Skip to main content
Log in

Non-conservative Solutions of the Euler-\(\alpha \) Equations

  • Published:
Journal of Mathematical Fluid Mechanics Aims and scope Submit manuscript

Abstract

The Euler-\(\alpha \) equations model the averaged motion of an ideal incompressible fluid when filtering over spatial scales smaller than \(\alpha \). We show that there exists \(\beta >1\) such that weak solutions to the two and three dimensional Euler-\(\alpha \) equations in the class \(C^0_t H^\beta _x\) are not unique and may not conserve the Hamiltonian of the system, thus demonstrating flexibility in this regularity class. The construction utilizes a Nash-style intermittent convex integration scheme. We also formulate an appropriate version of the Onsager conjecture for Euler-\(\alpha \), postulating that the threshold between rigidity and flexibility is the regularity class .

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. The pressure \(p_q\) is determined via the incompressibility of \(u_q\).

  2. The pressure corresponding to a stationary solution of the Euler-\(\alpha \) equations is not uniquely determined unless a mean-zero condition is imposed.

  3. In two dimensions, these error terms do not vanish but may be analyzed using Proposition 5.3 in the same way as the terms for which \(i\ne i'\).

References

  1. Arnold, V.: On the differential geometry of infinite dimensional lie groups and its applications to the hydrodynamics of perfect fluids. Ann. Fourier Inst. 16(1), 319–361 (1966)

    Article  MathSciNet  Google Scholar 

  2. Bahouri, H., Chemin, J.-Y., Danchin, R.: Fourier Analysis and Nonlinear Partial Differential Equations. Springer, Berlin (2011)

    Book  MATH  Google Scholar 

  3. Beekie, R., Buckmaster, T., Vicol, V.: Weak solutions of ideal MHD which do not conserve magnetic helicity. Ann. PDE 6(1), 40 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  4. Brué, E., Colombo, M., De Lellis, C.: Positive solutions of transport equations and classical nonuniqueness of characteristic curves. Arch. Ration. Mech. Anal. 240(2), 1055–1090 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  5. 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)

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

    Article  MathSciNet  MATH  Google Scholar 

  7. Buckmaster, T., De Lellis, C., Székelyhidi, L., Jr., Vicol, V.: Onsager’s conjecture for admissible weak solutions. Commun. Pure Appl. Math. 72(2), 229–274 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  8. 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 

  9. Buckmaster, T., Vicol, V.: Convex integration and phenomenologies in turbulence. EMS Surv. Math. Sci. 6(1), 173–263 (2019)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  11. 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)

  12. Busuioc, A.V., Iftimie, D., Lopes Filho, M.C., Nussenzveig Lopes, H.J.: Uniform time of existence for the alpha Euler equations. J. Funct. Anal. 271(5), 1341–1375 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  13. Busuioc, A.V., Iftimie, D.: Weak solutions for the \(\alpha \)-Euler equations and convergence to Euler. Nonlinearity 30(12), 4534–4557 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Busuioc, A.V., Iftimie, D., Lopes Filho, M.C., Nussenzveig Lopes, H.J.: Incompressible euler as a limit of complex fluid models with navier boundary conditions. J. Differ. Equ. 252(1), 624–640 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Busuioc, V.: On second grade fluids with vanishing viscosity. C. R. Acad. Sci. Paris Sér. I Math. 328(12), 1241–1246 (1999)

  16. Chen, S., Foias, C., Holm, D., Olson, E., Titi, E., Wynne, S.: Camassa–Holm equations as a closure model for turbulent channel and pipe flow. Phys. Rev. Lett. 81 (1998)

  17. Chen, S., Holm, D.D., Margolin, L.G., Zhang, R.: Direct numerical simulations of the Navier-Stokes alpha model. In: Predictability: quantifying uncertainty in models of complex phenomena, Los Alamos, NM, pp. 66–83 (1998)

  18. Cheskidov, A., Luo, X.: Sharp nonuniqueness for the Navier–Stokes equations. arXiv:2009.06596 (2020)

  19. Cheskidov, A., Luo, X.: Nonuniqueness of weak solutions for the transport equation at critical space regularity. Ann. PDE 7(1), 45 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  20. Constantin, P., Weinan, E., 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  ADS  MathSciNet  MATH  Google Scholar 

