A locally anisotropic regularity criterion for the Navier–Stokes equation in terms of vorticity

By Evan Miller

Abstract

In this paper, we will prove a regularity criterion that guarantees solutions of the Navier–Stokes equation must remain smooth so long as the vorticity restricted to a plane remains bounded in the scale critical space , where the plane may vary in space and time as long as the gradient of the unit vector orthogonal to the plane remains bounded. This extends previous work by Chae and Choe that guaranteed that solutions of the Navier–Stokes equation must remain smooth as long as the vorticity restricted to a fixed plane remains bounded in a family of scale critical spaces. This regularity criterion also can be seen as interpolating between Chae and Choe’s regularity criterion in terms of two vorticity components and Beirão da Veiga and Berselli’s regularity criterion in terms of the gradient of vorticity direction. In physical terms, this regularity criterion is consistent with key aspects of the Kolmogorov theory of turbulence, because it requires that finite-time blowup for solutions of the Navier–Stokes equation must be fully three dimensional at all length scales.

1. Introduction

The Navier–Stokes equation is the fundamental equation of fluid mechanics. The incompressible Navier–Stokes equation is given by

where is determined entirely by by convolution with the Poisson kernel,

The Navier–Stokes equation is best viewed as an evolution equation on the space of divergence free vector fields rather than as a system of equations, and this is the vantage point we will adopt in this paper.

Two other fundamentally important objects for the study of the Navier–Stokes equation are the strain and the vorticity. The strain is the symmetric part of , and is given by

The vorticity is the curl of the velocity, . It is a vector representation of the anti-symmetric party of , with

where

Note that this representation implies that for all ,

The evolution equation for the strain is given by

and the evolution equation for vorticity is given by

Before we proceed, we should define a number of spaces. We will take to be the homogeneous Hilbert space with norm

We will note that for all , we have

We will take the inhomogeneous Hilbert space to be the space with norm

Finally we will take the mixed Lebesgue space to be the Banach space

We will note in particular that

In his foundational work on the Navier–Stokes equation, Leray proved the global existence of weak solutions to the Navier–Stokes equation satisfying an energy inequality for arbitrary initial data Reference 15. Such solutions, however, are not known to be either smooth or unique. Kato and Fujita introduced the notion of mild solutions Reference 7, which are solutions that satisfy the Navier–Stokes equation,

in the sense of convolution with the heat kernel as in Duhamel’s formula. They used this notion of solution, in particular the higher regularity that can be extracted from the heat kernel, to show that the Navier–Stokes equation has unique, smooth solutions locally in time for arbitrary initial data . Mild solutions are only known to exist locally in time, however, and so this approach based on the heat semigroup cannot guarantee the existence of global-in-time smooth solutions of the Navier–Stokes equation. It is one of the biggest open questions in the field of nonlinear PDEs whether smooth solutions of the Navier–Stokes equations can develop singularities in finite time. If the norm remains bounded this guarantees that a solution of the Navier–Stokes equation must remain smooth, but there is no known bound on the growth of the norm of . One bound we do have for smooth solutions of the Navier–Stokes equation is the energy equality.

Proposition 1.1.

Suppose is a mild solution to the Navier–Stokes equation. Then for all

The energy equality gives us bounds on solutions of the Navier–Stokes equation in and , but this cannot guarantee regularity, because both of these bounds are supercritical with respect to the scale invariance of the Navier–Stokes equation. The solution set of the Navier–Stokes equation is invariant under the rescaling

If we do have a bound on in a scale invariant space then that is enough to guarantee regularity. In particular, the Ladyzhenskaya-Prodi-Serrin regularity criterion Reference 13Reference 23Reference 24 states that if a smooth solution of the Navier–Stokes equation develops a singularity in finite-time , then for all , then

It is easy to check that this regularity criterion is scale critical with respect to the rescaling in Equation 1.16. This result was extended to the endpoint case, , by Escauriaza, Seregin, and Šverák Reference 6, where they proved that if , then

There have been many extensions of these regularity criteria, far more than we can discuss here, so we will confine ourselves to discussing regularity criteria directly related to this paper. For a very thorough overview of the literature on regularity criteria for solutions to the Navier-Stokes equation, see Chapter 11 in Reference 14.

One regularity criterion proven by Chae and Choe Reference 2 has a particular geometric significance. Chae and Choe prove a scale critical regularity criterion on two components of vorticity. For a two dimensional solution of the Navier–Stokes equation in the plane, the vorticity is entirely in the direction, perpendicular to the plane. This means that if the vorticity restricted to the plane or, using rotational invariance, any fixed plane, remains bounded in a scale critical space, then the solution is not too far from being two dimensional. The size of the vorticity restricted to a plane can be interpreted as a measure of how fully three dimensional a solution of the Navier–Stokes equation is, so Chae and Choe’s regularity criterion can be interpreted as saying that blowup must be fully three dimensional. Chae and Choe’s result is the following.

Theorem 1.2.

Suppose is a mild solution of the Navier–Stokes equation, and let . Then for all , there exists depending only on such that for all

In particular, if , then

Remark 1.3.

