Skip to main content
Log in

Potential Singularity of the 3D Euler Equations in the Interior Domain

  • Published:
Foundations of Computational Mathematics Aims and scope Submit manuscript

Abstract

Whether the 3D incompressible Euler equations can develop a finite time singularity from smooth initial data is one of the most challenging problems in nonlinear PDEs. In this paper, we present some new numerical evidence that the 3D axisymmetric incompressible Euler equations with smooth initial data of finite energy develop a potential finite time singularity at the origin. This potential singularity is different from the blow-up scenario revealed by Luo and Hou (111:12968–12973, 2014) and (12:1722–1776, 2014), which occurs on the boundary. Our initial condition has a simple form and shares several attractive features of a more sophisticated initial condition constructed by Hou and Huang in (arXiv:2102.06663, 2021) and (435:133257, 2022). One important difference between these two blow-up scenarios is that the solution for our initial data has a one-scale structure instead of a two-scale structure reported in Hou and Huang (arXiv:2102.06663, 2021) and (435:133257, 2022). More importantly, the solution seems to develop nearly self-similar scaling properties that are compatible with those of the 3D Navier–Stokes equations. We will present numerical evidence that the 3D Euler equations seem to develop a potential finite time singularity. Moreover, the nearly self-similar profile seems to be very stable to the small perturbation of the initial data.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24
Fig. 25
Fig. 26
Fig. 27
Fig. 28
Fig. 29

Similar content being viewed by others

References

  1. J. Beale, T. Kato, and A. Majda. 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 

  2. O. N. Boratav and R. B. Pelz. Direct numerical simulation of transition to turbulence from a high-symmetry initial condition. Phys. Fluids, 6:2757–2784, 1994.

    Article  MATH  Google Scholar 

  3. M. Brenner, S. Hormoz, and A. Pumir. Potential singularity mechanism for the Euler equations. Phys. Rev. Fluids, 1:084503, 2016.

    Article  Google Scholar 

  4. J. Chen and T. Y. Hou. Finite time blowup of \(2\)D Boussinesq and \(3\)D Euler equations with \({C}^{1,\alpha }\) velocity and boundary. CMP, 383(3):1559–1667, 2021.

    MATH  Google Scholar 

  5. J. Chen and T. Y. Hou. On stability and instability of \(c^{1,\alpha }\) singular solutions of the \(3\)D Euler and \(2\)D Boussinesq equations. arXiv:2206.01296 [math.AP], 2022.

  6. J. Chen, T. Y. Hou, and D. Huang. Asymptotically self-similar blowup of the Hou-Luo model for the \(3\)D Euler equations. arXiv:2106.05422 [math.AP], 2021.

  7. J. Chen, T. Y. Hou, and D. Huang. On the finite time blowup of the De Gregorio model for the \(3\)D Euler equation. CPAM, https://doi.org/10.1002/cpa.21991, 2021.

    Article  MATH  Google Scholar 

  8. K. Choi, T. Y. Hou, A. Kiselev, G. Luo, V. Sverak, and Y. Yao. On the finite-time blowup of a \(1\)D model for the \(3\)D axisymmetric Euler equations. CPAM, 70(11):2218–2243, 2017.

    MATH  Google Scholar 

  9. K. Choi, A. Kiselev, , and Y. Yao. Finite time blow up for a \(1\)D model of \(2\)D Boussinesq system. CMP, 334(3):1667–1679, 2015.

    MathSciNet  MATH  Google Scholar 

  10. C. Collot, T. Ghoul, and N. Masmoudi. Singularity formation for Burgers equation with transverse viscosity. arXiv:1803.07826v2 [math.AP], 2020.

  11. P. Constantin, C. Fefferman, and A. Majda. Geometric constraints on potentially singular solutions for the \(3\)-D Euler equations. Commun. PDEs, 21:559–571, 1996.

    Article  MathSciNet  MATH  Google Scholar 

  12. J. Deng, T. Y. Hou, and X. Yu. Geometric properties and non-blowup of \(3\)D incompressible Euler flow. Commun. PDEs, 30:225–243, 2005.