  21. Dai, M.: Nonunique weak solutions in Leray-Hopf class for the three-dimensional Hall-MHD system. SIAM J. Math, Anal (2021)

    Book  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  23. De Lellis, C., Székelyhidi, L., Jr.: On turbulence and geometry: from Nash to Onsager. Notices Am. Math. Soc. 66(5), 677–685 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  24. Dunn, J.E., Fosdick, R.L.: Thermodynamics, stability, and boundedness of fluids of complexity 2 and fluids of second grade. Arch. Ration. Mech. Anal. 56, 191–252 (1974)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  26. Holm, D., Marsden, J., Ratiu, T.: Euler-Poincaré models of ideal fluids with nonlinear dispersion. Phys. Rev. Lett. 80, 4173–4176 (1998)

    Article  ADS  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  28. Hou, T.Y., Li, C.: On global well-posedness of the Lagrangian averaged Euler equations. SIAM J. Math. Anal. 38(3), 782–794 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  29. Isett, P., Vicol, V.: Hölder continuous solutions of active scalar equations. Ann. PDE 1(1), 1–77 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  30. Isett, P.: Hölder continuous Euler flows in three dimensions with compact support in time. Annals of Mathematics Studies, vol. 196, Princeton University Press, Princeton, NJ (2017)

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

    Article  MathSciNet  MATH  Google Scholar 

  32. Lellis, C.D., Kwon, H.: On non-uniqueness of Hölder continuous globally dissipative Euler flows (2020)

  33. Linshiz, J.S., Titi, E.S.: On the convergence rate of the Euler-\(\alpha \), an inviscid second-grade complex fluid, model to the Euler equations. J. Stat. Phys. 138(1–3), 305–332 (2010)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  34. Lopes Filho, M.C., Nussenzveig Lopes, H.J., Titi, E.S., Zang, A.: Convergence of the 2D Euler-\(\alpha \) to Euler equations in the Dirichlet case: indifference to boundary layers. Phys. D 292(293), 51–61 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  35. Marsden, J.E., Ratiu, T.S., Shkoller, S.: The geometry and analysis of the averaged Euler equations and a new diffeomorphism group. Geom. Funct. Anal. 10(3), 582–599 (2000)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  37. Mohseni, K., Kosović, B., Shkoller, S., Marsden, J.E.: Numerical simulations of the Lagrangian averaged Navier-Stokes equations for homogeneous isotropic turbulence. Phys. Fluids 15(2), 524–544 (2003). https://doi.org/10.1063/1.1533069

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. Novack, M.: Nonuniqueness of weak solutions to the 3-dimensional quasi-geostrophic equations. SIAM J. Math. Anal. 52(4), 3301–3349 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  39. Onsager, L.: Statistical hydrodynamics. Nuovo Cimento 9(6)2 (Convegno Internazionale di Meccanica Statistica), 279–287 (1949)

  40. Shkoller, S.: Geometry and curvature of diffeomorphism groups with \(H^1\) metric and mean hydrodynamics. J. Funct. Anal. 160(1), 337–365 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  41. Shkoller, S.: Analysis on groups of diffeomorphisms of manifolds with boundary and the averaged motion of a fluid. J. Differ. Geom. 55(1), 145–191 (2000)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors thank Vlad Vicol and Steve Shkoller for stimulating conversations. RB was partially supported by the NSF under Grant No. DMS-1911413 and No. DMS-1954357. MN was partially supported by the NSF under Grant No. DMS-1928930 while participating in a program hosted by the Mathematical Sciences Research Institute during the spring 2021 semester, and by the NSF under Grant No. DMS-1926686 while a member at the Institute for Advanced Study.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Rajendra Beekie.

Ethics declarations

Conflict of interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.

Additional information

Communicated by S. Shkoller.

Publisher's Note

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

Appendices

Proof of Lemma 1.5

In this section we provide a proof of Lemma 1.5.