Chae and Choe’s result is not stated in this form in Reference 2. Chae and Choe prove a regularity criterion on two components of the vorticity, . However, we can see that

This means that a regularity criterion on is equivalent to a regularity criterion on . Using the rotational invariance of the Navier–Stokes equation, this is equivalent to a regularity criterion on for any fixed unit vector . We note that the Navier–Stokes equation is rotationally invariant in the sense that if is a rotation matrix on , and is a solution of the Navier–Stokes equation, then is also a solution of the Navier–Stokes equation where

See chapter 1 in Reference 16 for further discussion.

We will note that because is a derivative of , the vorticity has the rescaling,

If is a solution of the vorticity equation, then so is for all . Theorem 1.2 is critical with respect to this rescaling.

Theorem 1.2 was then extended into Besov spaces by Chen and Zhang Reference 25 and more recently into the endpoint Besov space, , for by Guo, Kučera, and Skalák Reference 8. This is an improvement on Theorem 1.2, because we have an embedding , and is the largest translation invariant Banach space with the same scaling relation as . This is therefore, the furthest advance that can be made to Chae and Choe’s regularity criterion solely by loosening the assumptions on the space.

Another regularity criterion with geometric significance is the regularity criterion in terms of the positive part of the intermediate eigenvalue of the strain matrix. This was first proven by Neustupa and Penel in Reference 19Reference 20Reference 21 and independently by the author using different methods in Reference 17.

Theorem 1.4.

Suppose is a mild solution of the Navier–Stokes equation. Let be the eigenvalues of , and let . Then for all , there exists depending only on such that for all

In particular, if , then

This regularity criterion gives a geometric characterization of the structure of potential finite-time blowup solutions of the Navier–Stokes equation. Theorem 1.4 says that in order for blowup to occur, the flow needs to be stretching in two directions, while compressing more strongly in the third. Points where the strain has two positive eigenvalues, corresponding to stretching in two directions, and one negative eigenvalue, corresponding to compressing more strongly in a third, are the points that drive enstrophy growth and hence blowup. This provides more insight into the qualitative properties of blowup solutions than the regularity criteria that just involve the size of or . The author also proved the following corollary of Theorem 1.4 in Reference 17.

Corollary 1.5.

Suppose is a mild solution of the Navier–Stokes equation. Let , with almost everywhere. Then for all , there exists depending only on such that for all

In particular, if , then

This result follows from Theorem 1.4 because , so is the smallest eigenvalue of in magnitude. Therefore, if , then

Corollary 1.5 follows immediately from this fact and Theorem 1.4. We will show in Section 3 that the special case of Corollary 1.5 where we take to be a fixed unit vector is equivalent to Chae and Choe’s result, Theorem 1.2.

This raises an interesting question: can Chae and Choe’s result be extended to a regularity criterion on , where is a unit vector that is allowed to vary in space. Clearly, unlike in Corollary 1.5, we won’t be able to take an arbitrary unit vector, otherwise we would simply take , and regularity would be guaranteed for any solution of the Navier–Stokes equation. In fact, using Corollary 1.5, it is possible to improve Chae and Choe’s result to one that allows the unit vector to vary in space and time, so long as remains bounded. The main theorem of this paper is the following.

Theorem 1.6.

Suppose is a mild solution of the Navier–Stokes equation. Suppose , with almost everywhere, and suppose . Then for all for all

where is a constant independent of , taken as in Theorem 1.4 and Corollary 1.5.

In particular, if , then

Chae and Choe’s result, Theorem 1.2, states that for a solution of the Navier–Stokes equation to blowup in finite-time, the vorticity must become unbounded in every fixed plane. For a two dimensional flow in the plane, the vorticity is entirely in the direction, so Theorem 1.2 can be interpreted as requiring the blowup of vorticity to be globally three dimensional. Theorem 1.6, strengthens this result, by requiring the vorticity to blowup in every plane, where the plane may vary in space and time so long as the gradient of the vector orthogonal to the plane remains bounded, meaning the geometry of the blowup must be locally three dimensional. In particular, this means that in order to blowup in finite-time, a solution of the Navier–Stokes equation must be fully three dimensional at all length scales, which is consistent with the Kolmogorov phenomenological theory of turbulence, as we will discuss shortly.

This result highlights the strength of the strain formulation of the Navier–Stokes regularity problem. Using the strain formulation, we are able to get a stronger regularity criterion in terms of the vorticity than by working directly with the vorticity formulation. Theorem 1.6 follows as a fairly direct corollary of Corollary 1.5 using the relationship between the structure of the strain and vorticity that is imposed by the divergence free condition on the velocity. Because the regularity criterion on gives us geometric information that is fundamentally anisotropic in that it does not involve any fixed direction, it allows us to go from a component reduction regularity criterion that involves some fixed direction, and therefore guarantees regularity as long as the solution is not too far away from being globally two dimensional in some sense, to a component reduction type regularity criterion that only requires that the solution is not too far away from being locally two dimensional. The regularity criterion in terms of first proven by Neustupa and Penel therefore encodes fundamentally anisotropic information about the structure of possible blowup solutions that does not require the imposition of any arbitrary direction. For this reason it is quite powerful.

