler equations of an ideal incompressible fluid. Part II. arXiv:1805.09440 (2018)

  82. Vishik, M.: Instability and non-uniqueness in the Cauchy problem for the Euler equations of an ideal incompressible fluid. Part I. arXiv:1805.09426 (2018)

  83. Wu, B.: Partially regular weak solutions of the Navier–Stokes equations in \({\mathbb{R}}^4\times [0,\infty [\). Arch. Ration. Mech. Anal. 239(3), 1771–1808 (2021)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

AC was partially supported by the NSF grant DMS-1909849. The authors are grateful to the anonymous referees for very helpful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexey Cheskidov.

Additional information

Publisher's Note

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

Appendices

Appendix A: \(X^{p,q}\) weak solutions on the torus

In this section, we show that sub-critical and critical weak solutions in \( X^{p,q}([0,T] ; {\mathbb {T}}^d)\) are in fact Leray–Hopf. In particular, by the weak-strong uniqueness of Ladyzhenskaya–Prodi–Serrin, this implies the uniqueness part of Theorem 1.3. The content of Theorem A.1 is classical [32, 35, 46, 53] and we include a proof in the regime \(q>2\) for the convenience of the readers. Note that the proof applies to the case \(q=\infty \) which is most relevant to the results of this paper, but was omitted in [32].

Theorem A.1

Let \(d \ge 2\) be the dimension and \(u \in X^{p,q}([0,T] ; {\mathbb {T}}^d)\) be a weak solution of (1.1) with \(\frac{2}{p} + \frac{d}{q} = 1\), \(d \le q \le \infty \). Then u is a Leray–Hopf solution.

We prove Theorem A.1 for \(q>2\) only. The case \(d=q=2\), as discussed in [38], can be handled by the argument of [35] in 2D. The method we present here follows the duality approach in [53] and use only classical ingredients.

The first ingredient is an existence result for a linearized Navier–Stokes equation.

Theorem A.2

Let \(u \in X^{p,q}([0,T] ; {\mathbb {T}}^d)\) be a weak solution of (1.1) with \(\frac{2}{p} + \frac{d}{q} = 1\). For any divergence-free \(v_0 \in L^2({\mathbb {T}}^d)\), there exists a weak solution \(v \in C_w L^2 \cap L^2_t H^1\) to the linearized Navier–Stokes equation:

$$\begin{aligned} {\left\{ \begin{array}{ll} \partial _t v -\Delta v + u \cdot \nabla v + \nabla p = 0 &{}\\ {{\,\mathrm{div}\,}}v =0, \end{array}\right. } \end{aligned}$$
(A.1)

satisfying the energy inequality

$$\begin{aligned} \frac{1}{2}\Vert v(t)\Vert _2^2+ \int _{t_0}^t \Vert \nabla v(s)\Vert _2^2 \, ds\le \frac{1}{2}\Vert v(t_0)\Vert _2^2 , \end{aligned}$$

for all \(t\in [t_0,T]\), a.e. \(t_0\in [0,T]\) (including \(t_0=0\)).

Proof

This follows by a standard Galerkin method and can be found in many textbooks. See [67, Chapter 4] or [36] for details. \(\square \)

Let v be the weak solution given by Theorem A.2 with initial data u(0). The goal is to show \(u \equiv v\). Setting \(w = u -v\), the equation for w reads

$$\begin{aligned} \partial _t w -\Delta w + u \cdot \nabla w + \nabla q = 0 , \end{aligned}$$

and its weak formulation

$$\begin{aligned} \int _0^T \int _{{\mathbb {T}}^d} w\cdot ( \partial _t \varphi + \Delta \varphi + u\cdot \nabla \varphi ) \, dx dt =0 \quad \text {for any }\varphi \in {\mathcal {D}}_T, \end{aligned}$$
(A.2)

where we recall that the test function class \({\mathcal {D}}_T \) consists of smooth divergence-free functions vanishing for \(t \ge T\).

Fix \(F \in C^\infty _c([0,T] \times {\mathbb {T}}^d)\). Let \(\Phi :[0,T]\times {\mathbb {T}}^d \rightarrow {\mathbb {R}}^d\) and \(\chi :[0,T]\times {\mathbb {T}}^d \rightarrow {\mathbb {R}}\) satisfy the system of equations

$$\begin{aligned} {\left\{ \begin{array}{ll} -\partial _t \Phi -\Delta \Phi - u \cdot \nabla \Phi + \nabla \chi = F&{}\\ {{\,\mathrm{div}\,}}\Phi =0 &{}\\ \Phi (T) =0. \end{array}\right. } \end{aligned}$$
(A.3)

Note that the equation of \( \Phi \) is “backwards in time” and by a change of variable one can convert (A.3) into a more conventional form.