Proof

We follow the proof from [20]. Mollifying (1.8) in space with a standard radial, compactly supported mollifier \(\varphi _{\varepsilon }\), assuming differentiability in time, and integrating in space against \(u_i^{\varepsilon }=u *\varphi _\varepsilon \) gives

$$\begin{aligned}&\frac{1}{2} \frac{d}{dt} \left( \Vert u_i^{\varepsilon } \Vert _{L^2}^2 + \alpha ^2 \Vert \nabla u_i^{\varepsilon }\Vert _{L^2}^2 \right) \nonumber \\&= \int _{\mathbb {T}^n } (u_j u_i)^{\varepsilon } \partial _j u_i^{\varepsilon } + \alpha ^2 (u_j \partial _k u_i )^{\varepsilon } \partial _j \partial _k u_i^{\varepsilon } + \alpha ^2 (\partial _k u_j \partial _k u_i)^{\varepsilon } \partial _ju_i^{\varepsilon } - \alpha ^2 (\partial _k u_j \partial _i u_j)^{\varepsilon } \partial _k u_i^{\varepsilon } \, dx \, \end{aligned}$$
(A.1)

where we have used Remark 1.3.

We wish to show that the right-hand side vanishes when \(\varepsilon \) is sent to 0. We recall some standard estimates for mean-zero \(f \in B_{3, \infty }^{s}\) where \({1<s<2}\); see for example [2, 20].

$$\begin{aligned}&\Vert \nabla f(\cdot ) - \nabla f(\cdot - y)\Vert _{L^3} \lesssim |y|^{s -1} \Vert \nabla f \Vert _{B_{3 , \infty }^{s - 1}} \end{aligned}$$
(A.2a)
$$\begin{aligned}&\Vert f - f^{\varepsilon } \Vert _{L^3} \lesssim \varepsilon \Vert f \Vert _{B_{3, \infty }^{s}} \end{aligned}$$
(A.2b)
$$\begin{aligned}&\Vert \nabla f - \nabla f^{\varepsilon } \Vert _{L^3} \lesssim \varepsilon ^{s -1} \Vert \nabla f \Vert _{B_{3, \infty }^{s-1}} \end{aligned}$$
(A.2c)
$$\begin{aligned}&\Vert \nabla ^2 f^{\varepsilon } \Vert _{L^3} \lesssim \varepsilon ^{s-2} \Vert \nabla f \Vert _{B_{3, \infty }^{s - 1}} \end{aligned}$$
(A.2d)
$$\begin{aligned}&{\left\| \nabla f \right\| _{L^3} \lesssim \left\| \nabla f \right\| _{B^{s-1}_{3,\infty }}} \end{aligned}$$
(A.2e)
$$\begin{aligned}&\left\| \nabla f \right\| _{B^{s-1}_{3,\infty }} \lesssim \left\| f \right\| _{B^{s}_{3,\infty }} \end{aligned}$$
(A.2f)
$$\begin{aligned}&{\left\| f^\varepsilon \right\| _{B^{s}_{3,\infty }}\lesssim \left\| f \right\| _{B^{s}_{3,\infty }}} \, . \end{aligned}$$
(A.2g)

Finally, recall the double commutator identity from [20]:

$$\begin{aligned} (fg)^{\varepsilon } = f^{\varepsilon }g^{\varepsilon } - (f - f^{\varepsilon })(g -g^{\varepsilon }) + r_{\varepsilon }(f,g) \, , \end{aligned}$$
(A.3)

where \(f^{\varepsilon }(x) = f * \varphi _{\varepsilon }(x)\) and

$$\begin{aligned} r_{\varepsilon }(f,g) = \int _{\mathbb {T}^n} \varphi _{\varepsilon }(y) (f(x - y) - f(x))(g(x - y) - g(x)) \, dy \, . \end{aligned}$$
(A.4)