    Article  MATH  Google Scholar 

  13. W. E and C.-W. Shu. Small-scale structures in Boussinesq convection. Phys. Fluids, 6:49–58, 1994.

  14. T. M. Elgindi. Finite-time singularity formation for \({C}^{1,\alpha }\) solutions to the incompressible euler equations on \({R}^3\). Annals of Mathematics, 194(3):647–727, 2021.

    Article  MathSciNet  MATH  Google Scholar 

  15. T. M. Elgindi, T. Ghoul, and N. Masmoudi. On the stability of self-similar blow-up for \({C}^{1,\alpha }\) solutions to the incompressible Euler equations on \({R}^3\). arXiv:1910.14071, 2019.

  16. T. M. Elgindi and I. J. Jeong. The incompressible Euler equations under octahedral symmetry: Singularity formation in a fundamental domain. Adv. Math., 393:10891, 2021.

    Article  MathSciNet  MATH  Google Scholar 

  17. J. Gibbon. The three-dimensional Euler equations: Where do we stand? Physica D, 237:1894–1904, 2008.

    Article  MathSciNet  MATH  Google Scholar 

  18. R. Grauer and T. C. Sideris. Numerical computation of \(3\)D incompressible ideal fluids with swirl. Phys. Rev. Lett., 67:3511–3514, 1991.

    Article  Google Scholar 

  19. T. Y. Hou. The potentially singular behavior of the \(3\)D Navier–Stokes equations. arXiv:2107.06509 [physics.flu-dyn], 2021.

  20. T. Y. Hou and D. Huang. Potential singularity formation of \(3\)D axisymmetric Euler equations with degenerate variable viscosity coefficients. arXiv:2102.06663, 2021.

  21. T. Y. Hou and D. Huang. A potential two-scale traveling wave asingularity for \(3\)D incompressible Euler equations. Physica D, 435:133257, 2022.

    Article  MATH  Google Scholar 

  22. T. Y. Hou and C. Li. Dynamic stability of the three-dimensional axisymmetric Navier–Stokes equations with swirl. CPAM, 61(5):661–697, 2008.

    MathSciNet  MATH  Google Scholar 

  23. T. Y. Hou and R. Li. Dynamic depletion of vortex stretching and non-blowup of the 3-D incompressible Euler equations. J. Nonlinear Sci., 16:639–664, 2006.

    Article  MathSciNet  MATH  Google Scholar 

  24. T. Y. Hou and R. Li. Blowup or no blowup? the interplay between theory and numerics. Physica D., 237:1937–1944, 2008.

    Article  MathSciNet  MATH  Google Scholar 

  25. R. M. Kerr. Evidence for a singularity of the three-dimensional incompressible Euler equations. Phys. Fluids A, 5:1725–1746, 1993.

    Article  MathSciNet  MATH  Google Scholar 

  26. A. Kiselev. Small scales and singularity formation in fluid dynamics. In Proceedings of the International Congress of Mathematicians, volume 3, 2018.

  27. A. Kiselev, L. Ryzhik, Y. Yao, and A. Zlatos. Finite time singularity for the modified SQG patch equation. Ann. Math., 184:909–948, 2016.

    Article  MathSciNet  MATH  Google Scholar 

  28. A. Kiselev and V. Sverak. Small scale creation for solutions of the incompressible two dimensional Euler equation. Annals of Mathematics, 180:1205–1220, 2014.

    Article  MathSciNet  MATH  Google Scholar 

  29. L. Lafleche, A. F. Vasseur, and M. Vishik. Instability for axisymmetric blow-up solutions to incompressible Euler equations. J. Math. Pures Appl., 155:140–154, 2021.

    Article  MathSciNet  MATH  Google Scholar 

  30. J. Liu and W. Wang. Convergence analysis of the energy and helicity preserving scheme for axisymmetric flows. SINUM, 44(6):2456–2480, 2006.

    Article  MathSciNet  MATH  Google Scholar 

  31. G. Luo and T. Y. Hou. Potentially singular solutions of the \(3\)D axisymmetric Euler equations. Proceedings of the National Academy of Sciences, 111(36):12968–12973, 2014.

