Convex integration constructions in hydrodynamics

By Tristan Buckmaster and Vlad Vicol

Abstract

We review recent developments in the field of mathematical fluid dynamics which utilize techniques that go under the umbrella name convex integration. In the hydrodynamical context, these methods produce paradoxical solutions to the fluid equations which defy physical laws. These counterintuitive solutions have a number of properties that resemble predictions made by phenomenological theories of fluid turbulence. The goal of this review is to highlight some of these similarities while maintaining an emphasis on rigorous mathematical statements. We focus our attention on the construction of weak solutions for the incompressible Euler, Navier–Stokes, and magneto-hydrodynamic equations which violate these systems’ physical energy laws.

1. Introduction

We experience turbulent fluids on a day-to-day basis. The plume rising from a lit candle starts off as smooth and well organized (laminar) and quickly becomes wildly irregular, or chaotic. The air flow around a car in motion is typically laminar around the front of the car and becomes chaotic (turbulent) in the wake of the car. Hydrodynamic turbulence has received a tremendous amount of attention over the past century, both within the physics and mathematics literature. This has resulted in a number of phenomenological theories,⁠Footnote1 which have been very successful in making predictions about the statistics of turbulent flows. Nonetheless, to date, we do not have an unconditional, mathematically rigorous bridge between these phenomenological theories and properties of the solutions to the underlying partial differential equations (PDEs) which are meant to describe the fluid: the Navier–Stokes equations and their infinite Reynolds number limit, the Euler equations.

1

Some of these phenomenological theories may be traced back to the works of O. Reynolds, L. Prandtl, T. von Karman, G. I. Taylor, L. F. Richardson, W. Heisenberg, A. Kolmogorov, A. Obhukov, L. Onsager, L. Landau, E. Hopf, G. Batchelor, R. H. Kraichnan or P. G. Saffman, and many others. The topic is too vast to review here, and we refer the reader to Reference 9Reference 53Reference 74Reference 82Reference 133 for further references.

A slightly less ambitious goal, which nonetheless would offer tremendous insight into the nature of turbulent flows, is to start from experimental facts,⁠Footnote2 translate them into mathematical properties for solutions of the fundamental fluids PDEs, and then prove that there exist classes of solutions to these PDEs which exhibit the desired properties. In this process a certain degree of mathematical idealization is inevitable, and thus one should view such a program as showing that the PDE models are consistent with the physical reality of turbulent flows, justifying their predictive usage in computer simulations.

2

By this we mean quantitative predictions about hydrodynamic turbulence, which are confirmed experimentally both in a laboratory setting and in computer simulations to the point that these predictions are undoubted in the physics community. Examples of such experimental facts include the anomalous dissipation of energy in the infinite Reynolds number limit Reference 158Reference 172 or Kolmogorov’s law Reference 82Reference 159. In contrast, quantifying the intermittent nature of fully developed turbulent flows, for instance by measuring the scaling of the th-order structure functions for , is well known to exhibit variations depending on the experimental setup Reference 82Reference 157.

The remarkable outcome of such attempts is that sometimes beautiful and deep mathematical connections are revealed when trying to translate experimental facts into rigorous questions about PDEs. This article is about one such program, which has brought the method of convex integration (which has its roots in classical problems in geometry) into the hydrodynamic context, where it played the key role in the resolution of Onsager’s conjecture Reference 20Reference 93, the mathematical manifestation of an experimental fact in turbulence.

Building upon the seminal works of Scheffer Reference 147 and Shnirelman Reference 152Reference 153, De Lellis and Székelyhidi Jr. developed in Reference 55Reference 58 a systematic framework for applying convex integration in fluid dynamics. Since then, the mathematical techniques have evolved, in part by taking into account more detailed physical properties of the fluids equations, and this mathematical evolution has been the subject of a number of review papers Reference 22Reference 57Reference 60Reference 61Reference 162. The goal of this article is to put in perspective some of the more recent results in this program, such as the nonuniqueness of distributional solutions to the Navier–Stokes equations Reference 16Reference 23, and the existence of weak solutions to the ideal magneto-hydrodynamic (MHD) equations which do not conserve magnetic helicity Reference 10.

We discuss the three flavors of convex integration encountered in hydrodynamics, which originated in the works Reference 55Reference 58, and Reference 23, respectively. The first is the classical convex integration in which one relaxes the equation, constructs a suitable notion of subsolution, to which one then adds high frequency plane wave corrections of suitable amplitudes. Through a powerful abstract functional analytic machinery originating in the field of differential inclusions, this procedure may be shown to produce a sequence of approximate solutions which converge weak-* to a bounded weak solution. The second flavor is the convex integration scheme, which is both motivated and also resembles the earlier schemes of Nash and Kuiper for the isometric embedding problem. In this scheme, the approximate solutions are built by incrementally adding oscillatory perturbations of higher and higher frequencies. The oscillations themselves to leading order are exact stationary solutions of the underlying PDEs. The convergence to a limiting continuous weak solution holds because the error converges to zero in the uniform norm, and the approximating sequence converges absolutely in Hölder spaces. The third flavor of convex integration, intermittent convex integration, builds on the aforementioned Nash-type scheme, but the analysis is performed in Lebesgue spaces. This scheme explores the fact that if the building blocks are to leading order exact solutions of the underlying PDEs, then some of the most dangerous error terms in the iteration are linear in terms of the highest frequency perturbation, and thus via a decoupling argument these terms are smaller than initially expected when measured in the correct Lebesgue space. In this intermittent convex integration scheme, the cancellation of errors of smaller frequency⁠Footnote3 is achieved in an average sense, rather than pointwise, as is the case for the and variants of convex integration. One way to emphasize the different types of results that may be obtained via these three approaches, is to consider hydrodynamic PDEs with more than one conservation law such as the ideal MHD equations. For this system, it is easy to discern the type of results which may be obtained via the the various flavors of convex integration Reference 10Reference 78, as they relate to the conservation laws of the model.

3

The cancellation of the previous error (Reynolds stress) in convex integration schemes essentially happens through an infinite range backwards cascade, very much like how produces a contribution of amplitude at frequency , no matter how large is! The weak solutions constructed via convex integration schemes essentially retain the memory of an access to frequency infinity. In contrast, it is known that in 3D turbulent flows, kinetic energy mostly travels from low to high frequencies, and that the energy transfer is very local, due to Equation 2.12 below.

It is fair to say that the limitations of convex integrations schemes in fluid dynamics are not yet known, emphasizing the power and the flexibility of the machinery that De Lellis and Székelyhidi have developed. A large number of physically motivated, mathematically very interesting challenges remain to be explored, and we have mentioned a number of open problems and conjectures in Section 3.6, Section 4.4, and Section 5.3.

2. Energy equalities and their validity

In this section, we discuss the energy balance relations for the Navier–Stokes, Euler, and MHD equations, which are the fundamental models governing the motion of incompressible homogenous viscous, inviscid, magnetically conducting fluids, respectively. Throughout this survey, we focus on the space dimension three (3D), on occasion making reference to two dimensional (2D) models. The physical domain in which the hydrodynamical models are considered is the periodic box , putting aside the many interesting physical phenomena and the mathematical difficulties which arise when considering these models in domains with solid walls. For all PDEs considered, the initial data, the forcing terms, and thus solutions satisfy periodic boundary conditions, have zero mean on , and are incompressible.

2.1. The Navier–Stokes equations and the zeroth law of turbulence

The PDEs governing the motion of homogenous incompressible viscous fluid flows are the Navier–Stokes equations, a manifestation of Newton’s second law of motion and the conservation of mass. The unknowns are the velocity and pressure , satisfying

Here is the kinematic viscosity of the fluid and is a zero mean incompressible forcing term. We abuse notation and write for the Reynolds number. The system Equation 2.1 is supplemented with an initial datum which has zero mean, is incompressible, and is square integrable. Note that one may rewrite the nonlinear term in Equation 2.1a in divergence form as

which is important for defining distributional solutions to the system (see Definitions 2.1 and 2.2 below). The literature concerning these equations is vast, and we refer the interested reader to the books Reference 42Reference 67Reference 81Reference 114Reference 143Reference 170Reference 171 for an overview of the field.

2.1.1. The energy balance and weak solutions

Fix , an initial datum , and a forcing term which are smooth. Consider a smooth solution of the Navier–Stokes system Equation 2.1 with this datum, and take an inner product of the forced momentum equation Equation 2.1a with . Since is smooth, we obtain the pointwise energy balance

Integrating Equation 2.2 over and using that the divergence of a smooth periodic function vanishes, we obtain the kinetic energy balance

where we have denoted the kinetic energy by

Note that the second term on the left side of Equation 2.3 is signed, and it physically represents the energy dissipation rate; the term on the right side denotes the total work of the force. We emphasize that in deriving Equation 2.3 the following cancellation played a key role:

For , the energy balance Equation 2.3 implies that

for any . The inequality Equation 2.5 is the only known coercive a priori estimate for the 3D Navier–Stokes equations, and it gives an a priori bound for the solution in the so-called energy space , solely in terms of the input and . We emphasize however that the knowledge that lies in is not sufficient to establish that the cancellation relation Equation 2.4 holds (weakly in time). This point is important to keep in mind, and it will be revisited in Section 2.1.2 below.

Using Equation 2.5, along a sequence of approximate solutions for which Equation 2.3 is justified, Leray Reference 115 (and later Hopf Reference 89 in the case of domains with boundary) proved that for any finite energy initial datum there exists a global weak solution to the Navier–Stokes equation. More precisely, Leray proved the global existence in the following class of weak solutions.

Definition 2.1 (Leray–Hopf weak solution).

A vector field

is called a Leray–Hopf weak solution of the Navier–Stokes equations if for any the vector field is weakly divergence free, has zero mean, satisfies the Navier–Stokes equations distributionally:

for any divergence free test function , and satisfies the energy inequality Equation 2.5 for .⁠Footnote4

4

A stronger form of the energy inequality holds by Leray’s construction: Equation 2.5 holds for a.e. and all . Moreover, an a priori stronger form of weak solution may be constructed by modifying Leray’s proof. These so-called suitable weak solutions obey a version of the energy inequality which is localized in space and time, which allows one to prove that they are smooth except for a putative singular space-time set with one-dimensional (1D) parabolic Hausdorff measure which is zero Reference 25Reference 145Reference 146; see also the more recent reviews Reference 114Reference 140Reference 143.

While Leray–Hopf weak solutions are known to enjoy certain desirable properties such weak-strong uniqueness Reference 115Reference 142Reference 150Reference 176 (if a smooth solution exists, then any Leray–Hopf weak solution with the same initial datum is equal to this smooth one) or epochs of regularity Reference 115 (intervals for which the solution is smooth, leading to partial regularity in time Reference 111Reference 144Reference 145), their uniqueness to date remains open. Regularity, and hence also uniqueness, is only known to hold under the extra assumption that the solution is bounded in a scaling invariant space, such as for , known as the Ladyženskaja–Prodi–Serrin conditions Reference 72Reference 103Reference 142Reference 149. In fact, it has been conjectured Reference 97Reference 98Reference 112 that the uniqueness of Leray–Hopf weak solutions fails as soon as the Ladyženskaja–Prodi–Serrin conditions are violated. In spite of compelling numerical evidence Reference 87, to date this question remains open.

Note that the distributional form of the Navier–Stokes equations Equation 2.6 makes sense as soon as , without requiring that . Thus, as in Reference 150 one may define an even weaker notion of solution to Equation 2.1 as follows.

Definition 2.2 (Weak/mild solution).

A vector field is called a weak or distributional solution of the Navier–Stokes equations if for any the vector field is weakly divergence free, has zero mean, and Equation 2.6 holds for for any smooth divergence free test function .

As shown in Reference 75, the weak solutions of Definition 2.2 satisfy the integral equation

and are called mild solutions Reference 114, Definition 6.5; here is the Helmholtz projection. We emphasize that this class of solutions is weaker than the Leray–Hopf weak solutions in Definition 2.1 because solutions need not satisfy the energy inequality Equation 2.5, and their energy dissipation rate need not be finite. Similarly to Leray–Hopf ones, these weak solutions are known to be regular under the additional assumption that one of the Ladyženskaja–Prodi–Serrin conditions is satisfied Reference 75Reference 84Reference 85Reference 109Reference 113Reference 120. However, as opposed to Leray–Hopf weak solutions, for the weak solutions of Definition 2.2 it is by now known that uniqueness fails. This was established by the authors of this review in Reference 23 (see Theorem 4.1 below) and then improved in a joint work with Colombo Reference 16 showing that uniqueness fails even if partial regularity in time holds (see Theorem 4.3 below). The proofs in Reference 16Reference 23 rely on a version of the convex integration scheme, called intermittent convex integration, which will be discussed in Section 4 below.

2.1.2. Conditional proof of energy balance

We return to a question alluded to earlier: Consider a weak/mild solution of the Navier–Stokes equations which is known to lie in the Leray–Hopf energy space . Does such a solution automatically obey the energy equality/balance Equation 2.3? Currently, this question is open and only conditional criteria are available. The issue lies in justifying the cancellation Equation 2.4 in the sense of temporal distributions. The classical results state that if in addition to one also knows that Reference 118 or, more generally, with Reference 151, then Equation 2.3 holds; see also Reference 32Reference 33Reference 110Reference 116 for more recent refinements. We emphasize that these conditional results assume less integrability on than the Ladyženskaja–Prodi–Serrin conditions, and this is a common theme in hydrodynamics: the energy equality may be justified under much weaker conditions than those required for establishing the uniqueness of solutions, i.e., the energy equality is a weaker notion of rigidity than uniqueness Reference 104.

We recall a modern proof of the classical result of Reference 118Reference 151, namely that implies Equation 2.3; for details we refer the reader to Reference 29Reference 33Reference 154. The point is that by interpolation, the assumption on implies that , which means that . The former condition is known due to Reference 29Reference 68 to imply Equation 2.3. The more detailed argument is as follows.

Denote by the projection of to its Fourier frequencies which have modulus less than . Testing the Navier–Stokes equations with the smooth divergence free function , and using that is self-adjoint on and commutes with space and time derivatives, similarly to Equation 2.3 we obtain that

where we have denoted by the energy flux through frequencies by, i.e.,

First, we note that as soon as , by the dominated convergence theorem, the terms on the left side of Equation 2.7 converge as to the terms on the left side of Equation 2.3, when integrated in time. Second, if , then also the first term on the right side of Equation 2.7 converges to the term on the right side of Equation 2.3, when integrated in time.

The only question that remains is whether the time integrated flux term in Equation 2.7, , vanishes as . In order to address this, one first uses that is smooth, and thus the cancellation Equation 2.4 holds with replaced by . This allows one to rewrite

which reveals the importance of the quadratic commutator term

As was shown by Constantin, E, and Titi in Reference 41, we have

This may be combined with the bound to yield

which proves that as soon as for any , then as , thereby proving that the energy equality holds. In fact, this is exactly the proof given in Reference 41 for the rigid side of Onsager’s conjecture; we discuss this in Section 2.2 below.

Returning to our goal of proving the energy equality for the Navier–Stokes system Equation 2.3, we note that the assumption was not yet used. Using standard interpolation inequalities, the information that gives that and thus that and , where ; i.e., lies in the space . This information is not good enough to apply Equation 2.11 since ; it is however just good enough to prove that the flux vanishes as . To see this, we recall a more detailed estimate on the commutator term , and thus for the flux , which was obtained in Reference 29 by using the Bony paraproduct decomposition from Littlewood–Paley analysis:

Besides showing that the energy transfer from one dyadic scale to another is mainly local, the above estimate shows that implies , thereby completing the conditional proof of Equation 2.3. We note that estimate Equation 2.12 gives the best-known condition on which ensures that the total energy flux vanishes, and this condition is sharp in the case of the 1D Burgers equation Reference 154. We revisit these ideas in Section 2.2.

2.1.3. Anomalous dissipation of energy in the infinite Reynolds number limit

We have seen in the previous section that the energy equality Equation 2.3 is not necessarily satisfied for weak solutions of the Navier–Stokes equations, even if they lie in the Leray–Hopf space. The identity Equation 2.7 does however show upon passing that

holds in the sense of distributions in time, and may or may not be equal to . It is thus natural to define the average energy dissipation rate on the time interval as

It is clear from the above definition that if the Navier–Stokes solutions maintain a certain degree of regularity uniformly in , for instance if at fixed we have and if when , then for any we have .

The vanishing of the energy dissipation rate for turbulent solutions as (equivalently, as the Reynolds number goes to ) contradicts an experimental fact known as the zeroth law of turbulence, which loosely speaking states that⁠Footnote5

5

Here, as in laboratory experiments, we consider the long (infinite) time average of the energy dissipation rate, , to signify that the solution has reached a stationary regime. For the sake of brevity we avoid the important and subtle discussion about ensemble averages with respect to probability measures on that encode the statistics of the flow Reference 79Reference 80, and about the impromptu ergodic hypothesis which is classical in statistical mechanics. We refer the interested reader to Reference 82 or other texts in turbulence.

for an . The anomalous dissipation of energy postulated in Equation 2.15 is the fundamental ansatz of Kolmogorov’s 1941 theory of fully developed turbulence Reference 105Reference 106Reference 107, and it has been verified experimentally to a tremendous accuracy Reference 99Reference 141Reference 158. The literature on this topic is vast, and we refer the interested reader to Reference 22Reference 74Reference 82Reference 154 for references.

To date it remains open to prove that Equation 2.15 is a manifestation of the statistical behavior of solutions to the Navier–Stokes equations in the infinite Reynolds number limit. Nonetheless, Equation 2.15 provides unquestionable physical evidence that in the limit , one should expect that Navier–Stokes solutions do not remain uniformly smooth, and that (at best) they converge to nonsmooth, possibly nonunique, weak solutions of the Euler equations. Thus, in an attempt to translate predictions made by turbulence theories into mathematically rigorous questions, such as Onsager’s conjecture which we discuss next, it is natural to work within the framework of weak solutions of the Euler equations.

2.2. The Euler equations and Onsager’s conjecture

The classical model for the motion of an incompressible homogenous inviscid fluid are the Euler equations, obtained by formally letting in Equation 2.1

As with the Navier–Stokes equations, the literature concerning the incompressible Euler equations is immense, and we refer the interested reader to the books Reference 28Reference 119Reference 124Reference 125 for an overview of the standard results in the field.

2.2.1. The energy equality and weak solutions

In direct correspondence to the discussion presented in Section 2.1.1 for the Navier–Stokes equations, one may show that smooth solutions of the Euler equations conserve their kinetic energy

Indeed, Equation 2.17 is nothing but Equation 2.3 with and . Thus, any reasonable notion of solution to the Euler equations should at least have a finite norm.

On the other hand, as mentioned in Section 2.1.3 the zeroth law of turbulence motivates the study of weak solutions to the Euler equations, defined as:

Definition 2.3 (Weak solution).

A vector field is called a weak solution of the Euler equations on an open if for any the vector field is weakly divergence free, has zero mean, and satisfies Equation 2.16 in the sense of distributions; that is,

holds for any divergence free test function . The pressure can be recovered by the formula with of zero mean.