If we can use \(\varphi =\Phi \) as the test function in the weak formulation (A.2), then immediately

$$\begin{aligned} \int _{[0,T] \times {\mathbb {T}}^d} w \cdot F \, dx dt = 0 . \end{aligned}$$

Since \(F \in C^\infty _c([0,T] \times {\mathbb {T}}^d)\) is arbitrary, we have

$$\begin{aligned} w = 0 \quad \text {for }a.e. (t,x) \in [0,T] \times {\mathbb {T}}^d. \end{aligned}$$

So the question of whether \(u \equiv v\) reduces to showing a certain regularity of \(\Phi \). More specifically, we can prove the following theorem.

Theorem A.3

Let \(d \ge 2\) be the dimension and \(u \in X^{p,q}([0,T] ; {\mathbb {T}}^d)\) be a weak solution of (1.1) with \(\frac{2}{p} + \frac{d}{q} = 1\) with \(q>2\). For any \(F \in C^\infty _c([0,T] \times {\mathbb {T}}^d) \), the system (A.3) has a weak solution \( \Phi \in L^\infty _t L^2 \cap L^2_t H^1\) such that \(\Phi \) can be used as a test function in (A.2).

Proof

We will prove the weak solution \(\Phi \) satisfies the regularity

$$\begin{aligned} \partial _t \Phi ,\Delta \Phi ,u\cdot \nabla \Phi , \nabla \chi \in L^2_{t,x} , \end{aligned}$$
(A.4)

which implies \( \Phi \) can be used in (A.2) since \(w \in L^2_{t,x}\).

The solution \( \Phi \) will be constructed by the Galerkin method of the following finite-dimensional approximation

$$\begin{aligned} {\left\{ \begin{array}{ll} -\partial _t \Phi _n -\Delta \Phi _n - P_n\big [ u_n \cdot \nabla \Phi _n\big ] + \nabla \chi _n = F_n &{}\\ {{\,\mathrm{div}\,}}\Phi _n =0 &{}\\ \Phi _n (T) =0, \end{array}\right. } \end{aligned}$$
(A.5)

where \(\Phi _n, u_n, \chi _n, F_n\) are restricted to the first n Fourier modes, and \(P_n\) is the projection operator on those modes.

It suffices to verify the following a priori estimates as they are preserved in the limit as \(n \rightarrow \infty \). We also only need to show the estimates for \(\partial _t \Phi \), \( \Delta \Phi \), and \(u\cdot \nabla \Phi \) since the pressure \( \chi \) satisfies the equation

$$\begin{aligned} \Delta \chi = {{\,\mathrm{div}\,}}( u \cdot \nabla \Phi )+ {{\,\mathrm{div}\,}}F. \end{aligned}$$

Step 1: Energy bounds \( L^\infty _t L^2 \cap L^2_t H^1 \).

The solutions to the Galerkin approximation (A.5) enjoy the energy estimate

$$\begin{aligned} -\frac{1}{2} \frac{d}{dt} \Vert \Phi \Vert _2^2 + \Vert \nabla \Phi \Vert _2^2 \le \int _{{\mathbb {T}}^d} |F \cdot \Phi |\,dx, \end{aligned}$$

which implies the desired energy bounds.

Step 2: Higher bounds \( L^\infty _t H^1 \cap L^2_t H^2 \).

We now take the \(L^2\) inner product of (A.5) with \(\Delta \Phi \) to obtain

$$\begin{aligned} -\frac{1}{2} \frac{d}{dt} \Vert \nabla \Phi \Vert _2^2 + \Vert \Delta \Phi \Vert _2^2 \le \int _{{\mathbb {T}}^d} |F \cdot \Delta \Phi |\,dx + \int _{{\mathbb {T}}^d} |u \cdot \nabla \Phi \cdot \Delta \Phi |\,dx. \end{aligned}$$
  • Case 1: \( p<\infty \) and \(d<q\le \infty \).