We should also note that while Chae and Choe’s result holds for , the proof of Theorem 1.6 only holds with , , because the Hilbert space structure of is essential to the proof. It is possible that Theorem 1.6 can be generalized for all , but this would require much more delicate analysis, and likely somewhat different techniques.

There are several other scale critical, component reduction type regularity criteria. For instance, Kukavica and Ziane Reference 12 showed that if a smooth solution of the Navier–Stokes equation blows up in finite-time , and if , with , then

More recently, it was shown by Chemin and Zhang Reference 3 and Chemin, Zhang, and Zhang Reference 4 that if and , then

Neustupa, Novotný, and Penel had previously proven a regularity criterion on Reference 18, which was the first result of this type, but their regularity criteria was not scale critical; it required to be in the subcritical space , with . Like Theorem 1.2 due to Chae and Choe, all of these results say that the flow must be regular as long as it is not too three dimensional in the sense that either or remains bounded in the appropriate scale critical space. The rotational invariance of the Navier–Stokes equation implies that there is nothing special about the particular direction , so and can be replaced by and respectively for any unit vector , but the vector cannot be allowed to vary in space. These regularity criteria are therefore global anistotropic regularity criteria in that they require solutions which are globally anisotropic in some sense to remain smooth. Theorem 1.6 is significantly stronger because it is a locally anisotropic regularity criterion which only requires solutions to be locally anisotropic to remain smooth.

There is one recent regularity result that does involve local anisotropy. Kukavica, Rusin and Ziane proved that if is a suitable weak solution on , and for some domain , we have , with , then is Hölder continuous on Reference 11. This is a locally anisotropic regularity criterion, because we only require control on in some domain, not globally. This implies that the control on the anisotropy is local. This differs from Theorem 1.6, which involves an estimate over the whole space, and is therefore a global, not local regularity criterion, but a global regularity criterion that is locally anisotropic, on account of the direction being allowed to vary. We will discuss the relationship between Kukavica, Rusin, and Ziane’s regularity criterion and Theorem 1.6 further in section 3.

Another regularity criterion related to Theorem 1.6 is the regularity criterion in terms of the vorticity direction proven by Constantin and Fefferman, which states that the direction of the vorticity must vary rapidly in regions where the vorticity is large in order for a solution of the Navier–Stokes equation to blowup Reference 5. This was a very important advance in that the rapid change of the vorticity direction in regions of large vorticity has long been at least heuristically understood as a fundamental property of turbulent flow, and Constantin and Fefferman’s result shows that any Navier–Stokes blowup solution must be “turbulent” in this sense. Indeed, even Leonardo da Vinci’s qualitative studies of turbulence in the early 1500s (see Figure 1) show an understanding of the fundamental character of a rapid change in the orientation of vortices in turbulent regions of the fluid to the phenomenon of turbulence Reference 9.

Kolmogorov would rigourously describe the phenomenon depicted purely heuristically by da Vinci with his celebrated theory of turbulent cascades Reference 10, which remains central to our understanding of turbulence. Kolmogorov showed that turbulence should involve a transfer of energy from long length scales to shorter length scales, giving the appropriate scaling, and also that turbulence is locally isotropic, with no preferred vorticity direction down to the smallest length scales. Kolmogorov’s scaling law was extended to Fourier space by Obukhov in Reference 22, where he showed that turbulence should involve a scaling of the energy spectrum of the form , in the inertial range.

Constantin and Fefferman were the first to connect the change in the orientation of vortices in turbulent regions to the Navier–Stokes regularity problem, which suggests that blowup for the Navier–Stokes equation is not only of purely mathematical interest, but could have significant physical implications—particularly if there is finite-time blowup—on our phenomenological understanding of turbulence. Constantin and Fefferman’s result was later generalized by Beirão da Veiga and Berselli in Reference 1. In particular, as a corollary of their refined result that the vorticity direction must vary rapidly in regions with large vorticity, they proved a regularity criterion in terms of the gradient of the vorticity direction, .

Theorem 1.7.

Suppose is a mild solution of the Navier–Stokes equation that blows up in finite-time . Then for all .

This is related to Theorem 1.6, because if we take , then we find that Theorem 1.6 implies if , then

If we take the case of Theorem 1.7, then we can see that if a solution of the Navier–Stokes equation blows up in finite-time , then

We will note that Beirão da Veiga and Berselli’s result is stronger than the special case of Theorem 1.6 where we take because it not only specifies that , must become unbounded as we approach blowup time, but also a rate of blowup—namely that its integral to the fourth power must go to infinity. We can see that the two extremal cases of Theorem 1.6 are the case where is constant and hence , in which case we recover Theorem 1.2 from Chae and Choe with , , and the case where and hence , in which case we recover a weaker form of Theorem 1.7, from Beirão da Veiga and Berselli. Therefore this result could be said to interpolate between these two results, although suboptimally at one end. We will discuss why this interpolation is suboptimal in section 3.

In section 2, we will prove the main result of the paper, Theorem 1.6. In section 3, we will further discuss the relationship between this result and the previous literature.