2.2.2. Onsager’s conjecture and its variants

The validity of Equation 2.17 for weak (instead of smooth) solutions of the Euler equations is the subject of Onsager’s conjecture, one of the most celebrated connections between phenomenologies in turbulence and the rigorous mathematical analysis of PDEs arising in fluid dynamics. In Reference 138, Onsager considered the possibility that energy dissipation in the infinite Reynolds number limit is not caused by a remnant of viscous effects (i.e., from the term present on the right side of Equation 2.13), but instead, because the solutions of the limiting equation at , namely the Euler equation, are not sufficiently smooth to ensure that . As explained in Reference 73, the argument that the energy equality Equation 2.17 holds for a finite energy solution of the Euler equations if and only if the total energy flux vanishes (i.e., when time integrated) is essentially already contained in Reference 138. Onsager’s remarkable analysis (see also Reference 74) went further and made a precise statement about the threshold regularity of which is necessary in order to justify Equation 2.17; in modern mathematical terms Onsager’s conjecture⁠Footnote6 is now a theorem due to Reference 20Reference 41Reference 93:

6

For a discussion of Onsager’s conjecture for the 2D Euler equations, we refer the reader to Reference 29Reference 30Reference 44.

Theorem 2.4 (Onsager’s conjecture in Hölder spaces).
(a)

Any weak solution belonging to the Hölder space for conserves its kinetic energy.

(b)

For any there exist weak solutions which dissipate kinetic energy, i.e., the kinetic energy is a nonincreasing function of time.

The rigidity, part a of Theorem 2.4, is discussed in Section 2.2.3, while the flexibility, part b, is presented in Section 3.

Remark 2.5 (Onsager’s conjecture on other Banach scales).

In Theorem 2.4, the threshold between rigidity and flexibility is measured here on the Hölder scale, with threshold value . However, since the energy flux is trilinear in (see Equation 2.8), it is most natural to measure this dichotomy on an based space, such as the scale of spaces . The bound Equation 2.11 again suggests that the threshold value on the -based Banach scale is , and indeed the proof of Theorem 2.4 establishes this fact. Lastly, we note that the threshold between rigidity a and flexibility b may alternatively be measured on based spaces in , such as the , with . These spaces are classically related to Fourier analytic measurements of energy spectra (such as the Kolmogorov–Obhukov power spectrum) and to the scaling of second-order structure functions in turbulent flows Reference 82. On this Sobolev scale, however, the value of the threshold exponent is unclear. It is known for a while Reference 32Reference 160 that the kinetic energy is conserved if with . However, as a byproduct of the proof of the flexibility part b of Theorem 2.4, we only have the existence of weak solutions which violate the energy equation for . Based on physical considerations and on the experimentally measured anomalous scaling of second-order structure functions in fully developed turbulent flows Reference 82, it is safe to conjecture that the threshold exponent on the is strictly larger than ; it is, however, not clear whether it should be equal to or another smaller value.⁠Footnote7 Conservatively, we conjecture (see also Open Problem  in Reference 22) that Onsager’s threshold exponent on the based Sobolev scale is strictly larger than :

7

For a dyadic shell model of the 3D Euler equations, for which the energy flux is unidirectional, it was shown in Reference 31 that the threshold is sharp.

Conjecture 2.6 (Deviation from the Kolmogorov–Obhukov power spectrum).

There exists an and infinitely many weak solutions of the 3D Euler equations Equation 2.16, which dissipate kinetic energy.

2.2.3. The proof of rigidity in Onsager’s conjecture

Part a of Onsager’s conjecture was partially established by Eyink in Reference 73 and later proven in full by Constantin, E, and Titi in Reference 41. Indeed, the commutator bound Equation 2.11 established in Reference 41 shows that if with , then automatically by embedding we have , and so

thereby proving Equation 2.17. The argument in Reference 41 was further refined in Reference 68, where the energy dissipation measure was introduced, in Reference 29 where the bound Equation 2.12 is proven, and in the more recent paper Reference 156, which discusses several geometric constraints which ensure the conservation of energy for the threshold value . We note that the proof of rigidity on the based scale (i.e., for with ) also follows from the bound Equation 2.12 established in Reference 29 by additionally using the Bernstein inequality .

2.2.4. Helicity

We mention that besides the kinetic energy , the 3D Euler system also posses one more nontrivial invariant Reference 71 which is not coercive but has deep geometric meaning Reference 131; this is the helicity

where is the vorticity. Here we use the generalized helicity notation Reference 4

which will appear several times in Section 2.3. As opposed to the kinetic energy, which is well defined for weak solutions in the sense of Definition 2.3, the fluid helicity requires a minimal regularity of to be well defined, via the duality pairing between and . However, it is not known whether for any weak solution we have

Due to Reference 29, we know that if , then the fluid helicity for that weak solution is a constant function of time. In terms of rigidity, this condition is expected to be sharp. An alternative condition in which velocity and vorticity obey different assumptions was obtained in Reference 63.

On the other hand, in terms of flexibility we note that for the weak solutions constructed for part b of Theorem 2.4, the helicity is not well defined, as does not embed into . In fact, to date the following question remains widely open:

Conjecture 2.7 (Helicity flexibility).

There exists a weak solution of the 3D Euler equations Equation 2.16, with for some , such that the helicity is not a constant function of time.

The difficulty is that on the one hand for the weak solutions obtained from the convex integration constructions in Section 3, the kinetic energy can be always made to be a nonconstant function of time; on the other hand, if one works on the Hölder scale and wishes to have a well-defined helicity, then one has to take ; but in turn this implies that the kinetic energy must be conserved due to part a of Theorem 2.4. It is conceivable that in order to attack Conjecture 2.7 one has to work on the based Sobolev scale , and that the intermittent convex integration which we will describe in Section 4 should be employed instead. Even so, Conjecture 2.7 is strictly harder than Conjecture 2.6.

2.3. The MHD equations and Taylor’s conjecture

The incompressible MHD equations are the classical macroscopic model coupling Maxwell’s equations to the evolution of an electrically conducting incompressible fluid Reference 12Reference 52. The unknowns are the velocity field , the magnetic field , and the scalar pressure , which we take to have zero mean on . The ideal (i.e., inviscid and nonresistive) version of these equations is

The viscous ( and resistive () MHD equations are given by

Setting in Equation 2.19 we recover the Euler equations Equation 2.16, while letting in Equation 2.20 we arrive at the Navier–Stokes equations Equation 2.1. As such, the local-in-time theory for smooth solutions of the MHD equations, as well as the global-in-time theory for weak solutions, closely mimics the one for the Euler and Navier–Stokes systems; see Reference 69Reference 148.

2.3.1. Ideal MHD conservation laws and weak solutions

The ideal MHD equations Equation 2.19 posses a number of global invariants Reference 101, two of which are obtained via a direct analogy with the Euler equation, namely the total energy and the cross helicity, and one more which is intrinsic to the Lie transport of the magnetic field Equation 2.19a, namely the magnetic helicity. The fact that these conservation laws are not defined at the same level of spatial regularity, makes the analysis in a sense more challenging than for 3D Euler.

The total energy

gives the only known coercive conserved quantity (and in fact the Hamiltonian Reference 101Reference 173 of the system), for smooth solutions of Equation 2.19. In order to verify that is a constant function of time, it is convenient to rewrite Equation 2.19 in terms of the Elsässer variables

so that the system Equation 2.19 becomes

Testing the momentum equation for with , integrating over , and using that both and are incompressible, similarly to Equation 2.17 we obtain that if (and thus ), then

On the other hand, from Equation 2.22 we have that

which combined with Equation 2.24 shows that is conserved for smooth solutions.

Besides the total energy, the MHD system possesses one more Elsässer invariant Reference 2, the cross helicity

Here and throughout the paper we use the notation in Equation 2.18. Again, Equation 2.24 shows that is conserved for smooth solutions.

We note that both and are bounded functions of time as soon as , and in view of the positivity of the total energy, any meaningful notion of solution for Equation 2.19 should at the very least assume this amount of regularity. This motivates the notion of weak solution for the ideal MHD system (in analogy to Definition 2.3 for Euler):

Definition 2.8 (Weak/distributional solution).

We say is a weak solution of the ideal MHD system Equation 2.19 on an open interval , if for any , the vector fields and are divergence free in the sense of distributions, they have zero mean, and Equation 2.19 holds in the sense of distributions; i.e.,

hold for all divergence free test functions .

The validity of Equation 2.24 for weak solutions of the ideal MHD system gives rise to an Onsager-type analysis, which we discuss in Section 2.3.2 below.

As alluded to at the beginning of the section, the ideal MHD system has one more conservation law, the magnetic helicity, which is defined as

(see Reference 131Reference 177Reference 178), where is a vector potential for , i.e., a zero mean periodic field such that . As we work on the simply connected domain and is incompressible, the value of is independent of the gradient part of ; thus may be chosen without loss of generality such that , so that . Thus, the definition Equation 2.27 is consistent with the definition Equation 2.18 introduced earlier.

In order to see that for smooth solutions of Equation 2.19 we have

one may proceed as follows. Since the Biot–Savart operator is self-adjoint, and since for divergence free and we may rewrite , we have

In the second to last equality we have used the identity .

We emphasize that as opposed to the total energy and the cross helicity which require in order to be well defined, the magnetic helicity is well defined as soon as , a negative level of regularity. This difference is manifested in the context of reconnection events in magneto-hydrodynamic turbulence, via a phenomenon whose mathematical aspects are described by Taylor’s conjecture Reference 11Reference 132Reference 168Reference 169, see Section 2.3.4 below.

Remark 2.9 (2D Euler and Surface Quasi Geostrophic).

This situation encountered here in which the hydrodynamic model has conservation laws at different levels of regularity ( vs ) is not uncommon. Another occurrence is in the context of the 2D Surface Quasi Geostrophic (SQG) and 2D Euler equations. For both of these equations smooth solutions conserve the norm (in fact all the norms with ) of the temperature (vorticity for the potential velocity Reference 21) for SQG (resp., the scalar vorticity for 2D Euler). With respect to these so-called Casimirs, the respective Hamiltonians of these systems lie at a negative level of regularity: for SQG (resp., for 2D Euler). As such, much of the discussion presented in this paper concerning the MHD system has an analogue in the context of the SQG Reference 21Reference 94Reference 96 and 2D Euler equations Reference 29Reference 30.

2.3.2. Onsager-type dichotomies

Similarly to Onsager’s conjecture for the 3D Euler equations, it is natural to analyze the regularity threshold at which weak solutions of Equation 2.19 (in the sense of Definition 2.8) respect the ideal MHD conservation laws for the energy , the cross helicity , and the magnetic helicity Reference 1Reference 2. Given a Banach scale used to measure the regularity of weak solutions, we wish to identify a critical/threshold exponent, above which all weak solutions obey the given conservation law (rigidity), while below this exponent there exist weak solutions which violate it (flexibility).⁠Footnote8

8

See Reference 104, where this question is posed for general nonlinear, supercritical, Hamiltonian evolution equations.

Conjecture 2.10 (Onsager-type conjecture for the Elsässer energies).
(a)

Any weak solution of the ideal MHD system Equation 2.19 belonging (resp., for , conserves the total energy and the cross helicity .

(b)

For any there exist weak solutions (resp., , which dissipate the total energy , and for which the cross helicity is not a constant function of time.

The statement of the corresponding dichotomy for the magnetic helicity is slightly modified, to avoid spaces of negative regularity, which are not consistent with Definition 2.8. In particular, in the flexibility statement, we relax the integrability exponent not the regularity assumptions of the weak solutions.

Conjecture 2.11 (Onsager-type conjecture for the magnetic helicity).
(a)

Any weak solution of the ideal MHD system Equation 2.19 belonging to (resp., conserves the magnetic helicity .

(b)

For any , there exist weak solutions (resp., for which the magnetic helicity is not a constant function of time.

The rigidity parts of the above conjectures are discussed in Section 2.3.3 below, while progress towards the flexibility parts is outlined in Section 5. As opposed to Onsager’s conjecture for 3D Euler, which is by now a theorem, the flexibility statements part b of Conjecture 2.10 and part b of Conjecture 2.11 remain to date open. We only mention at this stage that for and any as constructed in part b of Theorem 2.4, the resulting pair is a weak solution with the regularity required by the flexible part of Conjecture 2.10, and for which the total energy is dissipated. However, for any such solution the cross helicity and the magnetic helicity are trivial since is trivial (i.e., identically equal to zero) and thus are constants in time.

2.3.3. The proof of rigidity in Onsager-type conjectures for MHD

We start by discussing part a of Conjecture 2.10. Recall that the conservation of and is equivalent to the conservation of the Elsässer energies, i.e., the validity of Equation 2.24. Inspecting the momentum equation in Equation 2.23, we see that the only difference to 3D Euler is that appears as the transport velocity for the evolution. With this modification in mind, one may define two fluxes (for the evolution) and (for the evolution) in analogy to Equation 2.8. The energy equality again boils down to whether or not. The former question has been resolved by repeating the argument in Reference 45, which gives a bound for the commutator as in Equation 2.10. This proof was carried through in Reference 26, where it is shown that and thus the conservation of energy and cross helicity holds for any weak solutions with . We also refer the reader to Reference 100 where the methods of Reference 29 and used to establish the endpoint case for rigidity in Conjecture 2.10, namely .

Concerning the rigidity part a of Conjecture 2.11, we note that in the integrand of Equation 2.29 is zeroth order in and . This suggests that justifying Equation 2.29 for weak solutions of Equation 2.19 should require less regularity on and than was required in order to justify Equation 2.24. Indeed, it is shown in Reference 26 that is conserved by weak solutions of Equation 2.19 as soon as with , so that the threshold regularity is . The idea is as follows. We want to use a version of Equation 2.29, with replaced by . Then, the conservation in time of is equivalent to the vanishing as (integrated in time) of the magnetic flux term, defined in analogy to Equation 2.8 as

Using the commutator estimate of Reference 41 (see Equation 2.10), one then immediately obtains

which concludes the proof when , upon integrating in time and passing . The endpoint case was reached in Reference 1Reference 76Reference 100, where it is proven that magnetic helicity is conserved for any weak solution (in the sense of Definition 2.8) as soon as . The slightly sharper statement concerning rigidity for the endpoint Besov space stated in part a of Conjecture 2.11 may be achieved by repeating the argument of Reference 29 to establish a bound for the magnetic flux at dyadic scales: , which is nearly the same as Equation 2.12, except that the term on the right side is absent.

We note that for the 2D MHD equations much stronger types of rigidity may be established (when compared to the 3D case discussed in this paper), and we refer interested readers to Reference 76Reference 77Reference 78.

2.3.4. Taylor’s conjecture for weak ideal limits

While turbulent low-density plasma configurations are observed to dissipate the total energy Reference 50Reference 127, it is commonly accepted knowledge in the plasma physics literature, and an experimental fact, that the magnetic helicity is conserved Reference 139 in the infinite conductivity limit. This striking phenomenon manifests itself mathematically as Taylor’s conjecture Reference 11Reference 132Reference 168Reference 169. Here we discuss its rigorous foundations, following Reference 26Reference 77.

We start from the viscous and resistive MHD system Equation 2.20, where . In analogy to the energy inequality Equation 2.5 for the Navier–Stokes equations, sufficiently smooth solutions of Equation 2.20 satisfy the following energy inequality for the total energy defined in Equation 2.21:

for . Based on Equation 2.30 it is classical Reference 69 to build a theory of Leray–Hopf weak solutions for Equation 2.20; these are solutions with which obey Equation 2.30 for a.e. and all . In physically realistic regimes, we are interested in ; however, in the ideal limit , Equation 2.30 only gives bounds for the norms of and . Following Reference 77, Definition 1.1 we recall the definition:

Definition 2.12 (Weak ideal limit).

Let be a sequence of vanishing viscosities and resistivities. Associated to a sequence of divergence free initial data converging weakly in , let be a sequence of Leray–Hopf weak solutions of Equation 2.20. Any pair of functions such that in , are called a weak ideal limit of the sequence .⁠Footnote9

9

Note however that a weak ideal limit need not be a weak solution of the ideal MHD equations in the sense of Definition 2.8, for the same reason that a vanishing viscosity limit of Leray–Hopf weak solutions of the 3D Navier–Stokes equations need not be a weak/distributional solution of the 3D Euler equations; only measure valued solutions are known to be achieved as subsequential limits Reference 66.

Taylor’s conjecture states that weak ideal limits of MHD Leray–Hopf weak solutions conserve magnetic helicity. This was proven recently in Reference 77, and we recall the statement:

Theorem 2.13 (Taylor’s conjecture; Theorem 1.2 in Reference 77).

Suppose is a weak ideal limit of a sequence of Leray–Hopf weak solutions. Then is a constant function of time. In particular, finite energy weak solutions of the ideal MHD equations Equation 2.19, which are weak ideal limits, conserve magnetic helicity.

At the heart of the proof of Theorem 2.13 given in Reference 77 (where they also consider domains that are not simply connected) lies the observation that for , by interpolation we have that MHD Leray–Hopf weak solutions lie in , which is a stronger space than the which is required in part a of Conjecture 2.11. Thus, in analogy to the proof of Equation 2.3 discussed earlier, at fixed one may justify the magnetic helicity equality

The bounds provided by Equation 2.30 immediately imply that the second term on the left side of Equation 2.31 may be bounded by , and thus this term vanishes in the ideal limit . The proof of Theorem 2.13 is then concluded by showing that due to the compactness of the embedding for the subsequential limits in Definition 2.12 we have strongly in , and so strongly in . By the definition of magnetic helicity in Equation 2.27, this information is sufficient to show that uniformly in time, and thus letting in Equation 2.31 concludes the proof.⁠Footnote10

10

We note that similar arguments were previously used to prove the conservation of the Hamiltonian for the 2D Euler Reference 30 and 2D SQG equations Reference 43, for weak solutions which arise from vanishing viscosity limits.

Remark 2.14.

In closing, we note that Theorem 2.13 does not contradict part b of Conjecture 2.11. Theorem 2.13 does indeed show that if is a weak ideal limit (cf. Definiton 2.12) and also a weak solution of the ideal MHD equations (cf. Definiton 2.8), then it conserves magnetic helicity, although it may have regularity as weak as . This is far less spatial integrability than the condition required in part a of Conjecture 2.11, which points to the fact that weak ideal limits are seeing a ghost of the energy inequality Equation 2.30. Nonetheless, the flexible part of Conjecture 2.11 is that there may exist weak solutions of ideal MHD which do not arise in the vanishing viscosity/resistivity limit, leaving open the possibility that for such solutions is not constant in time. This specific result is established in Theorem 5.2 below.

3. Convex integration and Nash schemes for the Euler equations

While the rigid part of Onsager’s conjecture (part a of Theorem 2.4) has been essentially understood since the 1990s, systematic progress towards the resolution of the flexible part b did not occur until the 2010s and the groundbreaking works Reference 55Reference 58 of De Lellis and Székelyhidi Jr. These works have developed the mathematical framework of the and the flavors of convex integration in fluid dynamics, and have laid out some of the key ideas which have eventually led to the solution of the flexible part of Onsager’s conjecture by Isett Reference 93 (in the context of solutions with compact support in time); and in a subsequent work by Buckmaster, De Lellis, Székelyhidi, and Vicol Reference 20 (for dissipative weak solutions). This sequence of developments has already been discussed in great detail in the review papers Reference 22Reference 57Reference 60Reference 61Reference 162, so in this section, we only give a succinct presentation. We emphasize that the convex integration scheme used to prove Theorems 3.1 and 3.2 has a number of key similarities to, but is also conceptually different from, the Nash-type convex integration method used to prove Theorems 3.33.4, and 3.5.