Then the right-hand side of (A.1) can be written as

$$\begin{aligned}&\int _{\mathbb {T}^n } (u_j u_i)^{\varepsilon } \partial _j u_i^{\varepsilon } + \alpha ^2 (u_j \partial _k u_i )^{\varepsilon } \partial _j \partial _k u_i^{\varepsilon } + \alpha ^2 (\partial _k u_j \partial _k u_i)^{\varepsilon } \partial _ju_i^{\varepsilon } - \alpha ^2 (\partial _k u_j \partial _i u_j)^{\varepsilon } \partial _k u_i^{\varepsilon } dx \nonumber \\&\quad = \int _{\mathbb {T}^n} u_j^{\varepsilon }u_i^{\varepsilon } \partial _j u_i^{\varepsilon } + \alpha ^2\left[ u_j^{\varepsilon }\partial _ku_i^{\varepsilon } \partial _j \partial _k u_i^{\varepsilon } + \partial _ku_j^{\varepsilon } \partial _k u_i^{\varepsilon } \partial _j u_i^{\varepsilon } - \partial _ku_j^{\varepsilon } \partial _i u_j^{\varepsilon } \partial _k u_i^{\varepsilon } \right] dx \nonumber \\&\qquad - \int _{\mathbb {T}^n} (u_j - u_j^{\varepsilon })(u_i - u_i^{\varepsilon }) \partial _j u_i^{\varepsilon } dx \nonumber \\&\qquad - \alpha ^2\int _{\mathbb {T}^n} \left[ ( u_j - u_j^{\varepsilon })(\partial _k u_i - \partial _k u_i^{ \varepsilon })\partial _j \partial _ku_i^{\varepsilon } + (\partial _k u_j - \partial _k u_j^{\varepsilon }) (\partial _k u_i - \partial _k u_i^{\varepsilon }) \partial _j u_i^{\varepsilon } \right. \nonumber \\&\qquad \left. - (\partial _k u_j - \partial _k u_j^{\varepsilon } )(\partial _i u_j - \partial _i u_j^{\varepsilon }) \partial _k u_i^{\varepsilon } \right] dx \nonumber \\&\qquad + \int _{\mathbb {T}^n} r_{\varepsilon }(u_j, u_i) \partial _j u_i^{\varepsilon } + \alpha ^2\left[ r_{\varepsilon } (u_j, \partial _ku_i ) \partial _j \partial _k u_i^{\varepsilon } + r_{\varepsilon } (\partial _ku_j ,\partial _k u_i) \partial _j u_i^{\varepsilon } - r_{\varepsilon }(\partial _ku_j , \partial _i u_j) \partial _k u_i^{\varepsilon } \right] dx \, . \end{aligned}$$
(A.5)

We have

$$\begin{aligned} \int _{\mathbb {T}^n} u_j^{\varepsilon } u_i^{\varepsilon } \partial _j u_i^{\varepsilon } dx = \int _{\mathbb {T}^n} u_j^{\varepsilon } \partial _k u_i^{\varepsilon } \partial _j \partial _k u_i^{\varepsilon } = 0 \end{aligned}$$

using integration by parts and that \(\textrm{div}u = 0\). Furthermore, the equality

$$\begin{aligned} \partial _k u_j^{\varepsilon } \partial _ku_i^{\varepsilon } \partial _j u_i^{\varepsilon } = \partial _k u_j^{\varepsilon } \partial _iu_j^{\varepsilon } \partial _k u_i^{\varepsilon } \end{aligned}$$

follows from switching the roles of i and j. Therefore the first four terms after the equal sign on the right-hand side of (A.5) vanish, and we can bound the remaining terms by

$$\begin{aligned}&\lesssim \Vert u - u^{\varepsilon }\Vert _{L^3}^2 \Vert \nabla u^{\varepsilon } \Vert _{L^3} + \Vert u - u^{\varepsilon }\Vert _{L^3} \Vert \nabla u - \nabla u^{\varepsilon } \Vert _{L^3} \Vert \nabla ^2 u^{\varepsilon } \Vert _{L^3} + \Vert \nabla u - \nabla u^{\varepsilon } \Vert _{L^3}^2 \Vert \nabla u^{\varepsilon } \Vert _{L^3} \\&\quad + \Vert r_{\varepsilon } (u, u) \Vert _{L^{\frac{3}{2}}} \Vert \nabla u^{\varepsilon } \Vert _{L^3} + \Vert r_{\varepsilon } (u, \nabla u)\Vert _{L^{\frac{3}{2}}} \Vert \nabla ^2 u^{\varepsilon }\Vert _{L^3} + \Vert r_{\varepsilon }( \nabla u, \nabla u) \Vert _{L^{\frac{3}{2}}} \Vert \nabla u^{\varepsilon } \Vert _{L^3 } \\&\lesssim \varepsilon ^{{2}} \Vert u \Vert _{B_{3, \infty }^{s} }^3 + \varepsilon ^{ {2(s-1)}} \Vert u \Vert _{B_{3 ,\infty }^{s}}^3. \end{aligned}$$

Thus, taking \(s > 1\) shows that the right-hand side of (A.1) converges to zero when \(\varepsilon \rightarrow 0\). Concluding the proof in the case of continuity in time may be done analogously as for the classical Euler equations, and we omit further details.