2. Proof of the main theorem

We will now prove Theorem 1.6, which is restated here for the reader’s convenience.

Theorem 2.1.

Suppose is a mild solution of the Navier–Stokes equation. Suppose , with almost everywhere, and suppose . Then for all

where is taken is in Theorem 1.4 and Corollary 1.5. In particular, if , then

Proof.

We will prove that this theorem is a corollary of Corollary 1.5. In particular, we will show that

First we will recall from Equation 1.6 that

Therefore we find that

Likewise we may compute that

and therefore

Putting together Equation 2.6 and Equation 2.8, we find that

We know that for all , so we find that

Now we will need to estimate . By definition we have

Integrating by parts with respect to , and applying the fact that , we find that

This estimate implies that

Applying the energy equality and our hypothesis on , we find that

Therefore we may conclude that

Applying Corollary 1.5, this completes the proof.

3. Relationship to the previous literature

In this section, we will further discuss the relationship between the results in this paper and the previous literature. In particular, we will show that Theorem 1.5, the regularity criterion on , is equivalent to Chae and Choe’s regularity criterion on in the special case where we take to be a fixed unit vector. In order to do this we will first need to introduce the Helmholtz decomposition of vector fields in into gradients and divergence free vector fields.

Proposition 3.1.

Suppose . For all , there exists a unique , and such that , and there exists depending only on , such that

and

Note because we do not have any assumptions of higher regularity, we will say that , if for all

and we will say that is a gradient if for all , we have

Using this decomposition, we will define the projection onto the space of divergence free vector fields and space of gradients in .

Definition 3.2.

For all , let the projection onto the space of divergence free vector fields, , be given by

and let the projection onto the space of gradients, , be given by

where are taken as in Proposition 3.1. Note in particular that Proposition 3.1 implies that and are bounded operators on .

We can now show that for any fixed unit vector and are equivalent norms.

Proposition 3.3.

Suppose , and Then for all unit vectors , and for all

where is taken as in Proposition 3.1. Furthermore,

Proof.

First we will observe that the rotational invariance of the space of divergence free vector fields means that we can take without loss of generality. It is a simple calculation to see that

Likewise we can see that

Clearly is a gradient, and we can also see that is divergence free because

This implies that

Applying Proposition 3.1 we can therefore conclude that

Likewise we can observe that

and apply Proposition 3.1 to find that

We will now finish the proof by applying the triangle inequality and concluding that

and that

This completes the proof.

Proposition 3.3 implies that in the range , Kukavica and Ziane’s regularity criterion on implies Chae and Choe’s regularity criterion on because

Likewise, Chemin, Zhang, and Zhang’s regularity criterion on implies Chae and Choe’s regularity criterion on when . We can see this by applying the fractional Sobolev embedding when , and Proposition 3.3 to find

We will also note that Theorem 1.6 differs in a fundamental way from the locally anisotropic regularity criterion proven by Kukavica, Rusin, and Ziane in Reference 11. This is true in particular because while

no such inequality holds if we are not on the whole space. For arbitrary , and arbitrary domains , there is no inequality of the form

The proof of Proposition 3.3 relies on the boundedness of the Helmholtz decomoposition, which in turn relies on the boundedness of the Riesz transform. This is something which only holds globally, not locally, because the Riesz transform is a nonlocal operator. controls globally in , but not locally in . This means that while Kukavica and Ziane’s regularity criterion for implies Chae and Choe’s regularity criterion for for , the local anisotropic regularity criterion with and the locally anisotropic regularity criterion on , where is allowed to vary do not have this same relationship.

Finally, we will note that when we set , Theorem 1.6 requires that if our solution of the Navier–Stokes equation blows up in finite-time , then

as long as on at most a set of measure zero. However, as we discussed in the introduction, Beirão da Veiga and Berselli proved the stronger result that under these conditions

For this reason, it seems like it ought to be possible to relax the condition in Theorem 1.6 from to . The methods used to prove Theorem 1.6 would at first glance suggest this would be possible, but the difficulty is that when integrating by parts, we cannot get all of the derivatives off of and onto , which is what leads to the sub-optimal result in this endpoint case.

Remark 3.4.

Suppose for all almost everywhere , we had a bound of the form

Then using the energy inequality we could conclude that

and we could relax the requirement in Theorem 1.6 to using precisely the same proof. The difficulty is that, as much as at first glance it would appear that an inequality of the form Equation 3.37 should hold using integration by parts, after multiple attempts to do so by the author it does not appear possible to push all the derivatives off of and onto in the manner desired. The asymmetry from the transpose, , does not seem to allow this. It is definitely possible, perhaps even likely, that the condition can be relaxed to , which would make the interpolation between Chae and Choe’s regularity criterion and Beirão da Veiga and Berselli’s optimal, but the proof would probably require some fundamentally new ideas, because it is unlikely to follow as a corollary of the regularity criterion on in Theorem 1.4.

Figures

Figure 1.

Leonardo da Vinci, Studies of Turbulent Water. Image credit: Royal Collection Trust/© Her Majesty Queen Elizabeth II 2021.

Graphic without alt text

Mathematical Fragments

Equation (1.6)
Equation (1.16)
Theorem 1.2.