3.1. The flexible part of Onsager’s conjecture: First paradoxical examples

In the seminal work Reference 147, Scheffer demonstrated the existence of nontrivial weak solutions of the 2D Euler system Equation 2.16, which lie in and have compact support in time and space! Strictly speaking the weak solutions of Scheffer are not dissipative, as dissipative solutions are required to have nonincreasing energy; nonetheless, Reference 147 is considered to be the first result concerning the flexible part b, of Onsager’s conjecture.

A different construction of a nontrivial weak solution to the 2D Euler equations, which are periodic in space and have compact support in time, was given by Shnirelman in Reference 153. The existence of dissipative weak solutions to the Euler equations was first proven by Shnirelman in Reference 152, where he constructs weak solutions which lie in .

These results, which were initially referred to as the Scheffer–Shnirelman paradox, represent not just a proof of nonuniqueness for weak solutions to the Euler equations, but a drastic failure of determinism within this class of solutions.

3.2. The results

The first example of a bounded in space and time, dissipative weak solution of the Euler equations (one for which the kinetic energy is a nonincreasing function of time), in any dimension , was obtained in a groundbreaking work by De Lellis and Székelyhidi Jr. Reference 55. Their main result is:

Theorem 3.1 (Theorem 4.1 in Reference 55).

For any open bounded space-time domain , there exists a weak solution of the Euler equations Equation 2.16 , in the sense of Definition 2.3, such that for a.e. , and for a.e. . Moreover, there exists a sequence of functions such that:

, and ,

in as ,

is uniformly bounded in ,

in for any .

The first part of the theorem establishes the existence of a weak solution which is compactly supported in space and time, while the second part is a manifestation of the proof: the limiting weak solution is obtained from a smooth approximating sequence , which solves a relaxed Euler system, whose right side vanishes in a weak sense as . This paper also introduced the ideas of a subsolution and of a Reynolds stress for the Euler system Equation 2.16. Maybe more important than the result itself, which was later improved by the same authors, is the fact that Reference 55 relates the construction of paradoxical weak solutions of the Euler equations with a classical technique in geometry, convex integration, and the notion of -principles for soft partial differential equations.

The method of convex integration can be traced back to the work of Nash, who used it to construct exotic counterexamples to the isometric embedding problem Reference 135. The method was later refined by Gromov Reference 86, and it evolved into a general method for solving soft/flexible geometric partial differential equations Reference 70. In the influential paper Reference 134, Müller and Šverák adapted convex integration to the theory of differential inclusions (see also Reference 102), leading to renewed interest in the method as a result of its greatly expanded applicability. Inspired by the works Reference 102Reference 134, and building on the plane-wave analysis introduced by Tartar Reference 165Reference 166 and Di Perna Reference 64, De Lellis and Székelyhidi Jr., in Reference 55, applied convex integration in the context of bounded weak solutions to the Euler equations.

We refer the interested reader to the review papers Reference 57Reference 61Reference 162 for a detailed discussion connecting convex integration in the context of differential inclusions, and also -principles, to the type of constructions that were initiated by Reference 55. In particular, analyzing the toy example presented in Reference 57, Section 5.1 is particularly accessible, yet insightful.

The work Reference 55, has since been extended and adapted by various authors to various problems arising in mathematical physics, including Reference 6Reference 27Reference 36Reference 37Reference 38Reference 39Reference 46Reference 47Reference 56Reference 155Reference 161Reference 163Reference 174, just to mention a few. Here we single out the work Reference 56 which considers the question of whether imposing additional admissibility criteria on the weak solutions of the Euler equations could rule out the construction of examples such as those in Theorem 3.1. Physically motivated admissibility criteria, based on energetic arguments such as those discussed in Section 2 (ordered by least restrictive to most restrictive) include the following.

(i)

Weak energy inequality: the kinetic energy satisfies for all .

(ii)

Strong energy inequality: the kinetic energy satisfies for all .

(iii)

Local energy inequality: the distribution , for , is nonnegative.

The main result of Reference 56 may be summarized as follows.

Theorem 3.2 (Theorem 1 in Reference 56).

For any dimension , there exists bounded, compactly supported divergence-free initial data , for which there exist infinitely many weak solutions of the Euler equations, such that the admissibility conditions i, ii, and iii hold.

This result shows that “wild” weak solutions of the Euler equations (as constructed by Theorem 3.1 and Theorem 3.2) cannot be ruled by the local energy inequality. Indeed, the Euler equations are not scalar conservation laws!

3.3. The result: A Nash-type convex integration scheme

The constructions described in Section 3.2 are based on writing the Euler equations as a differential inclusion, and then applying a machinery from Lipschitz differential inclusions, which either uses a Baire-category argument or, equivalently, an explicit convex integration approach. These methods face a serious difficulty in constructing continuous weak solutions of Equation 2.16, since it seems impossible to extract a uniform continuity estimate for approximating sequences . The breakthrough was made by De Lellis and Székelyhidi Jr. in their seminal papers Reference 58Reference 59, where they developed a new convex integration scheme, motivated by and resembling in part the earlier schemes of Nash and Kuiper Reference 108Reference 135. In Reference 58, De Lellis and Székelyhidi Jr. prove the existence of continuous weak solutions to the Euler equations satisfying a prescribed kinetic energy profile, which in particular may be decreasing:

Theorem 3.3 (Theorem 1.1 in Reference 58).

Assume is a smooth function. Then there is a continuous vector field and a continuous scalar field which solve the incompressible Euler equations Equation 2.16 in the sense of distributions, and such that

for all .

The proof of Theorem 3.3 departs from the arguments based on functional analysis, which were used to construct bounded weak solutions, and implements a hard analysis scheme, in which the constructions of the building blocks are not plane waves anymore, instead they are adapted to the geometry of steady states of the Euler equations (Beltrami flows), and the estimates involve precise singular integral bounds and Schauder estimates.

3.3.1. Overview of the proof of Theorem 3.3 and of Nash-type convex integration schemes

The summary given here is similar to the one we have given in Reference 22, Section 4.1.

For each by induction one constructs smooth functions , which solve the Euler–Reynolds system

The pressure is always given by . The stress is symmetric and has zero trace (since its trace is absorbed in to the pressure term). The goal is to construct the sequence such that converges uniformly to as , and that at the same time the sequence converges uniformly to a Hölder continuous weak solution to the Euler equations, which satisfies Equation 3.1. The iterates constructed via the convex integration scheme are approximately the spatial averages of the final solution at length scales , which are decreasing with . In view of the analogy to theories in fluid turbulence, one refers to the symmetric tensor as the Reynolds stress.

At each inductive step, the goal is to design a perturbation

such that the new velocity and pressure solve the Euler–Reynolds system Equation 3.2 at level , but with a smaller Reynolds stress . Subtracting the equations for and , we obtain the following decomposition of the Reynolds stress at level :

Note that not all the terms on the right side of Equation 3.4 are written in divergence form, which necessitates the use of a negative one order linear Fourier multiplier operator which formally acts as and outputs symmetric traceless matrices.⁠Footnote11

11

For of zero mean one may define

For an increasing sequence of frequency parameters , the approximate solutions at level are essentially localized at Fourier frequencies . On the other hand, the perturbation is constructed as a sum of highly oscillatory building blocks (denoted by in Equation 3.5 below) which live at the high frequency .

Roughly speaking, the principal part of the perturbation, which we label as , will be of the form

where ranges over a finite set, represents the building blocks oscillating at frequency , and the coefficient functions are chosen such that

Here denotes the trace-free part of the tensor product. The amplitude functions are designed in order to obtain a cancellation between the low frequencies of the quadratic term and the old Reynolds stress error , thereby reducing the size of the low frequency part of the oscillation error. More precisely, if we take into account that are chosen to have no low frequency (meaning ) contribution when , the need to minimize the low frequency part of the oscillation error, , dictates the choice Equation 3.6.

To leading order, with respect to the large parameter , the zero mean -periodic building blocks are chosen to be solutions of the stationary Euler equations, i.e., they satisfy for a suitable pressure , and . The importance of this choice becomes apparent when one takes into account the ansatz Equation 3.5, and uses it to compute the high frequency part of the oscillation error on the right side of Equation 3.4, namely . The building blocks used in Reference 58Reference 59, are the so-called Beltrami waves, which are families of complex eigenfunctions of the operator at the same eigenvalue, . Starting with Reference 51, the later works Reference 20Reference 92Reference 93 use the so-called Mikado flows, which are straight pipe flows with pairwise disjoint supports (see Section 3.5). These building blocks are used in an analogous fashion to the Nash twists and Kuiper corrugations employed in the embedding problem Reference 108Reference 135.

The principal part of the perturbation presented in Equation 3.5 needs to be modified in order to minimize the transport error in Equation 3.4, i.e., to ensure it is the divergence of a small Reynolds stress. This is achieved by flowing the building blocks along the ODE flow generated by (we will return to this issue in Section Equation 3.4). Additionally, in order to ensure that is divergence free, one introduces a corrector which ensures that is divergence free. The size of this incompressibility corrector is much smaller than the size of , roughly by a factor of , because the building blocks are divergence-free by definition, and the oscillate at the old frequency, .

In order to ensure that the inductive scheme converges to a Hölder continuous velocity with Hölder exponent , the amplitude of the perturbation is required to satisfy the bound

for some . Here, we note that it is convenient to use a superexponentially growing sequence of frequencies which obeys , for some . In view of Equation 3.6, this necessitates that the Reynolds stress obeys the estimate

Consistent with the definition

and with the bound Equation 3.7, the scheme of Reference 58 also propagates the estimate

It is not hard to see that if the bounds Equation 3.7Equation 3.10 are propagated throughout the scheme, then as we have that uniformly, where is a Hölder continuous weak solution of the Euler equations. Indeed, for any , the following series of increments is summable:

where the implicit constant is universal. Thus, we may define a limiting function which lies in . Moreover, is a weak solution of the Euler equation Equation 2.16, since by Equation 3.8 we have that in , and strongly in .

The main work is now to prove that for a velocity perturbation of the form Equation 3.5, and with amplitude functions that satisfy Equation 3.6, the bounds stated in Equation 3.7Equation 3.10 are indeed attainable inductively for all . We note that if the building blocks are normalized so that , then it follows from Equation 3.5Equation 3.6 and Equation 3.8 that the principal part of the velocity increment already satisfies the bound Equation 3.7. Since the incompressibility corrector is even smaller, Equation 3.7 is expected to hold. The difficult part is to prove Equation 3.8. In view of Equation 3.4, this amounts to bounding the oscillation error, the transport error, and the Nash error. This is the hard analysis part of the construction.

As a demonstration of the typical scalings present in convex integration schemes for the Euler equations, let us consider the Nash error. Heuristically, since is of frequency for every , and by appealing to Equation 3.9, we have that lives at frequency , and thus

where we recall that is a order Fourier multiplier which inverts the divergence operator. Applying Equation 3.7 and Equation 3.10, for we obtain

by using that . Thus, in order to ensure that satisfies the bound Equation 3.8 with replaced by , we require that for and we have

Thus, from this simple heuristic, we see that if is taken to be arbitrarily close to , then the Hölder regularity exponent may be taken to be arbitrarily close to the Onsager-critical Hölder regularity exponent, i.e., .

The construction described above provides a clear enemy towards reaching the desired Onsager threshold: the transport and oscillation errors in Equation 3.4. Designing a which minimizes these two errors simultaneously turns out to be a very difficult problem. This realization stimulated a series of advancements, through the works Reference 15Reference 17Reference 18Reference 51Reference 90, in which the authors incorporated more and more of the specifics of the 3D Euler equation into the convex integration scheme (by designing better and ), in order to obtain higher and higher Hölder regularity exponents. We mention a couple of these developments next.

3.4. Climbing the Onsager ladder

The first breakthrough after Reference 58Reference 59 was to produce a dissipative weak solution of the Euler system with a Hölder regularity exponent with . This was achieved by Isett Reference 90 and later simplified in Reference 19 by Buckmater, De Lellis, and Székelyhidi, the two papers resulting in the joint work Reference 17. The main improvement comes from obtaining a better bound for the transport error in Equation 3.4. In the proof of Theorem 3.3 one did not keep track of precise estimates for the material derivative of the Reynolds stress . Here is a mollification of at a length scale which lies in between and .⁠Footnote12 The realization of the schemes is that material derivatives are better behaved than either regular spatial or temporal derivatives: due to classical ODE arguments, material derivatives should cost a factor proportional to the Lipschitz norm of , i.e., in view of Equation 3.10. Compare this with a spatial derivative, whose cost is . Taking advantage of this observation, one can improve the estimate on the material derivative of , and thus improve the bounds for the transport error. Optimizing this new transport bound with the oscillation error yields the improved -Hölder exponent.

12

An inherent issue associated with convex integration schemes is that in order to control th order derivatives of the perturbation , one needs control derivatives on of an order strictly greater than . In order to avoid this loss of derivative, one replaces by a mollified velocity field and the stress by a mollified stress , where the mollification parameter is to be chosen suitably. This argument was already required in the schemes described in Section 3.3.

The -scheme is very versatile, and it was successfully used to construct weak solutions of 3D Euler with compact support on the whole space Reference 95, to establish the nonuniqueness of weak solutions for the 3D quasi-geostrophic equations Reference 137, the 2D SQG equations Reference 21Reference 94, active scalar equations with nonodd constitutive laws Reference 96, and the hypodissipative Navier–Stokes equations Reference 40 (this result was later improved in Reference 62 to take into account the techniques used to prove Theorem 3.5 below).

In Reference 15, the first author noted that one can construct infinitely many weak solutions of Equation 2.16 whose Hölder regularity exponent with respect to the space variable can be taken to be any with , but only almost everywhere in time. This new scheme concentrates the transport and oscillation errors on a zero-measure set of times. By taking advantage of this idea and by using a delicate bookkeeping scheme, Buckmaster, De Lellis, and Székelyhidi Reference 18 constructed nonconservative solutions in the space .

3.5. Resolution of the flexible side of Onsager’s conjecture

The flexible side of Onsager’s conjecture was finally resolved by Isett in Reference 93, who proved the existence of nonconservative weak solutions of 3D Euler in the regularity class , for any :

Theorem 3.4 (Theorem 1 in Reference 93).

For any there exists a nonzero weak solution , such that vanishes identically outside of a finite interval.

The proof of Isett builds upon the ideas in the above mentioned works, and utilizes two new key ingredients. The first, is the usage of Mikado flows which were introduced earlier by Daneri and Székelyhidi Reference 51. These are a rich family of pressureless stationary solutions of the 3D Euler equation (straight pipe flows), which have a better (when compared to Beltrami flows) self-interaction behavior in the oscillation error, when they are advected by a mean flow. The second key ingredient is that prior to adding the convex integration perturbation , it is very useful to replace the approximate solution with another pair , which has the property that is close to , but more importantly that vanishes on every other interval of size within . The velocity field (and consequently also the stress ) is obtained by smoothly gluing together exact solutions of the Euler equations, whose initial data are chosen to precisely match at suitably spaced instances of time. A proto-version of this gluing scheme may already be found in the work of Shnirelman Reference 153, who however works with Dirac masses in time, which produce unacceptable errors. In turn, working with the glued velocity and stress results in a major improvement of the size of the oscillation error in the convex integration step since different Mikado flows have disjoint supports, and thus do not interact on the time scale dictated by the Lipschitz norm of the mean flow .

The weak solutions constructed by Isett Reference 93 are not strictly dissipative. This issue was resolved in the paper Reference 20 by Buckmaster, De Lellis, Székelyhidi, and Vicol, who prove the precise statement of part b of Onsager’s conjecture:

Theorem 3.5 (Theorem 1.1 in Reference 20).

Let be a strictly positive smooth function. For any there exists a weak solution of the Euler equations Equation 2.16, whose kinetic energy at time equals .

We note that in Reference 92, Isett showed that one can further optimize the schemes of Reference 20Reference 93 in order to construct nonconservative weak solutions to the Euler equations that lie in the intersection of all Hölder spaces for .

3.6. Some open problems in the context of 3D Euler

Although the weak solutions constructed in Theorem 3.5 may be constructed to satisfy both the weak i and the strong ii energy inequality mentioned earlier in Theorem 3.2, they do not satisfy the local in space and time version of the energy inequality iii. In this direction, the recent results Reference 91 and Reference 54 achieve the regularity exponents , and , respectively. Extending these results to the full range remains open (see also Reference 22, Problem 3).

It is also an open problem to determine whether there exist nonconservative weak solutions to the Euler equations that have Hölder exponent exactly equal to . Such a result would not be in contradiction with Reference 29 (see also Reference 22, Problem 4).

At the moment of writing of this article, it is not known how to construct nonunique weak solutions of the 3D Euler equations with regularity, for some . It appears that such a result would require fundamental new ideas, beyond the ideas provided by the convex integration schemed described earlier. Lastly, we note that Conjectures 2.6 and 2.7 remain open.

4. Intermittent convex integration for the Navier–Stokes equations

In the previous section we have described various flexibility results obtained for the 3D Euler equations via the convex integration technique and the Nash-type convex integration scheme. Both produce infinitely many weak solutions of Equation 2.16 which are bounded in space and time. For this reason, these methods cannot be used to produce finite energy weak solutions of 3D Navier–Stokes equations (cf. Definition 2.2). Indeed, as discussed in Section 2.1.1, if is a weak solution of Equation 2.1 such that , then it is automatically smooth and thus unique Reference 75. Thus, in order to extend the applicability of convex integration techniques to the Navier–Stokes system requires a new approach.

Building on the Nash-type convex integration method in Reference 58 and drawing inspiration from the experimental reality that fully developed turbulent flows are intermittent⁠Footnote13 Reference 82Reference 133, a new technique (which we call intermittent convex integration) was developed by the authors in Reference 23. In physical space, intermittency causes concentrations that result in the formation of intermittent peaks. In frequency space, intermittency smears frequencies. Analytically, intermittency has the effect of saturating Bernstein inequalities between different spaces Reference 35. In the context of convex integration, intermittency reduces the strength of the linear dissipative term in order to ensure that the nonlinear term dominates. We refer the reader to Reference 22, Sections 2 and 7, where more heuristics (and references) about intermittency are presented and the fine details of the proof in Reference 23 are given.

13

Broadly speaking, intermittency is characterized as a deviation from the Kolmogorov 1941 scaling laws, which were derived under the assumptions of homogeneity and isotropy. Experimentally, it is seen that these assumptions need not hold at large Reynolds numbers. A common signature of intermittency is that the th order structure function exponents deviate from the Kolmogorov predicted value of for Reference 3Reference 99Reference 126Reference 136.

The goal of this section is to show how intermittent convex integration may be used to prove the nonuniqueness of weak/mild solutions (cf. Definition 2.2) to the 3D Navier–Stokes and to present a number of variants and improvements of the method from Reference 23 which were recently obtained in Reference 14Reference 16Reference 48Reference 122Reference 123Reference 129Reference 130. The application of intermittent convex integration to the MHD system is given in Section 5 below.

4.1. Nonuniqueness of weak solutions with finite kinetic energy

In Reference 23, we have proven the existence of infinitely many weak/mild solutions of the 3D Navier–Stokes equations Equation 2.1, in the sense of Definiton 2.2, with a prescribed kinetic energy profile:

Theorem 4.1 (Theorem 1.2 in Reference 23).