    Article  MATH  Google Scholar 

  32. G. Luo and T. Y. Hou. Toward the finite-time blowup of the \(3\)D axisymmetric Euler equations: a numerical investigation. Multiscale Modeling & Simulation, 12(4):1722–1776, 2014.

    Article  MathSciNet  MATH  Google Scholar 

  33. A. Majda and A. Bertozzi. Vorticity and incompressible flow, volume 27. Cambridge University Press, 2002.

  34. D. McLaughlin, G. Papanicolaou, C. Sulem, and P. Sulem. Focusing singularity of the cubic schrödinger equation. Physical Review A, 34(2):1200, 1986.

    Article  Google Scholar 

  35. A. F. Vasseur and M. Vishik. Blow-up solutions to \(3\)D Euler are hydrodynamically unstable. CMP, 378:557–568, 2020.

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The research was in part supported by NSF Grants DMS-1907977, DMS-1912654, DMS-2205590, and the Choi Family Gift Fund. I would like to thank Dr. De Huang for very helpful discussions regarding the design of the adaptive mesh strategy. I would also like to thank Professor Vladimir Sverak, Jiajie Chen, Dr. De Huang, and the referees for their very constructive comments and suggestions, which significantly improves the quality of this paper. Finally, I have benefited a lot from the AIM SQarRE “Towards a 3D Euler singularity”.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thomas Y. Hou.

Additional information

Publisher's Note

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

Appendix A: Construction of the Adaptive Mesh

Appendix A: Construction of the Adaptive Mesh

In this appendix, we describe our adaptive mesh strategy to study the singularity formation near the origin \((r,z) = (0,0)\). We will use the method described in Appendix B of [20] to construct our adaptive mesh maps \(r=r(\rho )\) and \(z=z(\eta )\). We will discretize the equations in the transformed variables \((\rho ,\eta )\) with \(n_1\) grid points along the z direction and \(n_2\) grid points along the r-direction.

The adaptive mesh strategy described in [20] was inspired by the adaptive mesh strategy introduced in [32]. On the other hand, the adaptive mesh strategy developed in [32] is simpler due to the fact that the singularity is located at a fixed stagnation point on the boundary \((r,z)=(1,0)\), \(\omega _1 >0\) for \(z>0\) and has a bell-shaped structure near the singularity. In our case, we have a traveling wave solution that approaches the origin with \(\omega _1\) changing sign in the most singular region. The singular region has a more complicated shape since the potentially singular solution produces a strong shearing flow traveling downstream. Thus, we need to introduce a moving frame and design our adaptive mesh map \(r(\rho )\) and \(z(\eta )\) to resolve the solution in different regions. Our adaptive mesh strategy does not require that the solution has a bell-shaped structure in the most singular region. To construct \(r(\rho )\), we use the distance dr between the location at which \(u_1\) achieves its maximum and the location at which \(u_{1r}\) achieves its maximum to define the boundary \(r_i\) for different singular regions (phases). Similarly, we construct \(z(\eta )\) by using the distance dz between the location at which \(\omega _1\) achieves its maximum and the location at which \(\omega _{1z}\) achieves its maximum to define the boundary \(z_j\) for different phases.

1.1 A.1. The Adaptive (Moving) Mesh Algorithm

To effectively and accurately compute the potential blow-up, we have designed a special meshing strategy that is dynamically adaptive to the singular structure of the solution. The adaptive mesh covering the half-period computational domain \(\mathcal {D}_1 = \{(r,z):0\le r\le 1,0\le z\le 1/2\}\) is characterized by a pair of analytic mesh mapping functions

$$\begin{aligned}r = r(\rho ),\quad \rho \in [0,1];\quad z = z(\eta ),\quad \eta \in [0,1].\end{aligned}$$