    Thanks to Proposition A.4,

    $$\begin{aligned} \int _{{\mathbb {T}}^d} |u \cdot \nabla \Phi \cdot \Delta \Phi |\,dx \lesssim \Vert u \Vert _q \Vert \nabla \Phi \Vert _2^{ \frac{q-d}{q}} \Vert \Delta \Phi \Vert _2^{ \frac{q+d}{q}}. \end{aligned}$$
    (A.6)

    Then Hölder’s, Poincaré’s, and Young’s inequalities yield

    $$\begin{aligned} -\frac{1}{2} \frac{d}{dt} \Vert \nabla \Phi \Vert _2^2 + \frac{1}{2}\Vert \Delta \Phi \Vert _2^2 \lesssim \Vert u\Vert _q^{\frac{2q}{q-d}}\Vert \nabla \Phi \Vert _2^2 + \Vert F\Vert _2^2. \end{aligned}$$

    Thanks to the integrability of \(\Vert u\Vert _q^{\frac{2q}{q-d}}\) (\(\Vert u\Vert _\infty ^2\) when \(q=\infty \)), Gronwall’s inequality immediately implies that \(\Phi \in L^\infty _t H^1 \cap L^2_t H^2\).

  • Case 2: \( p=\infty \) and \(q=d > 2\).

    Since \(u \in C_t L^{d}\), for any \(\varepsilon >0\) there exists a decomposition \(u = u_1 +u_2\) such that

    $$\begin{aligned} \Vert u_1 \Vert _{C_t L^{d} } \le \varepsilon \quad \text { and }\quad u_2 \in L^\infty _{t,x}, \end{aligned}$$

    and then the \(u_1\) portion of the nonlinear term in (A.6) can be absorbed by \(\Vert \Delta \Phi \Vert _2^2\), so we arrive at the same conclusion.

Step 3: Conclusion from maximal regularity of the heat equation.

Taking Leary’s projection \({\mathbb {P}}\) onto the divergence-free vector fields, the equation for \(\Phi \) can be rewritten as

$$\begin{aligned} {\left\{ \begin{array}{ll} -\partial _t \Phi -\Delta \Phi = {\mathbb {P}} (u \cdot \nabla \Phi ) + {\mathbb {P}} F&{}\\ \Phi (T) =0. \end{array}\right. } \end{aligned}$$

Therefore, by the maximal \(L^p_t L^q\) regularity of the heat equation (see for instance [67, Theorem 5.4]), we only need to show the estimates (A.4) for \(u \cdot \nabla \Phi \) to conclude the proof. By Sobolev interpolations we have \(\nabla \Phi \in L^r L^s \) for any \(\frac{2}{r} + \frac{d}{s} =\frac{d}{2} \) such that \( 2\le s \le \frac{2d}{d-2}\). We can find rs in this regime that satisfy the Hölder relations \( \frac{1}{p} + \frac{1}{r} = \frac{1}{2}\) and \( \frac{1}{q} + \frac{1}{s} = \frac{1}{2}\), where recall \(\frac{2}{p} + \frac{d}{q} =1\). Since \( u \in L^{p}_t L^q\) with some \(\frac{2}{p} + \frac{d}{q} =1\), this choice of rs implies that

$$\begin{aligned} u \cdot \nabla \Phi \in L^2([0,T] \times {\mathbb {T}}^d ) . \end{aligned}$$

\(\square \)

The last result is a classical estimate of the nonlinear term, cf. [67, pp. 172]. Notice that one of the embedding fails when \(q=2\) which is the reason we can only prove Theorem A.3 for \(q > 2\).

Proposition A.4

Let \( d \ge 2\) be the dimension and \(d \le q \le \infty \) such that \(q > 2\). For any smooth vector fields \(u,v \in C^\infty _0({\mathbb {T}}^d)\),

$$\begin{aligned} \int _{{\mathbb {T}}^d} |u \cdot \nabla v \cdot \Delta v | \,dx \lesssim \Vert u \Vert _{q} \Vert \nabla v \Vert _{2}^{\frac{q-d}{q}} \Vert \Delta v \Vert _{2}^{\frac{q+d}{q}}. \end{aligned}$$

Proof

We apply Hölder’s inequality with exponents \(\frac{1}{q} + \frac{1}{r} + \frac{1}{2} =1\), \(r = \frac{2q}{q-2} \in [2 , \frac{2d}{d-2})\),

$$\begin{aligned} \int _{{\mathbb {T}}^d} |u \cdot \nabla v \cdot \Delta v | \,dx \lesssim \Vert u \Vert _{q} \Vert \nabla v \Vert _{r} \Vert \Delta v \Vert _{ 2} . \end{aligned}$$

Since \( 2 \le r< \infty \), by the Sobolev embedding \(H^{s}({\mathbb {T}}^d ) \hookrightarrow L^{ r} ({\mathbb {T}}^d )\), \(s = d( \frac{1}{2} - \frac{1}{r})\),

$$\begin{aligned} \Vert \nabla v \Vert _{r} \lesssim \Vert \nabla v \Vert _{H^{s}} . \end{aligned}$$

Finally, by a standard Sobolev interpolation and the \(L^2\)-boundedness of Riesz transform, we have

$$\begin{aligned} \Vert \nabla v \Vert _{H^{s}} \lesssim \Vert \nabla v \Vert _{2}^{\frac{q- d}{q}}\Vert \Delta v \Vert _{2}^{\frac{d}{q} }, \end{aligned}$$

which concludes the proof. \(\square \)

Appendix B: Some technical tools

1.1 Improved Hölder’s inequality on \({\mathbb {T}}^d\)

We recall the following result due to Modena and Székelyhidi [60], which was inspired by [11, Lemma 3.7]. This lemma allows us to quantify the decorrelation in the usual Hölder’s inequality when we increase the oscillation of one function.