There exists , such that for any nonnegative smooth function and any , there exists a weak solution of the Navier–Stokes equations

such that holds for all .

We emphasize that the weak solutions constructed in Theorem 4.1 are not Leray–Hopf weak solutions, whose uniqueness remains famously one of the most challenging questions in fluid mechanics Reference 98.

Theorem 4.1 shows in particular that is not the only weak/mild solution which vanishes at a time slice, thereby implying the nonuniqueness of solutions in the sense of Definiton 2.2 (note that within the class of Leray–Hopf weak solutions is the only solution with datum). Theorem 4.1 also proves that weak solutions may come to rest in finite time, a question posed in Reference 150, pp. 88. Lastly, note that the function may be taken to be monotone decreasing so that rigidity fails for dissipative weak/mild solutions.

4.1.1. Outline of the proof of Theorem 4.1 and of intermittent convex integration schemes

The summary given here is similar to the one the authors gave in Reference 22, Section 4.2. For clarity of the presentation, we omit any discussion of the energy profile , and refer the reader to Reference 23 and Reference 22, Section 7 for details.

The structure of the proof resembles that of Nash-type convex integration schemes, as described in Section 3.3.1. In order to construct weak solutions of Equation 2.1 we proceed inductively, and for every we construct a solution to the Navier–Stokes–Reynolds system,

where the stress is traceless symmetric, and .

As in Section 3.3.1, we work with a superexponentially growing sequence of frequencies which obeys and , except that this time , so we think of the frequency as being much much larger than . Moreover, as in Equation 3.3, we denote the velocity increment by , and we heuristically view as a function whose Fourier support lies inside a ball of radius proportional to around the origin.

From experience with 3D Euler, we expect that has to be made as small as possible, so that if we wish the sequence of approximate velocities to converge strongly in as to a weak solution of the Navier–Stokes equations, then the sequence has to be constructed such that vanishes in as . This is a major difference when compared to the uniform in space convex integration used for the Euler equations: here Reynolds stresses (errors) are measured in , whereas velocity increments and approximate velocities are measured in ; this is essentially because is quadratic in . Thus, in analogy to Equation 3.7 and Equation 3.8, the inductive scheme is designed to propagate the bounds

for some , which will be taken to be very small (in terms of the parameter in ).

In order to streamline the notation, for the remainder of this section we omit the from all norms and simply denote as for any Banach space , because all norms are taken to be uniform in time. From the bound Equation 4.2a and our heuristic about the Fourier support of , we may expect that

which is suggestive of the fact that strongly in for . Moreover, viewing the momentum equation in Equation 4.1 as a forced Stokes system, from the maximal regularity of the Stokes equation we deduce that

with bounds that degenerate as . Thus, letting be sufficiently close to , we may deduce that strongly in for some . Thus, using Equation 4.2a and Equation 4.2b the regularity of the limiting weak solution which we claimed in Theorem 4.1 follows, except that is replaced with (the reader should not be confused by the fact that the in Theorem 4.1 is not the same parameter as the one in Equation 4.2a and Equation 4.2b).

Subtracting the equations Equation 4.1 at level and level , similarly to Equation 3.4, we obtain

In order to cancel the previous stress in the first term on the right side of Equation 4.3, as in the proof of Theorem 3.3, the principal part of perturbation, is taken to be of the form Equation 3.5, where the coefficient functions are chosen to satisfy the low mode cancellation identity Equation 3.6.

The main difficultly in implementing a convex integration scheme for the Navier–Stokes equations is ensuring that the dissipative term on the right side of Equation 4.3 can be treated as an error in comparison to the quadratic term . Note that the building blocks from the Euler construction (in conjunction with the ansatz Equation 3.5) do not have this desirable property, as they satisfy the heuristic pointwise in space bounds and (recall, and ).

The fundamental difference that the intermittent convex integration scheme presents over the Euler and schemes is that the building blocks are chosen to be intermittent, by which we mean that the size of their norms differs vastly for different values of . In particular, in view of Equation 4.2a and Equation 4.2b it is natural to normalize

and in order to control the term stress arising from , it will immediately become apparent that we need to ensure

for some . To see this, we use the order linear operator which inverts the divergence, and heuristically estimate the contribution of the dissipative term resulting from the principal perturbation to the Reynolds stress error :

where we have used the heuristic that only contains frequencies , whereas has frequencies as large as , and in the last inequality used Equation 4.2b and the Sobolev embedding . In order to ensure that the right side of Equation 4.6 produces an error which is compatible with the inductive stress assumption Equation 4.2b at level , and taking into account that , since , we must ensure that

The inequality Equation 4.7 justifies the need to ensure Equation 4.5. The former condition is, roughly speaking, also sufficient: first take and then , in terms of the in Equation 4.5.

The bounds Equation 4.4 and Equation 4.5 justify the need to have building blocks which are intermittent. In Fourier space, this means that their Fourier support is spread out (think, a Dirichlet kernel vs a plane wave), whereas in physical space this means that the supports have very small measure, but that the have very high amplitude on this support. We also note that in dimensions, with the normalization Equation 4.4, based on a simple scaling argument one may deduce that the smallest the norm of can be made is . In view of the requirement Equation 4.5, this shows that in the two dimensional case , the proof breaks down, as it should. Moreover, this heuristic shows that the higher the dimension is, the easier it is to achieve Equation 4.5.

With the above requirements in mind, intermittent Beltrami waves were introduced Reference 23 to serve as new ’s. These waves modify the usual Beltrami plane waves used in the Euler constructions, by adding oscillations that mimic the structure of a three dimensional Dirichlet kernel, and they are compactly supported in Fourier space. Based on an analogy with Mikado flows Reference 51, in the joint work with Colombo Reference 16 intermittent jets were introduced to serve as ’s. These flows achieve the same level of concentration in terms of Equation 4.5 as intermittent Beltrami waves, but have a number of advantageous properties since instead of being compactly supported in Fourier space, they are compactly supported in physical space, and as such may be chosen to satisfy for . This reduces the number of error terms which have to be estimated on the right side of Equation 4.3. Besides having a different scale of periodization and concentration (cf. the estimate Equation 4.9 below), one of the major differences between intermittent jets and Mikado flows is that intermittent jets are not anymore stationary solutions of the 3D Euler equations (neither were the intermittent Beltrami waves). More precisely, in order to obtain Equation 4.5 with , we need to (slowly) cut the pipe flows in the direction parallel to the pipe (in addition to the usual cutoff in the direction orthogonal to the pipe). This turns out to cause a number of additional difficulties, as explained in Equation 4.13 below.

For the case of intermittent jets, in order to parameterize the concentration of the , we introduce two parameters (a length in the direction parallel to the pipe) and (a length in the direction perpendicular to the pipe) such that

Each intermittent jet is defined to be supported on many cylinders of diameter and length . In particular, the measure of the support of is . We note that such scalings are consistent with the jet being of frequency . Finally, we normalize such that its norm is . Hence by scaling arguments, one expects an estimate of the form

for and . In particular, the smallness condition Equation 4.5 holds once we require that

for some . Note that Equation 4.10, together with the condition Equation 4.8, rules out geometric growth of the frequency , i.e., .

Next, consider the estimate Equation 4.2a. Using Hölder’s inequality to naïvely estimate the norm of the principal perturbation , we obtain

In view of Equation 4.9 with , we cannot however inductively propagate good estimates on the norm of , and as such, the above naïve estimate is not suitable in order to obtain Equation 4.2a. To obtain a better estimate, we will utilize the following heuristic observation: given a function with frequency contained in a ball of radius and a -periodic function , if , then

A precise statement for the above heuristic is given in Reference 23, Lemma 3.7. Hence using that is of frequency roughly , and that is -periodic with , and since by Equation 4.8 we have , we obtain from Equation 4.11 that

where we have used Equation 4.2b. This proves the feasibility of Equation 4.2a, which is crucial if we wish to obtain a finite energy weak solution in the limit.

We return to discuss the remaining terms in Equation 4.3. First, we note that the last term on the right side of Equation 4.3 may be bounded quite easily, by using , the Sobolev embedding , and an argument similar to Equation 4.12 but in :

where we have used Equation 4.2a, Equation 4.2b, and Equation 4.5. Since and , the above estimate is easily seen to be , as is required by Equation 4.2b at level .

The only terms from Equation 4.3 are the high frequency part of the oscillation error, and the temporal derivative term. We note crucially that in comparison to the Beltrami or the Mikado waves used for the 3D Euler constructions, the intermittent building blocks used in Reference 16Reference 23 introduce addition difficulties in handling the oscillation error, because the intermittent building blocks do not anymore solve stationary 3D Euler (to leading order). For the intermittent jets of Reference 16 we have

Similar to how the Nash error for the Euler equations was dealt with (see Equation 3.11), the high frequency error in Equation 4.13 experiences a gain when one inverts the divergence equation. Note however that this high frequency is not since the lowest active frequency in is ; nonetheless, as in Equation 4.8.

In order to take care of the main term in Equation 4.13, the intermittent jets are carefully designed to oscillate in time such that this term can be written as a temporal derivative

for some large parameter . This error can absorbed by introducing, in addition to the principal corrector, also a temporal corrector defined as

where is the Helmholtz projection and is the projection onto functions with mean zero. Thus pairing the oscillation error with the time derivative of the temporal corrector, we obtain

This identity is crucial for the intermittent convex integration scheme for 3D Navier–Stokes to close. In essence, the intermittent building blocks we have chosen in the construction do not solve to leading order the stationary 3D Euler equations, but instead to leading order they solve the time-dependent 3D Euler equations.

Finally, analogous to the Euler case, we define a divergence corrector to fix the fact that is not, as defined, divergence free. The total perturbation is then defined to be

Due to and , a number of new error terms arise on the right side of Equation 4.3, most of them being benign. The main new error term arises from the temporal oscillation in the definition of the intermittent jets, which introduce an error from the term which has a factor of in front. On the other hand, the oscillation error is inversely proportional to , and thus one has to carefully choose to optimize between these two errors. We omit these technical details and refer the reader to the summary given in Reference 22, Section 7.

4.1.2. Vanishing viscosity limits of finite energy mild solutions

A natural question to consider is whether the nonconservative weak solutions to the 3D Euler equations, which were obtained in Reference 20Reference 93, arise as vanishing viscosity limits of weak solutions to the Navier–Stokes equations. In this direction, as a direct consequence of the proof of Theorem 4.1, one obtains that the answer is yes, at least when one considers a sufficiently wide class of weak solutions to Equation 2.1:

Theorem 4.2 (Theorem 1.3 in Reference 23).

For let be a zero-mean weak solution of the 3D Euler equations. Then there exists , a sequence , and a uniformly bounded sequence of weak solutions to the Navier–Stokes equations in the sense of Definition 2.2, with strongly in .

The above result shows that being a strong limit of weak solutions to the Navier–Stokes equations, in the sense of Definition 2.2, cannot serve as a selection criterion for weak solutions of the Euler equation. Theorem 4.2 however does not rule out a selection criterion based on vanishing viscosity limits of Leray–Hopf weak solutions; we refer the reader to Reference 5Reference 7Reference 8Reference 175 where this question is discussed in the context of convex integration, symmetry breaking, and in the presence of solid boundaries.

4.2. Nonuniqueness of weak solutions with partial regularity in time

As mentioned above, intermittent jets were introduced in the joint work of the authors and Colombo Reference 16. The main goal of that paper was to give an example of a mild/weak solution to the Navier–Stokes equation whose singular set of times is both nonempty and has Hausdorff dimension strictly less than , i.e., it has partial regularity in time (a property that all Leray–Hopf weak solutions possess Reference 115). The main result in Reference 16 is:

Theorem 4.3 (Theorem 1.1 in Reference 16).

There exists such that the following holds. For , let be two strong solutions of the Navier–Stokes equations with initial data and of zero mean. There exists a weak solution of Equation 2.1, in the sense of Definition 2.2, such that

Moreover, for every such there exists a zero Lebesgue measure set of times with Hausdorff dimension less than , such that

In particular, the weak solution is almost everywhere smooth.

Note that Theorem 4.3 establishes the nonuniqueness of weak/mild solutions (cf. Defintion 2.2) to the Cauchy problem for the 3D Navier–Stokes equation, for any initial datum which permits the local-in-time solvability of 4.3 (e.g., or ). Indeed, for such one may uniquely define the solution in Theorem 4.3 for a suitable time , and then all one needs to do is choose as a shear flow whose kinetic at time is larger than that of . Due to the energy inequality, the weak solution provided by Theorem 4.3 is not the same as the global-in-time Leray–Hopf weak solution with datum , yielding nonuniqueness. In fact, this argument may be modified to hold for any incompressible , cf. Reference 16, Remark 1.4.

The proof of Theorem 4.3 builds on the one of Theorem 4.1, but additionally it inductively keeps track of good time intervals on which the approximate solutions are in fact smooth solutions of Equation 2.1 and are untouched in later inductive steps. This is achieved by employing the method of gluing introduced in Reference 20Reference 93. Taking the countable union of the good regions over each inductive step a fractal set is formed; on this set the solution is smooth, and the complement of this set has Hausdorff dimension strictly less than .

4.3. Further developments of intermittent convex integration schemes

The new flavor of convex integration introduced in Reference 23 and Reference 16, has recently been extended and improved, in order to obtain nonuniqueness and other flexibility-type results for various models arising in hydrodynamics and PDEs in general. We mention here a few of these results, as they present interesting applications of the idea that flexibility may also be attained via low integrability, not just low regularity.

The transport equation. Using a version of the classical Mikado flows but which also take into account spatial concentrations (called intermittent Mikado flows), the authors of Reference 128Reference 129Reference 130 have established the existence of nonrenormalized solutions, as well as Eulerian nonuniqueness, for the transport and continuity equations with Sobolev vector fields. These striking counterexamples point towards the sharpness of the classical results for renormalized solutions Reference 65. We also refer the reader to Reference 14 where a number of new results (both in terms of uniqueness and nonuniqueness) were obtained for positive solutions to the transport equation, and to Reference 34 where temporal intermittency and oscillation is introduced in the intermittent convex integration scheme in order to extend the integrability range.

Hyperviscous 3D Navier–Stokes. The authors of Reference 122 and Reference 16 have independently proven that intermittent convex integration is also applicable to establish the nonuniqueness of finite energy weak/mild solutions to the fractionally dissipative 3D Navier–Stokes equations with dissipation , and (the so-called Lions criticality threshold Reference 117). We note that similar results may also be established Reference 121 for the hypoviscous 2D Navier–Stokes equation .

The stochastic 3D Navier–Stokes equations. Nonuniqueness in law for the stochastic 3D Navier–Stokes system, with either an additive or a linear multiplicative noise driven by a Wiener process, was recently obtained in a remarkable paper Reference 88. See also the result Reference 179 which considers the stochastic fractionally dissipative Navier–Stokes equation in the full supercritical regime .

Other hydrodynamic models. In Reference 48, it is shown that intermittent convex integration methods can be adapted to prove the nonuniqueness of Leray–Hopf weak solutions for the 3D Hall-MHD system. We note that the nonuniqueness result in Reference 48 fundamentally relies on the presence of the Hall term which is dominant when compared to (see also the recent result Reference 49 for the electron-MHD system). Lastly, we mention that the technique of convex integration was applied to obtain nonuniqueness of distributional solutions for a model of non-Newtonian fluids called power law fluids Reference 24.

4.4. Some open problems in the context of 3D Navier–Stokes

To date it remains open to show that the regularity parameter in Theorems 4.1 or 4.3 may be taken to be “significant”, for instance . Due to Reference 83, the results cannot hold for . Note however that one may formally write the Navier–Stokes system in arbitrary dimensions , and that in higher dimensions we have stronger forms of intermittency/spatial concentration. Exploring this observation, in Reference 123 it was shown that for , the statement corresponding to Theorem 4.1 for time-independent solutions holds for . Moreover, in Reference 164, Theorem 29, it is shown that as we send the dimension , one may prove Theorem 4.1 with .

Equally challenging to increasing the regularity of the solutions in Theorems 4.1 or 4.3 seems to be to improve their integrability from to for some which “significantly” departs from . Note that due to Reference 84 we know that uniqueness holds for mild solutions in with ; establishing the sharpness of this criterion via convex integration appears to be out of reach with current methods; see Reference 22, Problem 8.

The energy inequality Equation 2.5 presents a formidable obstacle towards extending the results of Theorems 4.1 or 4.3 to the class of Leray–Hopf weak solutions. However, what happens if we give up on the energy equality and only retain the regularity/integrability provided by the energy class? Recall the discussion in Section 2.1.2. The weak/mild solutions of Definition 2.2 need not satisfy the energy inequality Equation 2.5 even if they lie in the Leray–Hopf class , as long as they do not belong to or another space with similar scaling Reference 151. This leaves open an intriguing possibility:

Problem 4.4 (Nonuniqueness of mild solutions in the energy class).

Is it possible to prove a stronger version of Theorem 4.1, by additionally requiring that ?

We again emphasize that Problem 4.4 does not require that the weak solution satisfies the energy inequality. Curiously, in terms of the parabolic scaling that is natural for the 3D Navier–Stokes system, , the space scales in the same way as the space , which is already present in Theorem 4.1. To date, however, it is not known how to parabolically trade temporal integrability for spatial regularity within the framework of an intermittent convex integration scheme for 3D Navier–Stokes.

5. Convex integration constructions for the MHD equations

As discussed in Section 2.3.3, the rigidity parts a in Conjecture 2.10 and a in Conjecture 2.11 have been established rigorously. In this section, we discuss partial progress towards the flexible parts of these Onsager-type dichotomies for ideal MHD Equation 2.19.

5.1. Bounded wild weak solutions of the MHD system

Concerning the flexible part b of Conjecture 2.10, we start with the example of Bronzi, Lopes Filho, and Nussenzveig Lopes Reference 13. The authors in Reference 13 impose a symmetry assumption which embeds the ideal MHD system into a two-and-a-half dimensional Euler flow. If is a weak solution of 3D Euler independent of , then setting the velocity field in Equation 2.19 to and the magnetic field to , we obtain a weak solution of the ideal MHD system, which is independent of . If , then such a weak solution has a nontrivial magnetic field. For as in part b of Theorem 2.4, the total energy (defined in Equation 2.21) of these weak solutions is dissipated, but both , and are hence constant functions of time (recall the definitions Equation 2.26 and Equation 2.27). Thus, the additional flexibility requirement of Conjecture 2.10—that is not conserved—seems to cause new difficulties when compared to the Euler case.

A more fundamental source of difficulties present in the analysis of the ideal MHD system is that as soon as , the magnetic helicity must be a constant function of time Reference 1Reference 76Reference 100. But typically, convex integration schemes are able to “break” all quadratic conservation laws which are well defined at the regularity level of the weak solutions constructed. Thus, it is very nontrivial to construct solutions which conserve the magnetic helicity but do not conserve the Elsässer energies.⁠Footnote14

14

One is faced with a similar difficulty in attacking Conjecture 2.7 or in trying to establish flexibility of the Casimirs for 2D SQG or 2D Euler; see Remark 2.9.

The first such result in the context of bounded weak solutions was recently obtained by Faraco, Lindberg, and Székelyhidi Reference 78 who use the convex integration scheme of Reference 55 to construct bounded weak solutions of Equation 2.19 which have compact support in space and time. The authors of Reference 78 also establish a number of rigidity results for 2D ideal MHD, but in terms of flexibility, their main conclusion is the 3D result:

Theorem 5.1 (Theorem 1.1 in Reference 78).