Suppose is a mild solution of the Navier–Stokes equation, and let . Then for all , there exists depending only on such that for all

In particular, if , then

Theorem 1.4.

Suppose is a mild solution of the Navier–Stokes equation. Let be the eigenvalues of , and let . Then for all , there exists depending only on such that for all

In particular, if , then

Corollary 1.5.

Suppose is a mild solution of the Navier–Stokes equation. Let , with almost everywhere. Then for all , there exists depending only on such that for all

In particular, if , then

Theorem 1.6.

Suppose is a mild solution of the Navier–Stokes equation. Suppose , with almost everywhere, and suppose . Then for all for all

where is a constant independent of , taken as in Theorem 1.4 and Corollary 1.5.

In particular, if , then

Theorem 1.7.

Suppose is a mild solution of the Navier–Stokes equation that blows up in finite-time . Then for all .

Equation (2.6)
Equation (2.8)
Proposition 3.1.

Suppose . For all , there exists a unique , and such that , and there exists depending only on , such that

and

Note because we do not have any assumptions of higher regularity, we will say that , if for all

and we will say that is a gradient if for all , we have

Proposition 3.3.

Suppose , and Then for all unit vectors , and for all

where is taken as in Proposition 3.1. Furthermore,

Remark 3.4.

Suppose for all almost everywhere , we had a bound of the form

Then using the energy inequality we could conclude that

and we could relax the requirement in Theorem 1.6 to using precisely the same proof. The difficulty is that, as much as at first glance it would appear that an inequality of the form 3.37 should hold using integration by parts, after multiple attempts to do so by the author it does not appear possible to push all the derivatives off of and onto in the manner desired. The asymmetry from the transpose, , does not seem to allow this. It is definitely possible, perhaps even likely, that the condition can be relaxed to , which would make the interpolation between Chae and Choe’s regularity criterion and Beirão da Veiga and Berselli’s optimal, but the proof would probably require some fundamentally new ideas, because it is unlikely to follow as a corollary of the regularity criterion on in Theorem 1.4.