These mesh mapping functions are both monotonically increasing and infinitely differentiable on [0, 1], and satisfy \(r(0) = 0,\; r(1) = 1,\; z(0) = 0,\; z(1) = 1/2.\) In particular, we construct these mapping functions by carefully designing their Jacobians/densities \(r_\rho = r'(\rho ),\quad z_\eta = z'(\eta )\) using analytic functions that are even functions at 0. The even symmetries ensure that the resulting mesh can be smoothly extended to the full-period cylinder \( \{(r,z):0\le r\le 1,-1/2\le z\le 1/2\}\). The density functions contain a small number of parameters, which are dynamically adjusted to the solution. Once the mesh mapping functions are constructed, the computational domain is covered with a tensor-product mesh:

$$\begin{aligned} \mathcal {G} = \{(r_i,z_j): 0\le i\le n_2,\ 0\le j\le n_1\}, \end{aligned}$$
(A.1)

where \(r_i^h = r(ih_\rho ),\; h_\rho = 1/n_2;\; z_j^h = z(jh_\eta ),\; h_\eta = 1/n_1.\) The precise definition and construction of the mesh mapping functions are described in Appendix B of [20].

Figure 30 gives an example of the densities \(r_\rho ,z_\eta \) (in log scale) we use in the computation. We design the densities \(r_\rho ,z_\eta \) to have three phases:

  • Phase 1 covers the inner profile of the smaller scale near the sharp front;

  • Phase 2 covers the outer profile of the larger scale of the solution;

  • Phase 3 covers the (far-field) solution away from the symmetry axis \(r=0\).

We add a phase 0 in the density \(r_\rho \) to cover the region near \(r=0\) and also add a phase 0 in the density \(z_\eta \) to cover the region near \(z=0\) in the late stage. In our computation, the number (percentage) of mesh points in each phase are fixed, but the physical location of each phase will change in time, dynamically adaptive to the structure of the solution. Between every two neighboring phases, there is also a smooth transition region that occupies a fixed percentage of mesh points.

Fig. 30
figure 30

The mapping densities \(r_\rho \) (left) and \(z_\eta \) (right) with phase numbers labeled. This figure is for illustration only. The parameters do not reflect the adaptive mesh used in our computation

1.2 A.2. Adaptive Mesh for the 3D Euler Equations

We use three different adaptive mesh strategies for three different time periods. The first time period corresponds to the time interval between \(t=0\) and \(t_1=0.002231338\) with \(\Vert \omega (t_1)\Vert _{L^\infty }/\Vert \omega (0)\Vert _{L^\infty } \approx 46.54325\) for the \(1536\times 1536\) grid and the number of time steps equal to 45000. The second time period corresponds to the time interval between \(t_1=0.002231338\) and \(t_2 = 0.002264353\) with \(\Vert \omega (t_2)\Vert _{L^\infty }/\Vert \omega (0)\Vert _{L^\infty } \approx 295.39986\) for the \(1536\times 1536\) grid and the number of time steps equal to 60000. The third time period is for \( t \ge t_2\).

In the first time period, since we use a very smooth initial condition whose support covers the whole domain, we use the following parameters \(r_1=0.001,\; r_2=0.05, \; r_3=0.2\) and \(s_{\rho _1}=0.001\), \(s_{\rho _2}=0.5\), \(s_{\rho _3}=0.85\) to construct the mapping \(r=r(\rho )\) using a four-phase map. Similarly, we use the following parameters \(z_1=0.1,\; z_2=0.25\) and \(s_{\eta _1}=0.5\), \(s_{\eta _2}=0.85\) to construct the mapping \(z=z(\eta )\) using a three-phase map. We then update the mesh \(z=z(\eta )\) dynamically using \(z_1=2 z(I)\) and \(z_2=10 z(I)\) with \(s_{\eta _1}=0.6\), \(s_{\eta _2}=0.9\) when \(I < 0.25n_1\), but keep \(r=r(\rho )\) unchanged during this early stage. Here I is the grid point index along the z-direction at which \(\omega _1\) achieves its maximum.