There exist bounded, compactly supported weak solutions of ideal MHD in , with both nontrivial, such that neither total energy nor cross helicity is conserved in time.

Since the solutions constructed in Theorem 5.1 have compact support in time, and since they conserve the magnetic helicity, we have that must vanish at all times , even though is not identically equal to zero.

Theorem 5.1 represents the first result towards the flexible part b of Conjecture 2.10. The analogous result in the Onsager program for 3D Euler would be Theorem 3.1 from Reference 55.

The proof of Theorem 5.1 uses the classical framework provided by the flavor of convex integration of De Lellis and Székelyhidi Reference 55 and the Tartar framework Reference 167, as axiomatized in Reference 163. Broadly speaking, the additional rigidity provided by the conservation of magnetic helicity is a manifestation of the weakly closed nature of the Maxwell equations for the magnetic field Reference 166. As such, when performing the plane-wave analysis, the interaction with the momentum equation for the velocity field , which has a large relaxation, has to be considered very carefully. For instance, in an earlier work Reference 76 Faraco and Lindberg show that there exist nontrival smooth strict subsolutions of 3D ideal MHD, with compact support in space-time, but that the interior of the 3D -convex hull is empty, which makes it difficult to implement a convex integration scheme starting from this subsolution. The authors of Reference 78 instead develop a variant of the convex integration scheme directly on differential two-forms, retaining consistency with the phase space geometry of the 3D ideal MHD system; see Reference 78, Sections 3, 4 for details.

5.2. Weak solutions which do not conserve magnetic helicity

In order to make progress on the flexible part b of Conjecture 2.11, one has to be able to construct weak solutions of ideal MHD Equation 2.19 which have finite total energy (as required by Defintion 2.8), but do not lie in , since otherwise they would conserve . Moreover, in view of Taylor’s conjecture (cf. Theorem 2.13), such weak solutions cannot be constructed as weak ideal limits of MHD Leray–Hopf weak solutions (the usual weak-compactness methods via smooth approximations fail anyway, for the same reasons they fail in 3D Euler). Clearly, the uniform in space convex integration developed in the proof of Theorem 5.1 is not well suited to achieve this goal. However, the intermittent convex integration scheme developed in the context of 3D Navier–Stokes (see Section 4) stands a chance, because it is exactly designed to explore the low integrability of weak solutions, via a careful choice of intermittent building blocks. This idea was explored by the authors and Beekie in Reference 10:

Theorem 5.2 (Theorem 1.4 in Reference 10).

There exists such that the following holds. There exist weak solutions of Equation 2.19, in the sense of Definition 2.8, which do not conserve magnetic helicity. In particular, there exist solutions as above with and . For these solutions the total energy and cross helicity are nontrivial, nonconstant functions of time.

Theorem 5.2 provides the first result on the flexible part b of Conjecture 2.11 by providing an example of a nonconservative weak solution to the ideal MHD equations which lies in for some , for which and are all nontrivial. Besides establishing the nonuniqueness of weak solutions to Equation 2.19 in the sense of Definition 2.8, Theorem 5.2 also gives an existence result for weak solutions to Equation 2.19 at this low integrability level. Lastly, we emphasize that in view of Theorem 2.13, the weak solutions of 3D ideal MHD which we construct in Theorem 5.2 cannot be obtained as weak ideal limits from Leray–Hopf weak solutions to Equation 2.20.

The proof of Theorem 5.2 builds on the intermittent convex integration schemes developed in Reference 16Reference 23. The main new difficulties arise from the specific geometric structure of the nonlinear terms in 3D MHD so that the intermittent building blocks used for 3D Navier–Stokes (intermittent Beltrami flows, intermittent jets, viscous eddies) are not useful for the ideal MHD system.

Informally, the building blocks used in the 3D Navier–Stokes proof are designed to handle the dissipative term , and as such require more than two-dimensional intermittency ( in Equation 4.5); moreover, the high-frequency component of the oscillation error is handled by introducing a temporal corrector. For ideal MHD, we do not have to worry about a dissipative term, and so the role of intermittency is different. Here intermittency is needed in order to treat the high frequency nonlinear oscillation errors arising from and , which cannot be fixed anymore by adding temporal correctors, essentially because the velocity intermittent building blocks and the magnetic intermittent building blocks have nontrivial overlap even if . To see this, we first note that the structure of the MHD nonlinearities requires the building blocks’ direction of oscillation, , to be orthogonal to two direction vectors and . These two orthogonal direction vectors are needed in order to simultaneously cancel both the previous velocity Reynolds stress, which is symmetric, and the previous magnetic stress, which is antisymmetric. This only permits the usage of one-dimensional intermittency in our building blocks, meaning the maximal smallness that can be gained in and is proportional to (compare to Equation 4.4 and Equation 4.5 for Navier–Stokes). This amount of gain is by itself not sufficient to close the scheme.

The main idea in the proof of Theorem 5.2 is to construct a set of intermittent building blocks adapted to the MHD geometry, which we call intermittent shear velocity flows and intermittent shear magnetic flows . Their spatial support is given by thickened planes spanned by the two orthogonal vectors and , their support has volume (which plays a role akin to the in Equation 4.8); they are periodized to scale , and their only direction of oscillation is given only by the vector which is orthogonal to both and . Using these orthogonality properties, the contributions to the oscillation errors from can be handled suitably. In order to treat , we note that the product of two rationally skew-oriented 1D intermittent building blocks is more intermittent than each one of them; it has two dimensional intermittency because the intersection of two thickened (nonparallel) planes is given by a thickened line, which has 2D smallness. That is, we may show that the Lebesgue measure of , , and is when . Suitably choosing allows one to treat the remaining oscillation errors. In summary, intermittency is used in Theorem 5.2 to treat nonlinear errors, instead of linear ones for Navier–Stokes; we refer to Reference 10 for details.

5.3. Some open problems in the context of the MHD system

The applicability of convex integration methods to the ideal MHD system is at an early development stage. For instance, although the authors of Reference 78 have recently implemented a convex integration scheme for Equation 2.19, to date it remains open to implement a Nash-type convex integration, for any . As such, the flexible part b of Conjecture 2.10 remains open.

As is the case with the intermittent convex integration schemes for 3D Navier–Stokes, it seems that fundamentally new ideas are needed to substantially increase the value of in Theorem 5.2, or to increase the integrability index of the solutions from Reference 10 from to with closer to . Because of this, part b of Conjecture 2.11 remains widely open.

Acknowledgments

The authors are grateful to Susan Friedlander for stimulating discussions and for fruitful suggestions in improving this article. We also thank Rajendra Beekie and Matthew Novack for helpful comments.

About the authors

Tristan Buckmaster is assistant professor of mathematics at Princeton University. He completed his PhD at the University of Leipzig/Max Planck Institute for Mathematics in the Sciences, Leipzig, Germany, in 2014. He then spent three years as a Courant Instructor at New York University and has subsequently joined Princeton University in 2017. His main area of interest is partial differential equations with a particular focus on equations related to hydrodynamics.

Vlad Vicol is professor of mathematics at New York University, the Courant Institute for Mathematical Sciences. He received his PhD from the University of Southern California in 2010. He taught at the University of Chicago and Princeton University before joining the faculty at the Courant Institute in 2018. His research focuses on the analysis of partial differential equations arising in fluid dynamics.

Table of Contents

  1. Abstract
  2. 1. Introduction
  3. 2. Energy equalities and their validity
    1. 2.1. The Navier–Stokes equations and the zeroth law of turbulence
    2. Definition 2.1 (Leray–Hopf weak solution).
    3. Definition 2.2 (Weak/mild solution).
    4. 2.2. The Euler equations and Onsager’s conjecture
    5. Definition 2.3 (Weak solution).
    6. Theorem 2.4 (Onsager’s conjecture in Hölder spaces).
    7. Conjecture 2.6 (Deviation from the Kolmogorov–Obhukov power spectrum).
    8. Conjecture 2.7 (Helicity flexibility).
    9. 2.3. The MHD equations and Taylor’s conjecture
    10. Definition 2.8 (Weak/distributional solution).
    11. Conjecture 2.10 (Onsager-type conjecture for the Elsässer energies).
    12. Conjecture 2.11 (Onsager-type conjecture for the magnetic helicity).
    13. Definition 2.12 (Weak ideal limit).
    14. Theorem 2.13 (Taylor’s conjecture; Theorem 1.2 in 77).
  4. 3. Convex integration and Nash schemes for the Euler equations
    1. 3.1. The flexible part of Onsager’s conjecture: First paradoxical examples
    2. 3.2. The results
    3. Theorem 3.1 (Theorem 4.1 in 55).
    4. Theorem 3.2 (Theorem 1 in 56).
    5. 3.3. The result: A Nash-type convex integration scheme
    6. Theorem 3.3 (Theorem 1.1 in 58).
    7. 3.4. Climbing the Onsager ladder
    8. 3.5. Resolution of the flexible side of Onsager’s conjecture
    9. Theorem 3.4 (Theorem 1 in 93).
    10. Theorem 3.5 (Theorem 1.1 in 20).
    11. 3.6. Some open problems in the context of 3D Euler
  5. 4. Intermittent convex integration for the Navier–Stokes equations
    1. 4.1. Nonuniqueness of weak solutions with finite kinetic energy
    2. Theorem 4.1 (Theorem 1.2 in 23).
    3. Theorem 4.2 (Theorem 1.3 in 23).
    4. 4.2. Nonuniqueness of weak solutions with partial regularity in time
    5. Theorem 4.3 (Theorem 1.1 in 16).
    6. 4.3. Further developments of intermittent convex integration schemes
    7. 4.4. Some open problems in the context of 3D Navier–Stokes
    8. Problem 4.4 (Nonuniqueness of mild solutions in the energy class).
  6. 5. Convex integration constructions for the MHD equations
    1. 5.1. Bounded wild weak solutions of the MHD system
    2. Theorem 5.1 (Theorem 1.1 in 78).
    3. 5.2. Weak solutions which do not conserve magnetic helicity
    4. Theorem 5.2 (Theorem 1.4 in 10).
    5. 5.3. Some open problems in the context of the MHD system
  7. Acknowledgments
  8. About the authors

Mathematical Fragments

Equation (2.1)
Equation (2.2)
Equation (2.3)
Equation (2.4)
Equation (2.5)
Definition 2.1 (Leray–Hopf weak solution).

A vector field

is called a Leray–Hopf weak solution of the Navier–Stokes equations if for any the vector field is weakly divergence free, has zero mean, satisfies the Navier–Stokes equations distributionally:

for any divergence free test function , and satisfies the energy inequality Equation 2.5 for .⁠Footnote4

4

A stronger form of the energy inequality holds by Leray’s construction: Equation 2.5 holds for a.e. and all . Moreover, an a priori stronger form of weak solution may be constructed by modifying Leray’s proof. These so-called suitable weak solutions obey a version of the energy inequality which is localized in space and time, which allows one to prove that they are smooth except for a putative singular space-time set with one-dimensional (1D) parabolic Hausdorff measure which is zero Reference 25Reference 145Reference 146; see also the more recent reviews Reference 114Reference 140Reference 143.

Definition 2.2 (Weak/mild solution).

A vector field is called a weak or distributional solution of the Navier–Stokes equations if for any the vector field is weakly divergence free, has zero mean, and Equation 2.6 holds for for any smooth divergence free test function .

Equation (2.7)
Equation (2.8)
Equation (2.10)
Equation (2.11)
Equation (2.12)
Equation (2.13)
Equation (2.15)
Equation (2.16)
Equation (2.17)
Definition 2.3 (Weak solution).

A vector field is called a weak solution of the Euler equations on an open if for any the vector field is weakly divergence free, has zero mean, and satisfies Equation 2.16 in the sense of distributions; that is,

holds for any divergence free test function . The pressure can be recovered by the formula with of zero mean.

Theorem 2.4 (Onsager’s conjecture in Hölder spaces).
(a)

Any weak solution belonging to the Hölder space for conserves its kinetic energy.

(b)

For any there exist weak solutions which dissipate kinetic energy, i.e., the kinetic energy is a nonincreasing function of time.

Conjecture 2.6 (Deviation from the Kolmogorov–Obhukov power spectrum).

There exists an and infinitely many weak solutions of the 3D Euler equations Equation 2.16, which dissipate kinetic energy.

Equation (2.18)
Conjecture 2.7 (Helicity flexibility).

There exists a weak solution of the 3D Euler equations Equation 2.16, with for some , such that the helicity is not a constant function of time.

Equation (2.19)
Equation (2.20)
Equation (2.21)
Equation (2.22)
Equation (2.23)
Equation (2.24)
Equation (2.26)
Definition 2.8 (Weak/distributional solution).

We say is a weak solution of the ideal MHD system Equation 2.19 on an open interval , if for any , the vector fields and are divergence free in the sense of distributions, they have zero mean, and Equation 2.19 holds in the sense of distributions; i.e.,

hold for all divergence free test functions .

Equation (2.27)
Equation (2.29)
Remark 2.9 (2D Euler and Surface Quasi Geostrophic).

This situation encountered here in which the hydrodynamic model has conservation laws at different levels of regularity ( vs ) is not uncommon. Another occurrence is in the context of the 2D Surface Quasi Geostrophic (SQG) and 2D Euler equations. For both of these equations smooth solutions conserve the norm (in fact all the norms with ) of the temperature (vorticity for the potential velocity Reference 21) for SQG (resp., the scalar vorticity for 2D Euler). With respect to these so-called Casimirs, the respective Hamiltonians of these systems lie at a negative level of regularity: for SQG (resp., for 2D Euler). As such, much of the discussion presented in this paper concerning the MHD system has an analogue in the context of the SQG Reference 21Reference 94Reference 96 and 2D Euler equations Reference 29Reference 30.

Conjecture 2.10 (Onsager-type conjecture for the Elsässer energies).
(a)

Any weak solution of the ideal MHD system Equation 2.19 belonging (resp., for , conserves the total energy and the cross helicity .

(b)

For any there exist weak solutions (resp., , which dissipate the total energy , and for which the cross helicity is not a constant function of time.

Conjecture 2.11 (Onsager-type conjecture for the magnetic helicity).
(a)

Any weak solution of the ideal MHD system Equation 2.19 belonging to (resp., conserves the magnetic helicity .

(b)

For any , there exist weak solutions (resp., for which the magnetic helicity is not a constant function of time.

Equation (2.30)
Definition 2.12 (Weak ideal limit).

Let be a sequence of vanishing viscosities and resistivities. Associated to a sequence of divergence free initial data converging weakly in , let be a sequence of Leray–Hopf weak solutions of Equation 2.20. Any pair of functions such that in , are called a weak ideal limit of the sequence .⁠Footnote9

9

Note however that a weak ideal limit need not be a weak solution of the ideal MHD equations in the sense of Definition 2.8, for the same reason that a vanishing viscosity limit of Leray–Hopf weak solutions of the 3D Navier–Stokes equations need not be a weak/distributional solution of the 3D Euler equations; only measure valued solutions are known to be achieved as subsequential limits Reference 66.

Theorem 2.13 (Taylor’s conjecture; Theorem 1.2 in Reference 77).

Suppose is a weak ideal limit of a sequence of Leray–Hopf weak solutions. Then is a constant function of time. In particular, finite energy weak solutions of the ideal MHD equations Equation 2.19, which are weak ideal limits, conserve magnetic helicity.

Equation (2.31)
Theorem 3.1 (Theorem 4.1 in Reference 55).

For any open bounded space-time domain , there exists a weak solution of the Euler equations Equation 2.16 , in the sense of Definition 2.3, such that for a.e. , and for a.e. . Moreover, there exists a sequence of functions such that:

, and ,

in as ,

is uniformly bounded in ,

in for any .

Theorem 3.2 (Theorem 1 in Reference 56).

For any dimension , there exists bounded, compactly supported divergence-free initial data , for which there exist infinitely many weak solutions of the Euler equations, such that the admissibility conditions i, ii, and iii hold.

Theorem 3.3 (Theorem 1.1 in Reference 58).

Assume is a smooth function. Then there is a continuous vector field and a continuous scalar field which solve the incompressible Euler equations Equation 2.16 in the sense of distributions, and such that

for all .

Equation (3.2)
Equation (3.3)
Equation (3.4)
Equation (3.5)
Equation (3.6)
Equation (3.7)
Equation (3.8)
Equation (3.9)
Equation (3.10)
Equation (3.11)
Theorem 3.4 (Theorem 1 in Reference 93).

For any there exists a nonzero weak solution , such that vanishes identically outside of a finite interval.

Theorem 3.5 (Theorem 1.1 in Reference 20).

Let be a strictly positive smooth function. For any there exists a weak solution of the Euler equations Equation 2.16, whose kinetic energy at time equals .

Theorem 4.1 (Theorem 1.2 in Reference 23).

There exists , such that for any nonnegative smooth function and any , there exists a weak solution of the Navier–Stokes equations

such that holds for all .

Equation (4.1)
Equation (4.2)
Equation (4.3)
Equation (4.4)
Equation (4.5)
Equation (4.6)
Equation (4.7)
Equation (4.8)
Equation (4.9)
Equation (4.10)
Equation (4.11)
Equation (4.12)
Equation (4.13)
Theorem 4.2 (Theorem 1.3 in Reference 23).

For let be a zero-mean weak solution of the 3D Euler equations. Then there exists , a sequence , and a uniformly bounded sequence of weak solutions to the Navier–Stokes equations in the sense of Definition 2.2, with strongly in .

Theorem 4.3 (Theorem 1.1 in Reference 16).

There exists such that the following holds. For , let be two strong solutions of the Navier–Stokes equations with initial data and of zero mean. There exists a weak solution of Equation 2.1, in the sense of Definition 2.2, such that

Moreover, for every such there exists a zero Lebesgue measure set of times with Hausdorff dimension less than , such that

In particular, the weak solution is almost everywhere smooth.

Problem 4.4 (Nonuniqueness of mild solutions in the energy class).

Is it possible to prove a stronger version of Theorem 4.1, by additionally requiring that ?

Theorem 5.1 (Theorem 1.1 in Reference 78).

There exist bounded, compactly supported weak solutions of ideal MHD in , with both nontrivial, such that neither total energy nor cross helicity is conserved in time.

Theorem 5.2 (Theorem 1.4 in Reference 10).

There exists such that the following holds. There exist weak solutions of Equation 2.19, in the sense of Definition 2.8, which do not conserve magnetic helicity. In particular, there exist solutions as above with and . For these solutions the total energy and cross helicity are nontrivial, nonconstant functions of time.