References

Reference [1]
Hugo Beirão da Veiga and Luigi C. Berselli, On the regularizing effect of the vorticity direction in incompressible viscous flows, Differential Integral Equations 15 (2002), no. 3, 345–356. MR1870646,
Show rawAMSref \bib{daViegaBerselli}{article}{ author={Beir\~{a}o da Veiga, Hugo}, author={Berselli, Luigi C.}, title={On the regularizing effect of the vorticity direction in incompressible viscous flows}, journal={Differential Integral Equations}, volume={15}, date={2002}, number={3}, pages={345--356}, issn={0893-4983}, review={\MR {1870646}}, }
Reference [2]
Dongho Chae and Hi-Jun Choe, Regularity of solutions to the Navier-Stokes equation, Electron. J. Differential Equations (1999), No. 05, 7. MR1673067,
Show rawAMSref \bib{ChaeChoe}{article}{ author={Chae, Dongho}, author={Choe, Hi-Jun}, title={Regularity of solutions to the Navier-Stokes equation}, journal={Electron. J. Differential Equations}, date={1999}, pages={No. 05, 7}, review={\MR {1673067}}, }
Reference [3]
Jean-Yves Chemin and Ping Zhang, On the critical one component regularity for 3-D Navier-Stokes systems (English, with English and French summaries), Ann. Sci. Éc. Norm. Supér. (4) 49 (2016), no. 1, 131–167, DOI 10.24033/asens.2278. MR3465978,
Show rawAMSref \bib{CheminZhang}{article}{ author={Chemin, Jean-Yves}, author={Zhang, Ping}, title={On the critical one component regularity for 3-D Navier-Stokes systems}, language={English, with English and French summaries}, journal={Ann. Sci. \'{E}c. Norm. Sup\'{e}r. (4)}, volume={49}, date={2016}, number={1}, pages={131--167}, issn={0012-9593}, review={\MR {3465978}}, doi={10.24033/asens.2278}, }
Reference [4]
Jean-Yves Chemin, Ping Zhang, and Zhifei Zhang, On the critical one component regularity for 3-D Navier-Stokes system: general case, Arch. Ration. Mech. Anal. 224 (2017), no. 3, 871–905, DOI 10.1007/s00205-017-1089-0. MR3621812,
Show rawAMSref \bib{CheminZhangZhang}{article}{ author={Chemin, Jean-Yves}, author={Zhang, Ping}, author={Zhang, Zhifei}, title={On the critical one component regularity for 3-D Navier-Stokes system: general case}, journal={Arch. Ration. Mech. Anal.}, volume={224}, date={2017}, number={3}, pages={871--905}, issn={0003-9527}, review={\MR {3621812}}, doi={10.1007/s00205-017-1089-0}, }
Reference [5]
Peter Constantin and Charles Fefferman, Direction of vorticity and the problem of global regularity for the Navier-Stokes equations, Indiana Univ. Math. J. 42 (1993), no. 3, 775–789, DOI 10.1512/iumj.1993.42.42034. MR1254117,
Show rawAMSref \bib{ConstantinFefferman}{article}{ author={Constantin, Peter}, author={Fefferman, Charles}, title={Direction of vorticity and the problem of global regularity for the Navier-Stokes equations}, journal={Indiana Univ. Math. J.}, volume={42}, date={1993}, number={3}, pages={775--789}, issn={0022-2518}, review={\MR {1254117}}, doi={10.1512/iumj.1993.42.42034}, }
Reference [6]
L. Iskauriaza, G. A. Serëgin, and V. Shverak, -solutions of Navier-Stokes equations and backward uniqueness (Russian, with Russian summary), Uspekhi Mat. Nauk 58 (2003), no. 2(350), 3–44, DOI 10.1070/RM2003v058n02ABEH000609; English transl., Russian Math. Surveys 58 (2003), no. 2, 211–250. MR1992563,
Show rawAMSref \bib{SverakSereginL3}{article}{ author={Iskauriaza, L.}, author={Ser\"{e}gin, G. A.}, author={Shverak, V.}, title={$L_{3,\infty }$-solutions of Navier-Stokes equations and backward uniqueness}, language={Russian, with Russian summary}, journal={Uspekhi Mat. Nauk}, volume={58}, date={2003}, number={2(350)}, pages={3--44}, issn={0042-1316}, translation={ journal={Russian Math. Surveys}, volume={58}, date={2003}, number={2}, pages={211--250}, issn={0036-0279}, }, review={\MR {1992563}}, doi={10.1070/RM2003v058n02ABEH000609}, }
Reference [7]
Hiroshi Fujita and Tosio Kato, On the Navier-Stokes initial value problem. I, Arch. Rational Mech. Anal. 16 (1964), 269–315, DOI 10.1007/BF00276188. MR166499,
Show rawAMSref \bib{KatoFujita}{article}{ author={Fujita, Hiroshi}, author={Kato, Tosio}, title={On the Navier-Stokes initial value problem. I}, journal={Arch. Rational Mech. Anal.}, volume={16}, date={1964}, pages={269--315}, issn={0003-9527}, review={\MR {166499}}, doi={10.1007/BF00276188}, }
Reference [8]
Zhengguang Guo, Petr Kučera, and Zdeněk Skalák, Regularity criterion for solutions to the Navier-Stokes equations in the whole 3D space based on two vorticity components, J. Math. Anal. Appl. 458 (2018), no. 1, 755–766, DOI 10.1016/j.jmaa.2017.09.029. MR3711930,
Show rawAMSref \bib{GuoBesov}{article}{ author={Guo, Zhengguang}, author={Ku\v {c}era, Petr}, author={Skal\'{a}k, Zden\v {e}k}, title={Regularity criterion for solutions to the Navier-Stokes equations in the whole 3D space based on two vorticity components}, journal={J. Math. Anal. Appl.}, volume={458}, date={2018}, number={1}, pages={755--766}, issn={0022-247X}, review={\MR {3711930}}, doi={10.1016/j.jmaa.2017.09.029}, }
Reference [9]
Martin Kemp, Leonardo da Vinci’s laboratory: studies in flow, Nature 571 (2019), 322–323.
Reference [10]