Proof of Lemma 4.1

In this section we provide a proof of Lemma 4.1.

Proof

We first construct \(\mathcal {K}_0\) and \(c_i^0\) by hand while allowing \(\left\| {\mathring{R}}\right\| _{0}\le 1\), afterwards constructing \(\mathcal {K}_n\) for \(1\le n \le N\) and choosing \(\varepsilon \). Consider a symmetric traceless matrix \({\mathring{R}}\), which without loss of generality may be written as

$$\begin{aligned} {\mathring{R}}= \begin{pmatrix} a &{} c &{} d\\ c &{} b &{} e\\ d &{} e &{} -a-b \end{pmatrix} \, . \end{aligned}$$
(B.1)

Notice that the set of such matrices is 5-dimensional, and combined with the condition on the sum of \((c_i^0)^2\) in (4.1) will require a set of at least six vectors. The extra three vectors will ensure that the coefficients are all positive. Ignoring for now the upper subscript 0 on the vectors \(k_i^0\in \mathcal {K}_0\), we set

$$\begin{aligned} k_1 = e_1 \, , \qquad k_2 = e_2 \, , \qquad k_3 = e_3 \, . \end{aligned}$$

For ease of notation, let us define

$$\begin{aligned} f(k_i) = 3k_i\otimes k_i - \text {Id} \, . \end{aligned}$$

Then it is clear that \(f(k_1)\), \(f(k_2)\), and \(f(k_3)\) are symmetric traceless matrices which only contain entries on the diagonal; specifically, we have

$$\begin{aligned} f(k_1) = \begin{pmatrix} 2 &{} 0 &{} 0\\ 0 &{} -1 &{} 0\\ 0 &{} 0 &{} -1 \end{pmatrix}\, , \qquad f(k_2) =\begin{pmatrix} -1 &{} 0 &{} 0\\ 0 &{} 2 &{} 0\\ 0 &{} 0 &{} -1 \end{pmatrix}\, , \qquad f(k_3) = \begin{pmatrix} -1 &{} 0 &{} 0\\ 0 &{} -1 &{} 0\\ 0 &{} 0 &{} 2 \end{pmatrix}\, . \qquad \end{aligned}$$
(B.2)

We shall use \(f(k_1)\) and \(f(k_2)\) to engender the entries of the matrix on the diagonal in (B.1), while we shall use a “balanced” sum of the form \(c(f(k_1)+f(k_2)+f(k_3))\) to ensure the second condition in (4.1). We further set

$$\begin{aligned} k_4&= \frac{3e_1+4e_2}{5}\, , \qquad k_5=\frac{3e_1+4e_3}{5}\, ,\qquad k_6=\frac{3e_2+4e_3}{5} \\ k_7&= \frac{3e_1-4e_2}{5}\, , \qquad k_8=\frac{3e_1-4e_3}{5}\, ,\qquad k_9=\frac{3e_2-4e_3}{5}\, . \end{aligned}$$

Then

$$\begin{aligned} f(k_4)= & {} \frac{1}{25}\begin{pmatrix} 2 &{} 36 &{} 0\\ 36 &{} 23 &{} 0\\ 0 &{} 0 &{} -25 \end{pmatrix}\, , \qquad f(k_5) = \frac{1}{25}\begin{pmatrix} 2 &{} 0 &{} 36\\ 0 &{} -25 &{} 0\\ 36 &{} 0 &{} 23 \end{pmatrix}\, , \qquad f(k_6) = \frac{1}{25}\begin{pmatrix} -25 &{} 0 &{} 0\\ 0 &{} 2 &{} 36\\ 0 &{} 36 &{} 23 \end{pmatrix}\nonumber \\ f(k_7)= & {} \frac{1}{25}\begin{pmatrix} 2 &{} -36 &{} 0\\ -36 &{} 23 &{} 0\\ 0 &{} 0 &{} -25 \end{pmatrix}\, , \quad f(k_8) = \frac{1}{25}\begin{pmatrix} 2 &{} 0 &{} -36\\ 0 &{} -25 &{} 0\\ -36 &{} 0 &{} 23 \end{pmatrix}\, , \quad f(k_9) = \frac{1}{25}\begin{pmatrix} -25 &{} 0 &{} 0\\ 0 &{} 2 &{} -36\\ 0 &{} -36 &{} 23 \end{pmatrix}. \nonumber \\ \end{aligned}$$
(B.3)