References

Reference [1]
H. Aluie, Hydrodynamic and magnetohydrodynamic turbulence: Invariants, cascades, and locality, Ph.D. thesis, Johns Hopkins University, 2009.
Reference [2]
H. Aluie and G. L. Eyink, Scale locality of magnetohydrodynamic turbulence, Physical Review Letters 104 (2010), no. 8, 081101.
Reference [3]
F. Anselmet, Y. Gagne, E. J. Hopfinger, and R. A. Antonia, High-order velocity structure functions in turbulent shear flows, J. Fluid Mech. 140 (1984), 63–89.
Reference [4]
V. I. Arnold and B. A. Khesin, Topological methods in hydrodynamics, Applied Mathematical Sciences, vol. 125, Springer-Verlag, New York, 1998. MR1612569,
Show rawAMSref \bib{ArnoldKhesin99}{book}{ author={Arnold, Vladimir I.}, author={Khesin, Boris A.}, title={Topological methods in hydrodynamics}, series={Applied Mathematical Sciences}, volume={125}, publisher={Springer-Verlag, New York}, date={1998}, pages={xvi+374}, isbn={0-387-94947-X}, review={\MR {1612569}}, }
Reference [5]
C. Bardos, M. C. Lopes Filho, D. Niu, H. J. Nussenzveig Lopes, and E. S. Titi, Stability of two-dimensional viscous incompressible flows under three-dimensional perturbations and inviscid symmetry breaking, SIAM J. Math. Anal. 45 (2013), no. 3, 1871–1885, DOI 10.1137/120862569. MR3068562,
Show rawAMSref \bib{BardosEtAl13}{article}{ author={Bardos, C.}, author={Lopes Filho, M. C.}, author={Niu, Dongjuan}, author={Nussenzveig Lopes, H. J.}, author={Titi, E. S.}, title={Stability of two-dimensional viscous incompressible flows under three-dimensional perturbations and inviscid symmetry breaking}, journal={SIAM J. Math. Anal.}, volume={45}, date={2013}, number={3}, pages={1871--1885}, issn={0036-1410}, review={\MR {3068562}}, doi={10.1137/120862569}, }
Reference [6]
C. Bardos, L. Sekelikhidi Jr., and È. Videmann, On the absence of uniqueness for the Euler equations: the effect of the boundary (Russian, with Russian summary), Uspekhi Mat. Nauk 69 (2014), no. 2(416), 3–22, DOI 10.1070/rm2014v069n02abeh004886; English transl., Russian Math. Surveys 69 (2014), no. 2, 189–207. MR3236935,
Show rawAMSref \bib{BSV14}{article}{ author={Bardos, C.}, author={Sekelikhidi, L., Jr.}, author={Videmann, \`E.}, title={On the absence of uniqueness for the Euler equations: the effect of the boundary}, language={Russian, with Russian summary}, journal={Uspekhi Mat. Nauk}, volume={69}, date={2014}, number={2(416)}, pages={3--22}, issn={0042-1316}, translation={ journal={Russian Math. Surveys}, volume={69}, date={2014}, number={2}, pages={189--207}, issn={0036-0279}, }, review={\MR {3236935}}, doi={10.1070/rm2014v069n02abeh004886}, }
Reference [7]
C. Bardos, L. Sekelikhidi Jr., and È. Videmann, Non-uniqueness for the Euler equations: the effect of the boundary (Russian, with Russian summary), Uspekhi Mat. Nauk 69 (2014), no. 2(416), 3–22, DOI 10.1070/rm2014v069n02abeh004886; English transl., Russian Math. Surveys 69 (2014), no. 2, 189–207. MR3236935,
Show rawAMSref \bib{BardosEtAl14}{article}{ author={Bardos, C.}, author={Sekelikhidi, L., Jr.}, author={Videmann, \`E.}, title={Non-uniqueness for the Euler equations: the effect of the boundary}, language={Russian, with Russian summary}, journal={Uspekhi Mat. Nauk}, volume={69}, date={2014}, number={2(416)}, pages={3--22}, issn={0042-1316}, translation={ journal={Russian Math. Surveys}, volume={69}, date={2014}, number={2}, pages={189--207}, issn={0036-0279}, }, review={\MR {3236935}}, doi={10.1070/rm2014v069n02abeh004886}, }
Reference [8]
C. Bardos, E. S. Titi, and E. Wiedemann, The vanishing viscosity as a selection principle for the Euler equations: the case of 3D shear flow (English, with English and French summaries), C. R. Math. Acad. Sci. Paris 350 (2012), no. 15-16, 757–760, DOI 10.1016/j.crma.2012.09.005. MR2981348,
Show rawAMSref \bib{BardoesTitiWiedemann12}{article}{ author={Bardos, Claude}, author={Titi, Edriss S.}, author={Wiedemann, Emil}, title={The vanishing viscosity as a selection principle for the Euler equations: the case of 3D shear flow}, language={English, with English and French summaries}, journal={C. R. Math. Acad. Sci. Paris}, volume={350}, date={2012}, number={15-16}, pages={757--760}, issn={1631-073X}, review={\MR {2981348}}, doi={10.1016/j.crma.2012.09.005}, }
Reference [9]
G. K. Batchelor, The theory of homogeneous turbulence, Cambridge Monographs on Mechanics and Applied Mathematics, Cambridge at the University Press, 1953. MR0052268,
Show rawAMSref \bib{Batchelor53}{book}{ author={Batchelor, G. K.}, title={The theory of homogeneous turbulence}, series={Cambridge Monographs on Mechanics and Applied Mathematics}, publisher={Cambridge at the University Press}, date={1953}, pages={x+197}, review={\MR {0052268}}, }
Reference [10]
R. Beekie, T. Buckmaster, and V. Vicol, Weak solutions of ideal MHD which do not conserve magnetic helicity, Ann. PDE 6 (2020), no. 1, Paper No. 1, 40, DOI 10.1007/s40818-020-0076-1. MR4105741,
Show rawAMSref \bib{BBV19}{article}{ author={Beekie, Rajendra}, author={Buckmaster, Tristan}, author={Vicol, Vlad}, title={Weak solutions of ideal MHD which do not conserve magnetic helicity}, journal={Ann. PDE}, volume={6}, date={2020}, number={1}, pages={Paper No. 1, 40}, issn={2524-5317}, review={\MR {4105741}}, doi={10.1007/s40818-020-0076-1}, }
Reference [11]
M. A. Berger, Introduction to magnetic helicity, Plasma Physics and Controlled Fusion 41 (1999), no. 12B, B167.
Reference [12]
D. Biskamp, Nonlinear magnetohydrodynamics, Cambridge Monographs on Plasma Physics, vol. 1, Cambridge University Press, Cambridge, 1993, DOI 10.1017/CBO9780511599965. MR1250152,
Show rawAMSref \bib{Biskamp97}{book}{ author={Biskamp, Dieter}, title={Nonlinear magnetohydrodynamics}, series={Cambridge Monographs on Plasma Physics}, volume={1}, publisher={Cambridge University Press, Cambridge}, date={1993}, pages={xiv+378}, isbn={0-521-40206-9}, review={\MR {1250152}}, doi={10.1017/CBO9780511599965}, }
Reference [13]
A. C. Bronzi, M. C. Lopes Filho, and H. J. Nussenzveig Lopes, Wild solutions for 2D incompressible ideal flow with passive tracer, Commun. Math. Sci. 13 (2015), no. 5, 1333–1343, DOI 10.4310/CMS.2015.v13.n5.a12. MR3344429,
Show rawAMSref \bib{BronziEtAl15}{article}{ author={Bronzi, Anne C.}, author={Lopes Filho, Milton C.}, author={Nussenzveig Lopes, Helena J.}, title={Wild solutions for 2D incompressible ideal flow with passive tracer}, journal={Commun. Math. Sci.}, volume={13}, date={2015}, number={5}, pages={1333--1343}, issn={1539-6746}, review={\MR {3344429}}, doi={10.4310/CMS.2015.v13.n5.a12}, }
Reference [14]
E. Brué, M. Colombo, and C. De Lellis, Positive solutions of transport equations and classical nonuniqueness of characteristic curves, arXiv preprint arXiv:2003.00539 (2020).
Reference [15]
T. Buckmaster, Onsager’s conjecture almost everywhere in time, Comm. Math. Phys. 333 (2015), no. 3, 1175–1198, DOI 10.1007/s00220-014-2262-z. MR3302631,
Show rawAMSref \bib{Buckmaster15}{article}{ author={Buckmaster, Tristan}, title={Onsager's conjecture almost everywhere in time}, journal={Comm. Math. Phys.}, volume={333}, date={2015}, number={3}, pages={1175--1198}, issn={0010-3616}, review={\MR {3302631}}, doi={10.1007/s00220-014-2262-z}, }
Reference [16]
T. Buckmaster, M. Colombo, and V. Vicol, Wild solutions of the Navier–Stokes equations whose singular sets in time have hausdorff dimension strictly less than , Journal of the EMS. In press. (2020).
Reference [17]
T. Buckmaster, C. De Lellis, P. Isett, and L. Székelyhidi Jr., Anomalous dissipation for -Hölder Euler flows, Ann. of Math. (2) 182 (2015), no. 1, 127–172, DOI 10.4007/annals.2015.182.1.3. MR3374958,
Show rawAMSref \bib{BDLISZ15}{article}{ author={Buckmaster, Tristan}, author={De Lellis, Camillo}, author={Isett, Philip}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={Anomalous dissipation for $1/5$-H\"{o}lder Euler flows}, journal={Ann. of Math. (2)}, volume={182}, date={2015}, number={1}, pages={127--172}, issn={0003-486X}, review={\MR {3374958}}, doi={10.4007/annals.2015.182.1.3}, }
Reference [18]
T. Buckmaster, C. De Lellis, and L. Székelyhidi Jr., Dissipative Euler flows with Onsager-critical spatial regularity, Comm. Pure Appl. Math. 69 (2016), no. 9, 1613–1670, DOI 10.1002/cpa.21586. MR3530360,
Show rawAMSref \bib{BDLSZ16}{article}{ author={Buckmaster, Tristan}, author={De Lellis, Camillo}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={Dissipative Euler flows with Onsager-critical spatial regularity}, journal={Comm. Pure Appl. Math.}, volume={69}, date={2016}, number={9}, pages={1613--1670}, issn={0010-3640}, review={\MR {3530360}}, doi={10.1002/cpa.21586}, }
Reference [19]
T. Buckmaster, C. De Lellis, and L. Székelyhidi Jr., Transporting microstructure and dissipative Euler flows, arXiv preprint arXiv:1302.2815 (2013).
Reference [20]
T. Buckmaster, C. de Lellis, L. Székelyhidi Jr., and V. Vicol, Onsager’s conjecture for admissible weak solutions, Comm. Pure Appl. Math. 72 (2019), no. 2, 229–274, DOI 10.1002/cpa.21781. MR3896021,
Show rawAMSref \bib{BDLSV19}{article}{ author={Buckmaster, Tristan}, author={de Lellis, Camillo}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, author={Vicol, Vlad}, title={Onsager's conjecture for admissible weak solutions}, journal={Comm. Pure Appl. Math.}, volume={72}, date={2019}, number={2}, pages={229--274}, issn={0010-3640}, review={\MR {3896021}}, doi={10.1002/cpa.21781}, }
Reference [21]
T. Buckmaster, S. Shkoller, and V. Vicol, Nonuniqueness of weak solutions to the SQG equation, Comm. Pure Appl. Math. 72 (2019), no. 9, 1809–1874, DOI 10.1002/cpa.21851. MR3987721,
Show rawAMSref \bib{BSV16}{article}{ author={Buckmaster, Tristan}, author={Shkoller, Steve}, author={Vicol, Vlad}, title={Nonuniqueness of weak solutions to the SQG equation}, journal={Comm. Pure Appl. Math.}, volume={72}, date={2019}, number={9}, pages={1809--1874}, issn={0010-3640}, review={\MR {3987721}}, doi={10.1002/cpa.21851}, }
Reference [22]
T. Buckmaster and V. Vicol, Convex integration and phenomenologies in turbulence, EMS Surv. Math. Sci. 6 (2019), no. 1, 173–263, DOI 10.4171/emss/34. MR4073888,
Show rawAMSref \bib{BuckmasterVicol19}{article}{ author={Buckmaster, Tristan}, author={Vicol, Vlad}, title={Convex integration and phenomenologies in turbulence}, journal={EMS Surv. Math. Sci.}, volume={6}, date={2019}, number={1}, pages={173--263}, issn={2308-2151}, review={\MR {4073888}}, doi={10.4171/emss/34}, }
Reference [23]
T. Buckmaster and V. Vicol, Nonuniqueness of weak solutions to the Navier–Stokes equation, Ann. of Math. 189 (2019), no. 1, 101–144.
Reference [24]
J. Burczak, S. Modena, and L. Székelyhidi Jr., Non uniqueness of power-law flows, arXiv preprint arXiv:2007.08011 (2020).
Reference [25]
L. Caffarelli, R. Kohn, and L. Nirenberg, Partial regularity of suitable weak solutions of the Navier–Stokes equations, Comm. Pure Appl. Math. 35 (1982), no. 6, 771–831, DOI 10.1002/cpa.3160350604. MR673830,
Show rawAMSref \bib{CaffarelliKohnNirenberg82}{article}{ author={Caffarelli, L.}, author={Kohn, R.}, author={Nirenberg, L.}, title={Partial regularity of suitable weak solutions of the Navier--Stokes equations}, journal={Comm. Pure Appl. Math.}, volume={35}, date={1982}, number={6}, pages={771--831}, issn={0010-3640}, review={\MR {673830}}, doi={10.1002/cpa.3160350604}, }
Reference [26]
R. E. Caflisch, I. Klapper, and G. Steele, Remarks on singularities, dimension and energy dissipation for ideal hydrodynamics and MHD, Comm. Math. Phys. 184 (1997), no. 2, 443–455, DOI 10.1007/s002200050067. MR1462753,
Show rawAMSref \bib{Caflisch1997}{article}{ author={Caflisch, Russel E.}, author={Klapper, Isaac}, author={Steele, Gregory}, title={Remarks on singularities, dimension and energy dissipation for ideal hydrodynamics and MHD}, journal={Comm. Math. Phys.}, volume={184}, date={1997}, number={2}, pages={443--455}, issn={0010-3616}, review={\MR {1462753}}, doi={10.1007/s002200050067}, }
Reference [27]
A. Castro, C. Córdoba, and D. Faraco, Mixing solutions for the Muskat problem, arXiv:1605.04822 (2016).
Reference [28]
J.-Y. Chemin, Perfect incompressible fluids, Oxford Lecture Series in Mathematics and its Applications, vol. 14, The Clarendon Press, Oxford University Press, New York, 1998. Translated from the 1995 French original by Isabelle Gallagher and Dragos Iftimie. MR1688875,
Show rawAMSref \bib{Chemin98}{book}{ author={Chemin, Jean-Yves}, title={Perfect incompressible fluids}, series={Oxford Lecture Series in Mathematics and its Applications}, volume={14}, note={Translated from the 1995 French original by Isabelle Gallagher and Dragos Iftimie}, publisher={The Clarendon Press, Oxford University Press, New York}, date={1998}, pages={x+187}, isbn={0-19-850397-0}, review={\MR {1688875}}, }
Reference [29]
A. Cheskidov, P. Constantin, S. Friedlander, and R. Shvydkoy, Energy conservation and Onsager’s conjecture for the Euler equations, Nonlinearity 21 (2008), no. 6, 1233–1252, DOI 10.1088/0951-7715/21/6/005. MR2422377,
Show rawAMSref \bib{CCFS08}{article}{ author={Cheskidov, A.}, author={Constantin, P.}, author={Friedlander, S.}, author={Shvydkoy, R.}, title={Energy conservation and Onsager's conjecture for the Euler equations}, journal={Nonlinearity}, volume={21}, date={2008}, number={6}, pages={1233--1252}, issn={0951-7715}, review={\MR {2422377}}, doi={10.1088/0951-7715/21/6/005}, }
Reference [30]
A. Cheskidov, M. C. L. Filho, H. J. N. Lopes, and R. Shvydkoy, Energy conservation in two-dimensional incompressible ideal fluids, Comm. Math. Phys. 348 (2016), no. 1, 129–143, DOI 10.1007/s00220-016-2730-8. MR3551263,
Show rawAMSref \bib{CLfNlS16}{article}{ author={Cheskidov, A.}, author={Filho, M. C. Lopes}, author={Lopes, H. J. Nussenzveig}, author={Shvydkoy, R.}, title={Energy conservation in two-dimensional incompressible ideal fluids}, journal={Comm. Math. Phys.}, volume={348}, date={2016}, number={1}, pages={129--143}, issn={0010-3616}, review={\MR {3551263}}, doi={10.1007/s00220-016-2730-8}, }
Reference [31]
A. Cheskidov, S. Friedlander, and N. Pavlović, Inviscid dyadic model of turbulence: the fixed point and Onsager’s conjecture, J. Math. Phys. 48 (2007), no. 6, 065503, 16, DOI 10.1063/1.2395917. MR2337019,
Show rawAMSref \bib{CheskidovFriedlanderPavlovic07}{article}{ author={Cheskidov, Alexey}, author={Friedlander, Susan}, author={Pavlovi\'{c}, Nata\v {s}a}, title={Inviscid dyadic model of turbulence: the fixed point and Onsager's conjecture}, journal={J. Math. Phys.}, volume={48}, date={2007}, number={6}, pages={065503, 16}, issn={0022-2488}, review={\MR {2337019}}, doi={10.1063/1.2395917}, }
Reference [32]
A. Cheskidov, S. Friedlander, and R. Shvydkoy, On the energy equality for weak solutions of the 3D Navier–Stokes equations, Advances in mathematical fluid mechanics, Springer, Berlin, 2010, pp. 171–175, DOI 10.1007/978-3-642-04068-9_10. MR2665030,
Show rawAMSref \bib{CFR10}{article}{ author={Cheskidov, Alexey}, author={Friedlander, Susan}, author={Shvydkoy, Roman}, title={On the energy equality for weak solutions of the 3D Navier--Stokes equations}, conference={ title={Advances in mathematical fluid mechanics}, }, book={ publisher={Springer, Berlin}, }, date={2010}, pages={171--175}, review={\MR {2665030}}, doi={10.1007/978-3-642-04068-9_10}, }
Reference [33]
A. Cheskidov and X. Luo, Energy equality for the Navier–Stokes equations in weak-in-time Onsager spaces, Nonlinearity 33 (2020), no. 4, 1388–1403, DOI 10.1088/1361-6544/ab60d3. MR4072378,
Show rawAMSref \bib{CheskidovLuo20}{article}{ author={Cheskidov, Alexey}, author={Luo, Xiaoyutao}, title={Energy equality for the Navier--Stokes equations in weak-in-time Onsager spaces}, journal={Nonlinearity}, volume={33}, date={2020}, number={4}, pages={1388--1403}, issn={0951-7715}, review={\MR {4072378}}, doi={10.1088/1361-6544/ab60d3}, }
Reference [34]
A. Cheskidov and X. Luo, Nonuniqueness of weak solutions for the transport equation at critical space regularity, arXiv:2004.09538 (2020).
Reference [35]
A. Cheskidov and R. Shvydkoy, Euler equations and turbulence: analytical approach to intermittency, SIAM J. Math. Anal. 46 (2014), no. 1, 353–374, DOI 10.1137/120876447. MR3152734,
Show rawAMSref \bib{CS2014}{article}{ author={Cheskidov, A.}, author={Shvydkoy, R.}, title={Euler equations and turbulence: analytical approach to intermittency}, journal={SIAM J. Math. Anal.}, volume={46}, date={2014}, number={1}, pages={353--374}, issn={0036-1410}, review={\MR {3152734}}, doi={10.1137/120876447}, }