A. Kolmogoroff, The local structure of turbulence in incompressible viscous fluid for very large Reynold’s numbers, C. R. (Doklady) Acad. Sci. URSS (N.S.) 30 (1941), 301–305. MR0004146,
Show rawAMSref \bib{Kolmogorov}{article}{ author={Kolmogoroff, A.}, title={The local structure of turbulence in incompressible viscous fluid for very large Reynold's numbers}, journal={C. R. (Doklady) Acad. Sci. URSS (N.S.)}, volume={30}, date={1941}, pages={301--305}, review={\MR {0004146}}, }
Reference [11]
Igor Kukavica, Walter Rusin, and Mohammed Ziane, Localized anisotropic regularity conditions for the Navier-Stokes equations, J. Nonlinear Sci. 27 (2017), no. 6, 1725–1742, DOI 10.1007/s00332-017-9382-5. MR3713929,
Show rawAMSref \bib{KukavicaLocal}{article}{ author={Kukavica, Igor}, author={Rusin, Walter}, author={Ziane, Mohammed}, title={Localized anisotropic regularity conditions for the Navier-Stokes equations}, journal={J. Nonlinear Sci.}, volume={27}, date={2017}, number={6}, pages={1725--1742}, issn={0938-8974}, review={\MR {3713929}}, doi={10.1007/s00332-017-9382-5}, }
Reference [12]
Igor Kukavica and Mohammed Ziane, Navier-Stokes equations with regularity in one direction, J. Math. Phys. 48 (2007), no. 6, 065203, 10, DOI 10.1063/1.2395919. MR2337002,
Show rawAMSref \bib{Kukavica}{article}{ author={Kukavica, Igor}, author={Ziane, Mohammed}, title={Navier-Stokes equations with regularity in one direction}, journal={J. Math. Phys.}, volume={48}, date={2007}, number={6}, pages={065203, 10}, issn={0022-2488}, review={\MR {2337002}}, doi={10.1063/1.2395919}, }
Reference [13]
O. A. Ladyženskaja, Uniqueness and smoothness of generalized solutions of Navier-Stokes equations (Russian), Zap. Naučn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI) 5 (1967), 169–185. MR0236541,
Show rawAMSref \bib{Ladyzhenskaya}{article}{ author={Lady\v {z}enskaja, O. A.}, title={Uniqueness and smoothness of generalized solutions of Navier-Stokes equations}, language={Russian}, journal={Zap. Nau\v {c}n. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI)}, volume={5}, date={1967}, pages={169--185}, review={\MR {0236541}}, }
Reference [14]
Pierre Gilles Lemarié-Rieusset, The Navier-Stokes problem in the 21st century, CRC Press, Boca Raton, FL, 2016, DOI 10.1201/b19556. MR3469428,
Show rawAMSref \bib{NS21}{book}{ author={Lemari\'{e}-Rieusset, Pierre Gilles}, title={The Navier-Stokes problem in the 21st century}, publisher={CRC Press, Boca Raton, FL}, date={2016}, pages={xxii+718}, isbn={978-1-4665-6621-7}, review={\MR {3469428}}, doi={10.1201/b19556}, }
Reference [15]
Jean Leray, Sur le mouvement d’un liquide visqueux emplissant l’espace (French), Acta Math. 63 (1934), no. 1, 193–248, DOI 10.1007/BF02547354. MR1555394,
Show rawAMSref \bib{Leray}{article}{ author={Leray, Jean}, title={Sur le mouvement d'un liquide visqueux emplissant l'espace}, language={French}, journal={Acta Math.}, volume={63}, date={1934}, number={1}, pages={193--248}, issn={0001-5962}, review={\MR {1555394}}, doi={10.1007/BF02547354}, }
Reference [16]
Andrew J. Majda and Andrea L. Bertozzi, Vorticity and incompressible flow, Cambridge Texts in Applied Mathematics, vol. 27, Cambridge University Press, Cambridge, 2002. MR1867882,
Show rawAMSref \bib{MajdaBertozzi}{book}{ author={Majda, Andrew J.}, author={Bertozzi, Andrea L.}, title={Vorticity and incompressible flow}, series={Cambridge Texts in Applied Mathematics}, volume={27}, publisher={Cambridge University Press, Cambridge}, date={2002}, pages={xii+545}, isbn={0-521-63057-6}, isbn={0-521-63948-4}, review={\MR {1867882}}, }
Reference [17]
Evan Miller, A regularity criterion for the Navier-Stokes equation involving only the middle eigenvalue of the strain tensor, Arch. Ration. Mech. Anal. 235 (2020), no. 1, 99–139, DOI 10.1007/s00205-019-01419-z. MR4062474,
Show rawAMSref \bib{MillerStrain}{article}{ author={Miller, Evan}, title={A regularity criterion for the Navier-Stokes equation involving only the middle eigenvalue of the strain tensor}, journal={Arch. Ration. Mech. Anal.}, volume={235}, date={2020}, number={1}, pages={99--139}, issn={0003-9527}, review={\MR {4062474}}, doi={10.1007/s00205-019-01419-z}, }
Reference [18]
Jirí Neustupa, Antonín Novotný, and Patrick Penel, An interior regularity of a weak solution to the Navier-Stokes equations in dependence on one component of velocity, Topics in mathematical fluid mechanics, Quad. Mat., vol. 10, Dept. Math., Seconda Univ. Napoli, Caserta, 2002, pp. 163–183. MR2051774,
Show rawAMSref \bib{NeustupaOneComp}{article}{ author={Neustupa, Jir\'{\i }}, author={Novotn\'{y}, Anton\'{\i }n}, author={Penel, Patrick}, title={An interior regularity of a weak solution to the Navier-Stokes equations in dependence on one component of velocity}, conference={ title={Topics in mathematical fluid mechanics}, }, book={ series={Quad. Mat.}, volume={10}, publisher={Dept. Math., Seconda Univ. Napoli, Caserta}, }, date={2002}, pages={163--183}, review={\MR {2051774}}, }
Reference [19]
Jiří Neustupa and Patrick Penel, Anisotropic and geometric criteria for interior regularity of weak solutions to the 3D Navier-Stokes equations, Mathematical fluid mechanics, Adv. Math. Fluid Mech., Birkhäuser, Basel, 2001, pp. 237–265. MR1865056,