We shall use \(f(k_4)\), \(f(k_5)\), \(f(k_6)\), \(f(k_7)\), \(f(k_8)\), and \(f(k_9)\) to engender the entries of (B.1) off the diagonal.

It is simple to check that the set \(\{f(k_1),f(k_2),f(k_4),f(k_5),f(k_6)\}\) is a linearly independent set in the space of symmetric traceless matrices. Therefore, there exist smooth functions \(\{\tilde{c}_1, \tilde{c}_2, \tilde{c}_4, \tilde{c}_5, \tilde{c}_6\}\) given by the solutions of a linear system of equations such that for all symmetric traceless matrices \({\mathring{R}}\) satisfying \(\left\| {\mathring{R}}\right\| _0\le 1\),

$$\begin{aligned} \tilde{c}_1 f(k_1) + \tilde{c}_2 f(k_2) + \tilde{c}_4 f(k_4) + \tilde{c}_5 f(k_5) + \tilde{c}_6 f(k_6) = {\mathring{R}}\, . \end{aligned}$$
(B.4)

Unfortunately the functions \(\tilde{c}_i\) are not strictly positive. Let us define the auxiliary parameter

$$\begin{aligned} c_0 = \max _{\begin{array}{c} \left\| {\mathring{R}}\right\| _0\le 1, i=4,5,6 \end{array}} |\tilde{c}_i({\mathring{R}})| \, . \end{aligned}$$

Then from (B.4) and (B.3), we have that

$$\begin{aligned} \left( 1 -\delta _{ij} \right)&\bigg ( \tilde{c}_1 f(k_1) + \tilde{c}_2 f(k_2) + \tilde{c}_4 f(k_4) + \tilde{c}_5 f(k_5) + \tilde{c}_6 f(k_6) \\&\qquad + 2c_0 \left( f(k_4) + f(k_5) + f(k_6) + f(k_7) + f(k_8) + f(k_9) \right) \bigg )^{ij} = \left( 1 -\delta _{ij} \right) {\mathring{R}}^{ij}\, ; \end{aligned}$$

that is, off the diagonal, the matrix on the left hand side of the equation is equal to \({\mathring{R}}\). The advantage now is that the coefficients on \(f(k_i)\) for \(4\le i \le 9\) are all positive. In order to ensure equality on the diagonal, we may replace the coefficients \(\tilde{c}_1\) and \(\tilde{c}_2\) on \(f(k_1)\) and \(f(k_2)\) with new coefficients \(\dot{c}_1\) and \(\dot{c}_2\) such that

$$\begin{aligned}&\bigg (\dot{c}_1 f(k_1) + \dot{c}_2 f(k_2) + \tilde{c}_4 f(k_4) + \tilde{c}_5 f(k_5) + \tilde{c}_6 f(k_6) \nonumber \\&\qquad \qquad + 2c_0 \left( f(k_4) + f(k_5) + f(k_6) + f(k_7) + f(k_8) + f(k_9) \right) \bigg )^{ij} = {\mathring{R}}^{ij}\, \end{aligned}$$
(B.5)

for all \(1\le i,j\le 3\). Since the coefficients \(\dot{c}_1\) and \(\dot{c}_2\) are not necessarily positive, we set

$$\begin{aligned} \dot{c}_0 = \max _{\left\| {\mathring{R}} \right\| _0\le 1,i=1,2} \left| \dot{c}_i ({\mathring{R}}) \right| \, . \end{aligned}$$