Reference [36]
E. Chiodaroli, A counterexample to well-posedness of entropy solutions to the compressible Euler system, J. Hyperbolic Differ. Equ. 11 (2014), no. 3, 493–519, DOI 10.1142/S0219891614500143. MR3261301,
Show rawAMSref \bib{Chiodaroli14}{article}{ author={Chiodaroli, Elisabetta}, title={A counterexample to well-posedness of entropy solutions to the compressible Euler system}, journal={J. Hyperbolic Differ. Equ.}, volume={11}, date={2014}, number={3}, pages={493--519}, issn={0219-8916}, review={\MR {3261301}}, doi={10.1142/S0219891614500143}, }
Reference [37]
E. Chiodaroli, C. De Lellis, and O. Kreml, Global ill-posedness of the isentropic system of gas dynamics, Comm. Pure Appl. Math. 68 (2015), no. 7, 1157–1190, DOI 10.1002/cpa.21537. MR3352460,
Show rawAMSref \bib{ChiodaroliDeLellisKremlOndvrej15}{article}{ author={Chiodaroli, Elisabetta}, author={De Lellis, Camillo}, author={Kreml, Ond\v {r}ej}, title={Global ill-posedness of the isentropic system of gas dynamics}, journal={Comm. Pure Appl. Math.}, volume={68}, date={2015}, number={7}, pages={1157--1190}, issn={0010-3640}, review={\MR {3352460}}, doi={10.1002/cpa.21537}, }
Reference [38]
E. Chiodaroli, E. Feireisl, and O. Kreml, On the weak solutions to the equations of a compressible heat conducting gas, Ann. Inst. H. Poincaré Anal. Non Linéaire 32 (2015), no. 1, 225–243, DOI 10.1016/j.anihpc.2013.11.005. MR3303948,
Show rawAMSref \bib{ChFeKr15}{article}{ author={Chiodaroli, Elisabetta}, author={Feireisl, Eduard}, author={Kreml, Ond\v {r}ej}, title={On the weak solutions to the equations of a compressible heat conducting gas}, journal={Ann. Inst. H. Poincar\'{e} Anal. Non Lin\'{e}aire}, volume={32}, date={2015}, number={1}, pages={225--243}, issn={0294-1449}, review={\MR {3303948}}, doi={10.1016/j.anihpc.2013.11.005}, }
Reference [39]
A. Choffrut, -principles for the incompressible Euler equations, Arch. Ration. Mech. Anal. 210 (2013), no. 1, 133–163, DOI 10.1007/s00205-013-0639-3. MR3073150,
Show rawAMSref \bib{Choffrut13}{article}{ author={Choffrut, Antoine}, title={$h$-principles for the incompressible Euler equations}, journal={Arch. Ration. Mech. Anal.}, volume={210}, date={2013}, number={1}, pages={133--163}, issn={0003-9527}, review={\MR {3073150}}, doi={10.1007/s00205-013-0639-3}, }
Reference [40]
M. Colombo, C. De Lellis, and L. De Rosa, Ill-posedness of Leray solutions for the hypodissipative Navier–Stokes equations, Comm. Math. Phys. 362 (2018), no. 2, 659–688, DOI 10.1007/s00220-018-3177-x. MR3843425,
Show rawAMSref \bib{CDLDR18}{article}{ author={Colombo, Maria}, author={De Lellis, Camillo}, author={De Rosa, Luigi}, title={Ill-posedness of Leray solutions for the hypodissipative Navier--Stokes equations}, journal={Comm. Math. Phys.}, volume={362}, date={2018}, number={2}, pages={659--688}, issn={0010-3616}, review={\MR {3843425}}, doi={10.1007/s00220-018-3177-x}, }
Reference [41]
P. Constantin, W. E, and E. S. Titi, Onsager’s conjecture on the energy conservation for solutions of Euler’s equation, Comm. Math. Phys. 165 (1994), no. 1, 207–209. MR1298949,
Show rawAMSref \bib{ConstantinETiti94}{article}{ author={Constantin, Peter}, author={E, Weinan}, author={Titi, Edriss S.}, title={Onsager's conjecture on the energy conservation for solutions of Euler's equation}, journal={Comm. Math. Phys.}, volume={165}, date={1994}, number={1}, pages={207--209}, issn={0010-3616}, review={\MR {1298949}}, }
Reference [42]
P. Constantin and C. Foias, Navier–Stokes equations, Chicago Lectures in Mathematics, University of Chicago Press, Chicago, IL, 1988. MR972259,
Show rawAMSref \bib{ConstantinFoias88}{book}{ author={Constantin, Peter}, author={Foias, Ciprian}, title={Navier--Stokes equations}, series={Chicago Lectures in Mathematics}, publisher={University of Chicago Press, Chicago, IL}, date={1988}, pages={x+190}, isbn={0-226-11548-8}, isbn={0-226-11549-6}, review={\MR {972259}}, }
Reference [43]
P. Constantin, M. Ignatova, and H. Q. Nguyen, Inviscid limit for SQG in bounded domains, SIAM J. Math. Anal. 50 (2018), no. 6, 6196–6207, DOI 10.1137/18M1192743. MR3882956,
Show rawAMSref \bib{ConstantinIgnatovaNguyen18}{article}{ author={Constantin, Peter}, author={Ignatova, Mihaela}, author={Nguyen, Huy Q.}, title={Inviscid limit for SQG in bounded domains}, journal={SIAM J. Math. Anal.}, volume={50}, date={2018}, number={6}, pages={6196--6207}, issn={0036-1410}, review={\MR {3882956}}, doi={10.1137/18M1192743}, }
Reference [44]
P. Constantin and F. Ramos, Inviscid limit for damped and driven incompressible Navier–Stokes equations in , Comm. Math. Phys. 275 (2007), no. 2, 529–551, DOI 10.1007/s00220-007-0310-7. MR2335784,
Show rawAMSref \bib{ConstantinRamos07}{article}{ author={Constantin, P.}, author={Ramos, F.}, title={Inviscid limit for damped and driven incompressible Navier--Stokes equations in $\mathbb {R}^2$}, journal={Comm. Math. Phys.}, volume={275}, date={2007}, number={2}, pages={529--551}, issn={0010-3616}, review={\MR {2335784}}, doi={10.1007/s00220-007-0310-7}, }
Reference [45]
P. Constantin, W. E, and E. S. Titi, Onsager’s conjecture on the energy conservation for solutions of Euler’s equation, Comm. Math. Phys. 165 (1994), no. 1, 207–209. MR1298949,
Show rawAMSref \bib{CET94}{article}{ author={Constantin, Peter}, author={E, Weinan}, author={Titi, Edriss S.}, title={Onsager's conjecture on the energy conservation for solutions of Euler's equation}, journal={Comm. Math. Phys.}, volume={165}, date={1994}, number={1}, pages={207--209}, issn={0010-3616}, review={\MR {1298949}}, }
Reference [46]
D. Cordoba, D. Faraco, and F. Gancedo, Lack of uniqueness for weak solutions of the incompressible porous media equation, Arch. Ration. Mech. Anal. 200 (2011), no. 3, 725–746, DOI 10.1007/s00205-010-0365-z. MR2796131,
Show rawAMSref \bib{CordobaFaracoGancedo11}{article}{ author={Cordoba, Diego}, author={Faraco, Daniel}, author={Gancedo, Francisco}, title={Lack of uniqueness for weak solutions of the incompressible porous media equation}, journal={Arch. Ration. Mech. Anal.}, volume={200}, date={2011}, number={3}, pages={725--746}, issn={0003-9527}, review={\MR {2796131}}, doi={10.1007/s00205-010-0365-z}, }
Reference [47]
G. Crippa, N. Gusev, S. Spirito, and E. Wiedemann, Non-uniqueness and prescribed energy for the continuity equation, Commun. Math. Sci. 13 (2015), no. 7, 1937–1947, DOI 10.4310/CMS.2015.v13.n7.a12. MR3393180,
Show rawAMSref \bib{CrippaEtAl14}{article}{ author={Crippa, Gianluca}, author={Gusev, Nikolay}, author={Spirito, Stefano}, author={Wiedemann, Emil}, title={Non-uniqueness and prescribed energy for the continuity equation}, journal={Commun. Math. Sci.}, volume={13}, date={2015}, number={7}, pages={1937--1947}, issn={1539-6746}, review={\MR {3393180}}, doi={10.4310/CMS.2015.v13.n7.a12}, }
Reference [48]
M. Dai, Non-uniqueness of Leray–Hopf weak solutions of the 3d Hall-MHD system, arXiv:1812.11311 (2018).
Reference [49]
M. Dai and H. Liu, Anomalous dissipation of energy and magnetic helicity for the electron-mhd system, arXiv:1911.03953 (2019).
Reference [50]
V. Dallas and A. Alexakis, The signature of initial conditions on magnetohydrodynamic turbulence, The Astrophysical Journal Letters 788 (2014), no. 2, L36.
Reference [51]
S. Daneri and L. Székelyhidi Jr., Non-uniqueness and h-principle for Hölder-continuous weak solutions of the Euler equations, Arch. Ration. Mech. Anal. 224 (2017), no. 2, 471–514, DOI 10.1007/s00205-017-1081-8. MR3614753,
Show rawAMSref \bib{DSZ17}{article}{ author={Daneri, Sara}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={Non-uniqueness and h-principle for H\"{o}lder-continuous weak solutions of the Euler equations}, journal={Arch. Ration. Mech. Anal.}, volume={224}, date={2017}, number={2}, pages={471--514}, issn={0003-9527}, review={\MR {3614753}}, doi={10.1007/s00205-017-1081-8}, }
Reference [52]
P. A. Davidson, An introduction to magnetohydrodynamics, Cambridge Texts in Applied Mathematics, Cambridge University Press, Cambridge, 2001, DOI 10.1017/CBO9780511626333. MR1825486,
Show rawAMSref \bib{Davidson01}{book}{ author={Davidson, P. A.}, title={An introduction to magnetohydrodynamics}, series={Cambridge Texts in Applied Mathematics}, publisher={Cambridge University Press, Cambridge}, date={2001}, pages={xviii+431}, isbn={0-521-79487-0}, review={\MR {1825486}}, doi={10.1017/CBO9780511626333}, }
Reference [53]
P. A. Davidson, Y. Kaneda, K. Moffatt, and K. R. Sreenivasan (eds.), A voyage through turbulence, Cambridge University Press, Cambridge, 2011, DOI 10.1017/CBO9781139018241. MR2894536,
Show rawAMSref \bib{Voyage11}{collection}{ title={A voyage through turbulence}, editor={Davidson, Peter A.}, editor={Kaneda, Yukio}, editor={Moffatt, Keith}, editor={Sreenivasan, Katepalli R.}, publisher={Cambridge University Press, Cambridge}, date={2011}, pages={xvi+434}, isbn={978-0-521-14931-0}, review={\MR {2894536}}, doi={10.1017/CBO9781139018241}, }
Reference [54]
C. De Lellis and H. Kwon, On non-uniqueness of Hölder continuous globally dissipative Euler flows, arXiv preprint arXiv:2006.06482 (2020).
Reference [55]
C. De Lellis and L. Székelyhidi Jr., The Euler equations as a differential inclusion, Ann. of Math. (2) 170 (2009), no. 3, 1417–1436, DOI 10.4007/annals.2009.170.1417. MR2600877,
Show rawAMSref \bib{DLSZ09}{article}{ author={De Lellis, Camillo}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={The Euler equations as a differential inclusion}, journal={Ann. of Math. (2)}, volume={170}, date={2009}, number={3}, pages={1417--1436}, issn={0003-486X}, review={\MR {2600877}}, doi={10.4007/annals.2009.170.1417}, }
Reference [56]
C. De Lellis and L. Székelyhidi Jr., On admissibility criteria for weak solutions of the Euler equations, Arch. Ration. Mech. Anal. 195 (2010), no. 1, 225–260, DOI 10.1007/s00205-008-0201-x. MR2564474,
Show rawAMSref \bib{DeLellisSzekelyhidi10}{article}{ author={De Lellis, Camillo}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={On admissibility criteria for weak solutions of the Euler equations}, journal={Arch. Ration. Mech. Anal.}, volume={195}, date={2010}, number={1}, pages={225--260}, issn={0003-9527}, review={\MR {2564474}}, doi={10.1007/s00205-008-0201-x}, }
Reference [57]
C. De Lellis and L. Székelyhidi Jr., The -principle and the equations of fluid dynamics, Bull. Amer. Math. Soc. (N.S.) 49 (2012), no. 3, 347–375, DOI 10.1090/S0273-0979-2012-01376-9. MR2917063,
Show rawAMSref \bib{DLSZ12}{article}{ author={De Lellis, Camillo}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={The $h$-principle and the equations of fluid dynamics}, journal={Bull. Amer. Math. Soc. (N.S.)}, volume={49}, date={2012}, number={3}, pages={347--375}, issn={0273-0979}, review={\MR {2917063}}, doi={10.1090/S0273-0979-2012-01376-9}, }
Reference [58]
C. De Lellis and L. Székelyhidi Jr., Dissipative continuous Euler flows, Invent. Math. 193 (2013), no. 2, 377–407, DOI 10.1007/s00222-012-0429-9. MR3090182,
Show rawAMSref \bib{DLSZ13}{article}{ author={De Lellis, Camillo}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={Dissipative continuous Euler flows}, journal={Invent. Math.}, volume={193}, date={2013}, number={2}, pages={377--407}, issn={0020-9910}, review={\MR {3090182}}, doi={10.1007/s00222-012-0429-9}, }
Reference [59]
C. De Lellis and L. Székelyhidi Jr., Dissipative Euler flows and Onsager’s conjecture, J. Eur. Math. Soc. (JEMS) 16 (2014), no. 7, 1467–1505, DOI 10.4171/JEMS/466. MR3254331,
Show rawAMSref \bib{DeLellisSzekelyhidi12a}{article}{ author={De Lellis, Camillo}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={Dissipative Euler flows and Onsager's conjecture}, journal={J. Eur. Math. Soc. (JEMS)}, volume={16}, date={2014}, number={7}, pages={1467--1505}, issn={1435-9855}, review={\MR {3254331}}, doi={10.4171/JEMS/466}, }
Reference [60]
C. De Lellis and L. Székelyhidi Jr., High dimensionality and h-principle in PDE, Bull. Amer. Math. Soc. (N.S.) 54 (2017), no. 2, 247–282, DOI 10.1090/bull/1549. MR3619726,
Show rawAMSref \bib{DLSZ17}{article}{ author={De Lellis, Camillo}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={High dimensionality and h-principle in PDE}, journal={Bull. Amer. Math. Soc. (N.S.)}, volume={54}, date={2017}, number={2}, pages={247--282}, issn={0273-0979}, review={\MR {3619726}}, doi={10.1090/bull/1549}, }
Reference [61]
C. De Lellis and L. Székelyhidi Jr., On turbulence and geometry: from Nash to Onsager, Notices Amer. Math. Soc. 66 (2019), no. 5, 677–685. MR3929468,
Show rawAMSref \bib{DLSZ19}{article}{ author={De Lellis, Camillo}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={On turbulence and geometry: from Nash to Onsager}, journal={Notices Amer. Math. Soc.}, volume={66}, date={2019}, number={5}, pages={677--685}, issn={0002-9920}, review={\MR {3929468}}, }
Reference [62]
L. De Rosa, Infinitely many Leray–Hopf solutions for the fractional Navier–Stokes equations, Comm. Partial Differential Equations 44 (2019), no. 4, 335–365, DOI 10.1080/03605302.2018.1547745. MR3941228,
Show rawAMSref \bib{DR19}{article}{ author={De Rosa, Luigi}, title={Infinitely many Leray--Hopf solutions for the fractional Navier--Stokes equations}, journal={Comm. Partial Differential Equations}, volume={44}, date={2019}, number={4}, pages={335--365}, issn={0360-5302}, review={\MR {3941228}}, doi={10.1080/03605302.2018.1547745}, }
Reference [63]
L. De Rosa, On the helicity conservation for the incompressible Euler equations, Proc. Amer. Math. Soc. 148 (2020), no. 7, 2969–2979, DOI 10.1090/proc/14952. MR4099784,
Show rawAMSref \bib{DeRosa20}{article}{ author={De Rosa, Luigi}, title={On the helicity conservation for the incompressible Euler equations}, journal={Proc. Amer. Math. Soc.}, volume={148}, date={2020}, number={7}, pages={2969--2979}, issn={0002-9939}, review={\MR {4099784}}, doi={10.1090/proc/14952}, }
Reference [64]
R. J. DiPerna, Compensated compactness and general systems of conservation laws, Trans. Amer. Math. Soc. 292 (1985), no. 2, 383–420, DOI 10.2307/2000221. MR808729,
Show rawAMSref \bib{diperna_compensated_1985}{article}{ author={DiPerna, Ronald J.}, title={Compensated compactness and general systems of conservation laws}, journal={Trans. Amer. Math. Soc.}, volume={292}, date={1985}, number={2}, pages={383--420}, issn={0002-9947}, review={\MR {808729}}, doi={10.2307/2000221}, }
Reference [65]
R. J. DiPerna and P.-L. Lions, Ordinary differential equations, transport theory and Sobolev spaces, Invent. Math. 98 (1989), no. 3, 511–547, DOI 10.1007/BF01393835. MR1022305,
Show rawAMSref \bib{DiPernaLions89}{article}{ author={DiPerna, R. J.}, author={Lions, P.-L.}, title={Ordinary differential equations, transport theory and Sobolev spaces}, journal={Invent. Math.}, volume={98}, date={1989}, number={3}, pages={511--547}, issn={0020-9910}, review={\MR {1022305}}, doi={10.1007/BF01393835}, }
Reference [66]
R. J. DiPerna and A. J. Majda, Oscillations and concentrations in weak solutions of the incompressible fluid equations, Comm. Math. Phys. 108 (1987), no. 4, 667–689. MR877643,
Show rawAMSref \bib{DiPernaMajda87}{article}{ author={DiPerna, Ronald J.}, author={Majda, Andrew J.}, title={Oscillations and concentrations in weak solutions of the incompressible fluid equations}, journal={Comm. Math. Phys.}, volume={108}, date={1987}, number={4}, pages={667--689}, issn={0010-3616}, review={\MR {877643}}, }
Reference [67]
C. R. Doering and J. D. Gibbon, Applied analysis of the Navier–Stokes equations, Cambridge Texts in Applied Mathematics, Cambridge University Press, Cambridge, 1995, DOI 10.1017/CBO9780511608803. MR1325465,
Show rawAMSref \bib{DoeringGibbon95}{book}{ author={Doering, Charles R.}, author={Gibbon, J. D.}, title={Applied analysis of the Navier--Stokes equations}, series={Cambridge Texts in Applied Mathematics}, publisher={Cambridge University Press, Cambridge}, date={1995}, pages={xiv+217}, isbn={0-521-44557-4}, isbn={0-521-44568-X}, review={\MR {1325465}}, doi={10.1017/CBO9780511608803}, }
Reference [68]
J. Duchon and R. Robert, Inertial energy dissipation for weak solutions of incompressible Euler and Navier–Stokes equations, Nonlinearity 13 (2000), no. 1, 249–255, DOI 10.1088/0951-7715/13/1/312. MR1734632,
Show rawAMSref \bib{DuchonRobert00}{article}{ author={Duchon, Jean}, author={Robert, Raoul}, title={Inertial energy dissipation for weak solutions of incompressible Euler and Navier--Stokes equations}, journal={Nonlinearity}, volume={13}, date={2000}, number={1}, pages={249--255}, issn={0951-7715}, review={\MR {1734632}}, doi={10.1088/0951-7715/13/1/312}, }