Show rawAMSref \bib{NeuPen1}{article}{ author={Neustupa, Ji\v {r}\'{\i }}, author={Penel, Patrick}, title={Anisotropic and geometric criteria for interior regularity of weak solutions to the 3D Navier-Stokes equations}, conference={ title={Mathematical fluid mechanics}, }, book={ series={Adv. Math. Fluid Mech.}, publisher={Birkh\"{a}user, Basel}, }, date={2001}, pages={237--265}, review={\MR {1865056}}, }
Reference [20]
Jiří Neustupa and Patrick Penel, The role of eigenvalues and eigenvectors of the symmetrized gradient of velocity in the theory of the Navier-Stokes equations (English, with English and French summaries), C. R. Math. Acad. Sci. Paris 336 (2003), no. 10, 805–810, DOI 10.1016/S1631-073X(03)00174-2. MR1990019,
Show rawAMSref \bib{NeuPen2}{article}{ author={Neustupa, Ji\v {r}\'{\i }}, author={Penel, Patrick}, title={The role of eigenvalues and eigenvectors of the symmetrized gradient of velocity in the theory of the Navier-Stokes equations}, language={English, with English and French summaries}, journal={C. R. Math. Acad. Sci. Paris}, volume={336}, date={2003}, number={10}, pages={805--810}, issn={1631-073X}, review={\MR {1990019}}, doi={10.1016/S1631-073X(03)00174-2}, }
Reference [21]
Jiří Neustupa and Patrick Penel, Regularity of a weak solution to the Navier-Stokes equation in dependence on eigenvalues and eigenvectors of the rate of deformation tensor, Trends in partial differential equations of mathematical physics, Progr. Nonlinear Differential Equations Appl., vol. 61, Birkhäuser, Basel, 2005, pp. 197–212, DOI 10.1007/3-7643-7317-2_15. MR2129619,
Show rawAMSref \bib{NeuPen3}{article}{ author={Neustupa, Ji\v {r}\'{\i }}, author={Penel, Patrick}, title={Regularity of a weak solution to the Navier-Stokes equation in dependence on eigenvalues and eigenvectors of the rate of deformation tensor}, conference={ title={Trends in partial differential equations of mathematical physics}, }, book={ series={Progr. Nonlinear Differential Equations Appl.}, volume={61}, publisher={Birkh\"{a}user, Basel}, }, date={2005}, pages={197--212}, review={\MR {2129619}}, doi={10.1007/3-7643-7317-2\_15}, }
Reference [22]
A. Obukhoff, On the energy distribution in the spectrum of a turbulent flow, C. R. (Doklady) Acad. Sci. URSS (N.S.) 32 (1941), 19–21. MR0005852,
Show rawAMSref \bib{Obukhov}{article}{ author={Obukhoff, A.}, title={On the energy distribution in the spectrum of a turbulent flow}, journal={C. R. (Doklady) Acad. Sci. URSS (N.S.)}, volume={32}, date={1941}, pages={19--21}, review={\MR {0005852}}, }
Reference [23]
Giovanni Prodi, Un teorema di unicità per le equazioni di Navier-Stokes (Italian), Ann. Mat. Pura Appl. (4) 48 (1959), 173–182, DOI 10.1007/BF02410664. MR126088,
Show rawAMSref \bib{Prodi}{article}{ author={Prodi, Giovanni}, title={Un teorema di unicit\`a per le equazioni di Navier-Stokes}, language={Italian}, journal={Ann. Mat. Pura Appl. (4)}, volume={48}, date={1959}, pages={173--182}, issn={0003-4622}, review={\MR {126088}}, doi={10.1007/BF02410664}, }
Reference [24]
James Serrin, On the interior regularity of weak solutions of the Navier-Stokes equations, Arch. Rational Mech. Anal. 9 (1962), 187–195, DOI 10.1007/BF00253344. MR136885,
Show rawAMSref \bib{Serrin}{article}{ author={Serrin, James}, title={On the interior regularity of weak solutions of the Navier-Stokes equations}, journal={Arch. Rational Mech. Anal.}, volume={9}, date={1962}, pages={187--195}, issn={0003-9527}, review={\MR {136885}}, doi={10.1007/BF00253344}, }
Reference [25]
Zhang Zhifei and Chen Qionglei, Regularity criterion via two components of vorticity on weak solutions to the Navier-Stokes equations in , J. Differential Equations 216 (2005), no. 2, 470–481, DOI 10.1016/j.jde.2005.06.001. MR2165305,
Show rawAMSref \bib{ChenZhangBesov}{article}{ author={Zhifei, Zhang}, author={Qionglei, Chen}, title={Regularity criterion via two components of vorticity on weak solutions to the Navier-Stokes equations in $\mathbb {R}^3$}, journal={J. Differential Equations}, volume={216}, date={2005}, number={2}, pages={470--481}, issn={0022-0396}, review={\MR {2165305}}, doi={10.1016/j.jde.2005.06.001}, }

Article Information

MSC 2020
Primary: 35Q30 (Navier-Stokes equations)
Author Information
Evan Miller
Department of Mathematics and Statistics, McMaster University, Hamilton, Ontario, Canada
emiller@msri.org
ORCID
MathSciNet
Communicated by
Catherine Sulem
Journal Information
Proceedings of the American Mathematical Society, Series B, Volume 8, Issue 6, ISSN 2330-1511, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2021 by the author under Creative Commons Attribution 3.0 License (CC BY 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/bproc/74
  • MathSciNet Review: 4214337
  • Show rawAMSref \bib{4214337}{article}{ author={Miller, Evan}, title={A locally anisotropic regularity criterion for the Navier--Stokes equation in terms of vorticity}, journal={Proc. Amer. Math. Soc. Ser. B}, volume={8}, number={6}, date={2021}, pages={60-74}, issn={2330-1511}, review={4214337}, doi={10.1090/bproc/74}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.