In the second time period, we use the following parameters \(s_{\rho _1}=0.05\), \(s_{\rho _2}=0.6\), \(s_{\rho _3}=0.9\), \(r_2 = r(J) + 2dr\), \(r_1=\max ((s_{\rho _1}/s_{\rho _2})r_2,r(J_r) - 5dr)\), and \(r_3 = \max (3r(J),(r_2 - r_1)(s_{\rho _3}-s_{\rho _2})/(s_{\rho _2}-s_{\rho _1}) + r_2)\), where J is the grid index at which \(u_1\) achieves its maximum along the r-direction, \(J_r\) is the grid index at which \(u_{1,r}\) achieves its maximum along the r-direction, and \(dr = r(J) - r(J_r)\). We update the mapping \(r(\rho )\) dynamically when \(J_r < 0.2n_2\). The adaptive mesh map for \(z(\eta )\) in the second time period remains the same as in the first time period.

In the third time period, we need to allocate more grid points to resolve the sharp front. We use the following parameters \(s_{\rho _1}=0.05\), \(s_{\rho _2}=0.65\), \(s_{\rho _3}=0.9\), \(r_2 = r(J) + 10dr\), \(r_1=\max ((s_{\rho _1}/s_{\rho _2})r_2,r(J_r) - 3dr)\), and \(r_3 = \max (2.3r(J),(r_2 - r_1)(s_{\rho _3}-s_{\rho _2})/(s_{\rho _2}-s_{\rho _1}) + r_2)\). To construct the mesh map \(z(\eta \), we use the following parameters \(s_{\eta _1}=0.05\), \(s_{\eta _2}=0.65\), \(s_{\eta _3}=0.9\), \(z_2 = z(I_w) + 2dz\), \(z_1=\max ((s_{\eta _1}/s_{\eta _2})z_2,z(I_{wz}) - 16dz)\), and \(z_3 = \max (2.3z(I_w),(z_2 - z_1)(s_{\eta _3}-s_{\eta _2})/(s_{\eta _2}-s_{\eta _1}) + z_2)\), where \(I_w\) is the grid index at which \(\omega _1\) achieves its maximum along the z-direction, \(I_{wz}\) is the grid index at which \(\omega _{1,z}\) achieves its maximum along the z-direction, and \(dz = z(I_w) - r(I_{wz})\). We will update \(r(\rho )\) dynamically when \(J_r < 0.2 n_2\) and update \(z(\eta )\) when \(I_z < 0.23 n_1\).

The three time periods for different grids will be defined similarly through the number of time steps by using a linear scaling relationship. For example, for the \(1024\times 1024\) grid, the first time period will be between \(t=0\) and the time step \(30000=45000 \cdot (1024/1536)\). The second time period will be between time steps 30000 and \(40000 = 60000\cdot (1024/1536)\). The third time period will be beyond time step 40000. The three time periods for other grids are defined similarly.

As we mentioned in the Introduction, we use a second-order Runge-Kutta method to discretize the Euler and Navier–Stokes equations in time with an adaptive time stepping strategy. We discretize the Euler and Navier–Stokes equations in the transformed domain \((\eta ,\rho )\) using a uniform mesh with \(h_1=1/n_1\) and \(h_2=1/n_2\). We choose our adaptive time step as follows:

$$\begin{aligned} k= & {} \min (k_1,k_2), \;\;k_1 = \min (\;0.2\min (h1,h_2)/\text{umax },\;10^{-3}\Vert u_1\Vert _{L^\infty }^{-1},\;2.5\cdot 10^{-7}),\\ k_2= & {} 0.1\min (\min (h_1z_\eta )^2/\nu ,\min (h_2r_\rho )^2/\nu ), \end{aligned}$$

where \(\text{ umax } = \max (\Vert u^r/r_\rho \Vert _{L^\infty },\Vert u^z/z_\eta \Vert _{L^\infty })\) is the maximum velocity in the transformed domain.

Rights and permissions

Springer Nature or its licensor 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

Hou, T.Y. Potential Singularity of the 3D Euler Equations in the Interior Domain. Found Comput Math 23, 2203–2249 (2023). https://doi.org/10.1007/s10208-022-09585-5

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10208-022-09585-5

Keywords

Mathematics Subject Classification

Navigation