Then from (B.2) and (B.5), we have that

$$\begin{aligned}&\bigg (\dot{c}_1 f(k_1) + \dot{c}_2 f(k_2) + 2\dot{c}_0 \left( f(k_1)+f(k_2)+f(k_3) \right) \\&\qquad + \tilde{c}_4 f(k_4) + \tilde{c}_5 f(k_5) + \tilde{c}_6 f(k_6) + 2c_0 \left( f(k_4) + f(k_5) + f(k_6) + f(k_7) + f(k_8) + f(k_9) \right) \bigg )^{ij} = {\mathring{R}}^{ij}\, \end{aligned}$$

Aggregating coefficients for each of the nine tensors on the left-hand side, we see that each is strictly positive. We still have to ensure the second condition in (4.1); however, this is easily achieved by replacing the constant \(\dot{c}_0\) with a non-negative, bounded function \(c_0({\mathring{R}}):B_1(0)\rightarrow [0,\infty )\) which depends on \({\mathring{R}}\) and imposes that the sum of the coefficients is equal for all \({\mathring{R}}\in B_1(0)\). Thus we have achieved everything in (4.1) for \(n=0\). Note that the value of \(\mathcal {C}_\textrm{sum}\) could in principle be computed explicitly and depends on the choice of the function \(c_0({\mathring{R}})\).

To construct \(\mathcal {K}_n\) for \(n=1\), choose a rational rotation matrix \(O_1\) with \(0 < \left\| O_1 - \text {Id} \right\| _0 \ll 1\), and set \(\mathcal {K}_1= O_1 \mathcal {K}_0\). Plugging the rotated vectors \(O_1 k_i^0\) into (4.1), we find that the effect of \(O_1\) is that

$$\begin{aligned} \sum _{i=1}^9 \left( c_i^0\right) ^2\left( {\mathring{R}}\right) \left( 3 k_i^0 \otimes k_i^0 - \text {Id} \right) = {\mathring{R}}\quad \rightarrow \quad \sum _{i=1}^9 \left( c_i^0\right) ^2\left( {\mathring{R}}\right) O_1\left( 3 k_i^0 \otimes k_i^0 - \text {Id} \right) O_1^T = O_1 {\mathring{R}}O_1^T \, . \end{aligned}$$

Using the equalities \(\text {tr}(AB)=\text {tr}(BA)\) and \(O^T=O^{-1}\), we have that \(O_1 {\mathring{R}}O_1^T\) is still traceless and symmetric. Define

$$\begin{aligned} c^1_i\left( {\mathring{R}}\right) = c^0_i\left( O_1^T {\mathring{R}}O_1 \right) \, \end{aligned}$$

for \(\left\| {\mathring{R}}\right\| \) sufficiently small so that \(c^0_i\left( O_1 {\mathring{R}}O_1^T\right) \) is well-defined and strictly positive for each i. Then this specifies a value \(\varepsilon _1\) such that for all \(\left\| {\mathring{R}}\right\| _0 \le \varepsilon _1\),

$$\begin{aligned} \sum _{i=1}^9 \left( c_i^1\right) ^2\left( {\mathring{R}}\right) \left( 3 k_i^1 \otimes k_i^1 - \text {Id} \right)&= \sum _{i=1}^9 \left( c^0_i\right) ^2 \left( O_1^T {\mathring{R}}O_1\right) O_1 \left( 3 k_i^0 \otimes k_i^0 - \text {Id} \right) O_1^T \\&= {\mathring{R}}\, , \end{aligned}$$

showing that the first equality in (4.1) is satisfied for \(n=1\). The second equality in (4.1) comes immediately from the construction of the \(c_i^0\)’s and \(c^1_i\)’s. In addition, \(\mathcal {K}_0\) and \(\mathcal {K}_1\) will be disjoint if \(0 < \left\| O_1 - \text {Id} \right\| _0\ll 1\). Iterating this procedure and taking the minimum value of \(\varepsilon _n\) concludes the proof.

\(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Beekie, R., Novack, M. Non-conservative Solutions of the Euler-\(\alpha \) Equations. J. Math. Fluid Mech. 25, 22 (2023). https://doi.org/10.1007/s00021-022-00757-5

Download citation

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00021-022-00757-5

Navigation