Lemma B.1

Let \(p \in [1,\infty ]\) and \(a,f :{\mathbb {T}}^d \rightarrow {\mathbb {R}}\) be smooth functions. Then for any \(\sigma \in {\mathbb {N}}\) ,

$$\begin{aligned} \Big | \Vert a f(\sigma \cdot ) \Vert _{p } - \Vert a \Vert _{p} \Vert f \Vert _{p } \Big |\lesssim \sigma ^{-\frac{1}{p}} \Vert a\Vert _{C^1} \Vert f \Vert _{ p }. \end{aligned}$$
(B.1)

The proof is based on the interplay between the Poincare’s inequality and the fast oscillation of \(f (\sigma \cdot )\) and can be found in [60, Lemma 2.1].

1.2 Tensor-valued antidivergence \({\mathcal {R}}\)

For any \(f\in C^\infty ({\mathbb {T}}^d)\), there exists a \(v \in C^\infty _0({\mathbb {T}}^d)\) such that

And we denote v by \(\Delta ^{-1}f\). Note that if \(f\in C^\infty _0({\mathbb {T}}^d) \), then by rescaling we have

$$\begin{aligned} \Delta ^{-1} \big ( f(\sigma \cdot ) \big ) = \sigma ^{-2} v(\sigma \cdot ) \quad \text { for }\sigma \in {\mathbb {N}}. \end{aligned}$$

We recall the following antidivergence operator \({\mathcal {R}}\) introduced in [26].

Definition B.2

\({\mathcal {R}} : C^\infty ({\mathbb {T}}^d ,{\mathbb {R}}^d) \rightarrow C^\infty ({\mathbb {T}}^d, {\mathcal {S}}^{d \times d }_0)\) is defined by

$$\begin{aligned} ({\mathcal {R}} v )_{ij} = {\mathcal {R}}_{ijk} v_k \end{aligned}$$
(B.2)

where

$$\begin{aligned} {\mathcal {R}}_{ijk} = \frac{2-d}{d-1}\Delta ^{-2} \partial _i \partial _j \partial _k - \frac{1}{d-1}\Delta ^{-1} \partial _k \delta _{ij} + \Delta ^{-1} \partial _i \delta _{jk} + \Delta ^{-1} \partial _j \delta _{ik} . \end{aligned}$$

It is clear that \({\mathcal {R}} \) is well-defined since \({\mathcal {R}}_{ijk}\) is symmetric in ij and taking the trace gives

$$\begin{aligned} {{\,\mathrm{Tr}\,}}{\mathcal {R}} v&= \frac{2-d}{ d -1} \Delta ^{-1} \partial _k v_k + \frac{-d}{d-1} \Delta ^{-1} \partial _k v_k + \Delta ^{-1} \partial _k v_k + \Delta ^{-1} \partial _k v_k\\&=(\frac{2-d}{ d -1} + \frac{-d}{d-1}+ 2) \Delta ^{-1} \partial _k v_k=0. \end{aligned}$$

By a direct computation, one can also show that

and

$$\begin{aligned} {\mathcal {R}} \Delta v = \nabla v + \nabla v ^T \quad \text {for any divergence-free }v \in C^\infty ({\mathbb {T}}^d ,{\mathbb {R}}^d). \end{aligned}$$
(B.3)

We can show that \( {\mathcal {R}}\) is bounded on \(L^{p} ({\mathbb {T}}^d) \) for any \(1\le p \le \infty \).

Theorem B.3

Let \(1 \le p \le \infty \). For any vector field \(f \in C^\infty ({\mathbb {T}}^d,{\mathbb {R}}^d)\), there holds

$$\begin{aligned} \Vert {\mathcal {R}} f \Vert _{L^{p}({\mathbb {T}}^d )} \lesssim \Vert f \Vert _{L^{p}({\mathbb {T}}^d )}. \end{aligned}$$

In particular, if \(f \in C^\infty _0({\mathbb {T}}^d,{\mathbb {R}}^d)\), then

$$\begin{aligned} \Vert {\mathcal {R}} f(\sigma \cdot ) \Vert _{L^{p}({\mathbb {T}}^d )} \lesssim \sigma ^{-1}\Vert f \Vert _{L^{p}({\mathbb {T}}^d )} \quad \text {for any }\sigma \in {\mathbb {N}}. \end{aligned}$$

Proof

Once the first bound is established, the second bound follows from the definition of \( {\mathcal {R}}\). It suffices to only consider f with zero mean since \({\mathcal {R}} (C) =0\) for any constant C. Then we only need to show that the operator

$$\begin{aligned} \Delta ^{-1} \Delta ^{-1} \partial _i \partial _j \partial _k \end{aligned}$$

is bounded on \(L^p({\mathbb {T}}^d)\) for \(1\le p \le \infty \) since the argument applies also to \(\Delta ^{-1} \partial _i \).

When \(1< p < \infty \), this follows from the boundedness of the Riesz transforms and the Poincare inequality.

When \(p= \infty \), the Sobolev embedding \(W^{1,d+1}({\mathbb {T}}^d) \hookrightarrow L^{\infty }({\mathbb {T}}^d )\) implies that

$$\begin{aligned} \Vert \Delta ^{-1} \Delta ^{-1} \partial _i \partial _j \partial _k f \Vert _{L^\infty ({\mathbb {T}}^d) }\lesssim & {} \Vert \Delta ^{-1} \Delta ^{-1} \partial _i \partial _j \partial _k f \Vert _{ W^{1,d+1} ({\mathbb {T}}^d) } \\\lesssim & {} \Vert f \Vert _{L^{d + 1}({\mathbb {T}}^d) }\le \Vert f \Vert _{L^{\infty } ({\mathbb {T}}^d) } \end{aligned}$$

where we have used the boundedness of the Riesz transforms once again.

When \(p= 1\), one can use a duality approach and use the boundedness in \(L^\infty \) since integrating by parts yields

$$\begin{aligned} \langle \Delta ^{-1} \Delta ^{-1} \partial _i \partial _j \partial _k f ,\varphi \rangle = -\langle f ,\Delta ^{-1} \Delta ^{-1} \partial _i \partial _j \partial _k \varphi \rangle \quad \text { if } \varphi \in C^\infty _0({\mathbb {T}}^d). \end{aligned}$$

\(\square \)

1.3 Bilinear antidivergence \({\mathcal {B}}\)

We can also introduce the bilinear version \({\mathcal {B}} : C^\infty ({\mathbb {T}}^d, {\mathbb {R}}^d) \times C^\infty ({\mathbb {T}}^d ,{\mathbb {R}}^{d\times d} ) \rightarrow C^\infty ({\mathbb {T}}^d, {\mathcal {S}}^{d \times d}_0) \) of \({\mathcal {R}}\). This bilinear antidivergence \({\mathcal {B}}\) allows us to gain derivative when the later argument has zero mean and a small period.

Let

$$\begin{aligned} ( {\mathcal {B}}( v,A ))_{i j} = v_l {\mathcal {R}}_{ijk}A_{lk} - {\mathcal {R}}( \partial _i v_l {\mathcal {R}}_{ijk}A_{lk} ) \end{aligned}$$

or by a slight abuse of notations

$$\begin{aligned} {\mathcal {B}}( v,A ) = v {\mathcal {R}} A - {\mathcal {R}}( \nabla v {\mathcal {R}} A ) . \end{aligned}$$

Theorem B.4

Let \(1 \le p \le \infty \). For any \(v \in C^\infty ({\mathbb {T}}^d,{\mathbb {R}}^d)\) and \(A \in C^\infty _0({\mathbb {T}}^d,{\mathbb {R}}^{d\times d})\),

(B.4)

and

$$\begin{aligned} \Vert {\mathcal {B}} (v ,A) \Vert _{L^{p}({\mathbb {T}}^d )} \lesssim \Vert v \Vert _{C^{1}({\mathbb {T}}^d )} \Vert {\mathcal {R}} A \Vert _{L^{p}({\mathbb {T}}^d )}. \end{aligned}$$

Proof

A direct compuation gives

where we have used the fact that A has zero mean and \({\mathcal {R}}\) is symmetric.

Integrating by parts, we have

which implies that

The second estimate follows immediately from the definition of \({\mathcal {B}}\) and Theorem B.3. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cheskidov, A., Luo, X. Sharp nonuniqueness for the Navier–Stokes equations. Invent. math. 229, 987–1054 (2022). https://doi.org/10.1007/s00222-022-01116-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00222-022-01116-x

Navigation