Reference [69]
G. Duvaut and J.-L. Lions, Inéquations en thermoélasticité et magnétohydrodynamique (French), Arch. Rational Mech. Anal. 46 (1972), 241–279, DOI 10.1007/BF00250512. MR346289,
Show rawAMSref \bib{DuvautLions72}{article}{ author={Duvaut, G.}, author={Lions, J.-L.}, title={In\'{e}quations en thermo\'{e}lasticit\'{e} et magn\'{e}tohydrodynamique}, language={French}, journal={Arch. Rational Mech. Anal.}, volume={46}, date={1972}, pages={241--279}, issn={0003-9527}, review={\MR {346289}}, doi={10.1007/BF00250512}, }
Reference [70]
Y. Eliashberg and N. Mishachev, Introduction to the -principle, Graduate Studies in Mathematics, vol. 48, American Mathematical Society, Providence, RI, 2002, DOI 10.1090/gsm/048. MR1909245,
Show rawAMSref \bib{EliashbergMishachev02}{book}{ author={Eliashberg, Y.}, author={Mishachev, N.}, title={Introduction to the $h$-principle}, series={Graduate Studies in Mathematics}, volume={48}, publisher={American Mathematical Society, Providence, RI}, date={2002}, pages={xviii+206}, isbn={0-8218-3227-1}, review={\MR {1909245}}, doi={10.1090/gsm/048}, }
Reference [71]
A. Enciso, D. Peralta-Salas, and F. Torres de Lizaur, Helicity is the only integral invariant of volume-preserving transformations, Proc. Natl. Acad. Sci. USA 113 (2016), no. 8, 2035–2040, DOI 10.1073/pnas.1516213113. MR3474089,
Show rawAMSref \bib{EncisoEtAl16}{article}{ author={Enciso, Alberto}, author={Peralta-Salas, Daniel}, author={Torres de Lizaur, Francisco}, title={Helicity is the only integral invariant of volume-preserving transformations}, journal={Proc. Natl. Acad. Sci. USA}, volume={113}, date={2016}, number={8}, pages={2035--2040}, issn={0027-8424}, review={\MR {3474089}}, doi={10.1073/pnas.1516213113}, }
Reference [72]
L. Eskauriaza, G. A. Seregin, and V. Sverak, -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{EscauriazaSerginSverak03}{article}{ author={Eskauriaza, L.}, author={Seregin, G. A.}, author={Sverak, 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 [73]
G. L. Eyink, Energy dissipation without viscosity in ideal hydrodynamics. I. Fourier analysis and local energy transfer, Phys. D 78 (1994), no. 3-4, 222–240, DOI 10.1016/0167-2789(94)90117-1. MR1302409,
Show rawAMSref \bib{Eyink94}{article}{ author={Eyink, Gregory L.}, title={Energy dissipation without viscosity in ideal hydrodynamics. I. Fourier analysis and local energy transfer}, journal={Phys. D}, volume={78}, date={1994}, number={3-4}, pages={222--240}, issn={0167-2789}, review={\MR {1302409}}, doi={10.1016/0167-2789(94)90117-1}, }
Reference [74]
G. L. Eyink and K. R. Sreenivasan, Onsager and the theory of hydrodynamic turbulence, Rev. Modern Phys. 78 (2006), no. 1, 87–135, DOI 10.1103/RevModPhys.78.87. MR2214822,
Show rawAMSref \bib{EyinkSreeniviasan06}{article}{ author={Eyink, Gregory L.}, author={Sreenivasan, Katepalli R.}, title={Onsager and the theory of hydrodynamic turbulence}, journal={Rev. Modern Phys.}, volume={78}, date={2006}, number={1}, pages={87--135}, issn={0034-6861}, review={\MR {2214822}}, doi={10.1103/RevModPhys.78.87}, }
Reference [75]
E. B. Fabes, B. F. Jones, and N. M. Rivière, The initial value problem for the Navier–Stokes equations with data in , Arch. Rational Mech. Anal. 45 (1972), 222–240, DOI 10.1007/BF00281533. MR316915,
Show rawAMSref \bib{FabesJonesRiviere72}{article}{ author={Fabes, E. B.}, author={Jones, B. F.}, author={Rivi\`ere, N. M.}, title={The initial value problem for the Navier--Stokes equations with data in $L^{p}$}, journal={Arch. Rational Mech. Anal.}, volume={45}, date={1972}, pages={222--240}, issn={0003-9527}, review={\MR {316915}}, doi={10.1007/BF00281533}, }
Reference [76]
D. Faraco and S. Lindberg, Magnetic helicity and subsolutions in ideal mhd, arXiv:1801.04896 (2018).
Reference [77]
D. Faraco and S. Lindberg, Proof of Taylor’s conjecture on magnetic helicity conservation, Comm. Math. Phys. 373 (2020), no. 2, 707–738, DOI 10.1007/s00220-019-03422-7. MR4056647,
Show rawAMSref \bib{FaracoLindberg18b}{article}{ author={Faraco, Daniel}, author={Lindberg, Sauli}, title={Proof of Taylor's conjecture on magnetic helicity conservation}, journal={Comm. Math. Phys.}, volume={373}, date={2020}, number={2}, pages={707--738}, issn={0010-3616}, review={\MR {4056647}}, doi={10.1007/s00220-019-03422-7}, }
Reference [78]
D. Faraco, S. Lindberg, and L. Székelyhidi, Jr., Bounded solutions of ideal MHD with compact support in space-time, arXiv preprint arXiv:1909.08678 (2019).
Reference [79]
C. Foiaş, Statistical study of Navier–Stokes equations. I, Rend. Sem. Mat. Univ. Padova 48 (1972), 219–348. MR352733,
Show rawAMSref \bib{Foias72}{article}{ author={Foia\c {s}, C.}, title={Statistical study of Navier--Stokes equations. I}, journal={Rend. Sem. Mat. Univ. Padova}, volume={48}, date={1972}, pages={219--348}, issn={0041-8994}, review={\MR {352733}}, }
Reference [80]
C. Foiaş, Statistical study of Navier–Stokes equations. II, Rend. Sem. Mat. Univ. Padova 49 (1973), 9–123. MR352733,
Show rawAMSref \bib{Foias73}{article}{ author={Foia\c {s}, C.}, title={Statistical study of Navier--Stokes equations. II}, journal={Rend. Sem. Mat. Univ. Padova}, volume={49}, date={1973}, pages={9--123}, issn={0041-8994}, review={\MR {352733}}, }
Reference [81]
C. Foias, O. Manley, R. Rosa, and R. Temam, Navier–Stokes equations and turbulence, Encyclopedia of Mathematics and its Applications, vol. 83, Cambridge University Press, Cambridge, 2001, DOI 10.1017/CBO9780511546754. MR1855030,
Show rawAMSref \bib{FoiasManleyRosaTemam01}{book}{ author={Foias, C.}, author={Manley, O.}, author={Rosa, R.}, author={Temam, R.}, title={Navier--Stokes equations and turbulence}, series={Encyclopedia of Mathematics and its Applications}, volume={83}, publisher={Cambridge University Press, Cambridge}, date={2001}, pages={xiv+347}, isbn={0-521-36032-3}, review={\MR {1855030}}, doi={10.1017/CBO9780511546754}, }
Reference [82]
U. Frisch, Turbulence, Cambridge University Press, Cambridge, 1995. The legacy of A. N. Kolmogorov. MR1428905,
Show rawAMSref \bib{Frisch95}{book}{ author={Frisch, Uriel}, title={Turbulence}, note={The legacy of A. N. Kolmogorov}, publisher={Cambridge University Press, Cambridge}, date={1995}, pages={xiv+296}, isbn={0-521-45103-5}, review={\MR {1428905}}, }
Reference [83]
H. Fujita and T. 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{FujitaKato64}{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 [84]
G. Furioli, P. G. Lemarié-Rieusset, and E. Terraneo, Unicité dans et d’autres espaces fonctionnels limites pour Navier–Stokes (French, with English and French summaries), Rev. Mat. Iberoamericana 16 (2000), no. 3, 605–667, DOI 10.4171/RMI/286. MR1813331,
Show rawAMSref \bib{FurioliLemarieRieussetTerraneo00}{article}{ author={Furioli, Giulia}, author={Lemari\'{e}-Rieusset, Pierre G.}, author={Terraneo, Elide}, title={Unicit\'{e} dans $L^3(\mathbb {R}^3)$ et d'autres espaces fonctionnels limites pour Navier--Stokes}, language={French, with English and French summaries}, journal={Rev. Mat. Iberoamericana}, volume={16}, date={2000}, number={3}, pages={605--667}, issn={0213-2230}, review={\MR {1813331}}, doi={10.4171/RMI/286}, }
Reference [85]
P. Germain, Multipliers, paramultipliers, and weak-strong uniqueness for the Navier–Stokes equations, J. Differential Equations 226 (2006), no. 2, 373–428, DOI 10.1016/j.jde.2005.10.007. MR2237686,
Show rawAMSref \bib{Germain06}{article}{ author={Germain, Pierre}, title={Multipliers, paramultipliers, and weak-strong uniqueness for the Navier--Stokes equations}, journal={J. Differential Equations}, volume={226}, date={2006}, number={2}, pages={373--428}, issn={0022-0396}, review={\MR {2237686}}, doi={10.1016/j.jde.2005.10.007}, }
Reference [86]
M. L. Gromov, Convex integration of differential relations. I (Russian), Izv. Akad. Nauk SSSR Ser. Mat. 37 (1973), 329–343. MR0413206,
Show rawAMSref \bib{Gromov73}{article}{ author={Gromov, M. L.}, title={Convex integration of differential relations. I}, language={Russian}, journal={Izv. Akad. Nauk SSSR Ser. Mat.}, volume={37}, date={1973}, pages={329--343}, issn={0373-2436}, review={\MR {0413206}}, }
Reference [87]
J. Guillod and V. Šverák, Numerical investigations of non-uniqueness for the Navier–Stokes initial value problem in borderline spaces, arXiv:1704.00560 (2017).
Reference [88]
M. Hofmanova, R. Zhu, and X. Zhu, Non-uniqueness in law of stochastic 3d Navier–Stokes equations, arXiv:1912.11841 (2019).
Reference [89]
E. Hopf, Über die Anfangswertaufgabe für die hydrodynamischen Grundgleichungen (German), Math. Nachr. 4 (1951), 213–231, DOI 10.1002/mana.3210040121. MR50423,
Show rawAMSref \bib{Hopf51}{article}{ author={Hopf, Eberhard}, title={\"{U}ber die Anfangswertaufgabe f\"{u}r die hydrodynamischen Grundgleichungen}, language={German}, journal={Math. Nachr.}, volume={4}, date={1951}, pages={213--231}, issn={0025-584X}, review={\MR {50423}}, doi={10.1002/mana.3210040121}, }
Reference [90]
P. Isett, Hölder continuous Euler flows with compact support in time, ProQuest LLC, Ann Arbor, MI, 2013. Thesis (Ph.D.)–Princeton University. MR3153420,
Show rawAMSref \bib{IsettThesis}{book}{ author={Isett, Philip}, title={H\"{o}lder continuous Euler flows with compact support in time}, note={Thesis (Ph.D.)--Princeton University}, publisher={ProQuest LLC, Ann Arbor, MI}, date={2013}, pages={227}, isbn={978-1303-09748-5}, review={\MR {3153420}}, }
Reference [91]
P. Isett, Nonuniqueness and existence of continuous, globally dissipative Euler flows, arXiv:1710.11186 (2017).
Reference [92]
P. Isett, On the endpoint regularity in Onsager’s conjecture, arXiv:1706.01549 (2017).
Reference [93]
P. Isett, A proof of Onsager’s conjecture, Ann. of Math. (2) 188 (2018), no. 3, 871–963, DOI 10.4007/annals.2018.188.3.4. MR3866888,
Show rawAMSref \bib{Isett18}{article}{ author={Isett, Philip}, title={A proof of Onsager's conjecture}, journal={Ann. of Math. (2)}, volume={188}, date={2018}, number={3}, pages={871--963}, issn={0003-486X}, review={\MR {3866888}}, doi={10.4007/annals.2018.188.3.4}, }
Reference [94]
P. Isett and A. Ma, A direct approach to nonuniqueness and failure of compactness for the SQG equation, arXiv:2007.03078 (2020).
Reference [95]
P. Isett and S.-J. Oh, On nonperiodic Euler flows with Hölder regularity, Arch. Ration. Mech. Anal. 221 (2016), no. 2, 725–804, DOI 10.1007/s00205-016-0973-3. MR3488536,
Show rawAMSref \bib{IsettOh16}{article}{ author={Isett, Philip}, author={Oh, Sung-Jin}, title={On nonperiodic Euler flows with H\"{o}lder regularity}, journal={Arch. Ration. Mech. Anal.}, volume={221}, date={2016}, number={2}, pages={725--804}, issn={0003-9527}, review={\MR {3488536}}, doi={10.1007/s00205-016-0973-3}, }
Reference [96]
P. Isett and V. Vicol, Hölder continuous solutions of active scalar equations, Ann. PDE 1 (2015), no. 1, Art. 2, 77, DOI 10.1007/s40818-015-0002-0. MR3479065,
Show rawAMSref \bib{IsettVicol15}{article}{ author={Isett, Philip}, author={Vicol, Vlad}, title={H\"{o}lder continuous solutions of active scalar equations}, journal={Ann. PDE}, volume={1}, date={2015}, number={1}, pages={Art. 2, 77}, issn={2524-5317}, review={\MR {3479065}}, doi={10.1007/s40818-015-0002-0}, }
Reference [97]
H. Jia and V. Šverák, Local-in-space estimates near initial time for weak solutions of the Navier–Stokes equations and forward self-similar solutions, Invent. Math. 196 (2014), no. 1, 233–265, DOI 10.1007/s00222-013-0468-x. MR3179576,
Show rawAMSref \bib{JiaSverak14}{article}{ author={Jia, Hao}, author={\v {S}ver\'{a}k, Vladim\'{\i }r}, title={Local-in-space estimates near initial time for weak solutions of the Navier--Stokes equations and forward self-similar solutions}, journal={Invent. Math.}, volume={196}, date={2014}, number={1}, pages={233--265}, issn={0020-9910}, review={\MR {3179576}}, doi={10.1007/s00222-013-0468-x}, }
Reference [98]
H. Jia and V. Sverak, Are the incompressible 3d Navier–Stokes equations locally ill-posed in the natural energy space?, J. Funct. Anal. 268 (2015), no. 12, 3734–3766, DOI 10.1016/j.jfa.2015.04.006. MR3341963,
Show rawAMSref \bib{JiaSverak15}{article}{ author={Jia, Hao}, author={Sverak, Vladimir}, title={Are the incompressible 3d Navier--Stokes equations locally ill-posed in the natural energy space?}, journal={J. Funct. Anal.}, volume={268}, date={2015}, number={12}, pages={3734--3766}, issn={0022-1236}, review={\MR {3341963}}, doi={10.1016/j.jfa.2015.04.006}, }
Reference [99]
Y. Kaneda, T. Ishihara, M. Yokokawa, K. Itakura, and A. Uno, Energy dissipation rate and energy spectrum in high resolution direct numerical simulations of turbulence in a periodic box, Physics of Fluids 15 (2003), no. 2, L21–L24.
Reference [100]
E. Kang and J. Lee, Remarks on the magnetic helicity and energy conservation for ideal magneto-hydrodynamics, Nonlinearity 20 (2007), no. 11, 2681–2689, DOI 10.1088/0951-7715/20/11/011. MR2361251,
Show rawAMSref \bib{KangLee07}{article}{ author={Kang, Ensil}, author={Lee, Jihoon}, title={Remarks on the magnetic helicity and energy conservation for ideal magneto-hydrodynamics}, journal={Nonlinearity}, volume={20}, date={2007}, number={11}, pages={2681--2689}, issn={0951-7715}, review={\MR {2361251}}, doi={10.1088/0951-7715/20/11/011}, }
Reference [101]
B. Khesin, D. Peralta-Salas, and C. Yang, A basis of casimirs in 3d magnetohydrodynamics, arXiv:1901.04404 (2019).
Reference [102]
B. Kirchheim, S. Müller, and V. Šverák, Studying nonlinear pde by geometry in matrix space, Geometric analysis and nonlinear partial differential equations, Springer, Berlin, 2003, pp. 347–395. MR2008346,
Show rawAMSref \bib{KircheimMullerSverak03}{article}{ author={Kirchheim, Bernd}, author={M\"{u}ller, Stefan}, author={\v {S}ver\'{a}k, Vladim\'{\i }r}, title={Studying nonlinear pde by geometry in matrix space}, conference={ title={Geometric analysis and nonlinear partial differential equations}, }, book={ publisher={Springer, Berlin}, }, date={2003}, pages={347--395}, review={\MR {2008346}}, }
Reference [103]
A. A. Kiselev and O. A. Ladyženskaya, On the existence and uniqueness of the solution of the nonstationary problem for a viscous, incompressible fluid (Russian), Izv. Akad. Nauk SSSR. Ser. Mat. 21 (1957), 655–680. MR0100448,
Show rawAMSref \bib{KiselevLadyzhenskaya57}{article}{ author={Kiselev, A. A.}, author={Lady\v {z}enskaya, O. A.}, title={On the existence and uniqueness of the solution of the nonstationary problem for a viscous, incompressible fluid}, language={Russian}, journal={Izv. Akad. Nauk SSSR. Ser. Mat.}, volume={21}, date={1957}, pages={655--680}, issn={0373-2436}, review={\MR {0100448}}, }
Reference [104]
S. Klainerman, On Nash’s unique contribution to analysis in just three of his papers, Bull. Amer. Math. Soc. (N.S.) 54 (2017), no. 2, 283–305, DOI 10.1090/bull/1560. MR3619727,
Show rawAMSref \bib{Klainerman17}{article}{ author={Klainerman, Sergiu}, title={On Nash's unique contribution to analysis in just three of his papers}, journal={Bull. Amer. Math. Soc. (N.S.)}, volume={54}, date={2017}, number={2}, pages={283--305}, issn={0273-0979}, review={\MR {3619727}}, doi={10.1090/bull/1560}, }
Reference [105]
A. N. Kolmogoroff, Dissipation of energy in the locally isotropic turbulence, C. R. (Doklady) Acad. Sci. URSS (N.S.) 32 (1941), 16–18. MR0005851,
Show rawAMSref \bib{Kolmogorov41c}{article}{ author={Kolmogoroff, A. N.}, title={Dissipation of energy in the locally isotropic turbulence}, journal={C. R. (Doklady) Acad. Sci. URSS (N.S.)}, volume={32}, date={1941}, pages={16--18}, review={\MR {0005851}}, }
Reference [106]
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{Kolmogorov41a}{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 [107]
A. N. Kolmogoroff, On degeneration of isotropic turbulence in an incompressible viscous liquid, C. R. (Doklady) Acad. Sci. URSS (N. S.) 31 (1941), 538–540. MR0004568,
Show rawAMSref \bib{Kolmogorov41b}{article}{ author={Kolmogoroff, A. N.}, title={On degeneration of isotropic turbulence in an incompressible viscous liquid}, journal={C. R. (Doklady) Acad. Sci. URSS (N. S.)}, volume={31}, date={1941}, pages={538--540}, review={\MR {0004568}}, }
Reference [108]
N. H. Kuiper, On -isometric imbeddings. I, II, Nederl. Akad. Wetensch. Proc. Ser. A. 58 = Indag. Math. 17 (1955), 545–556, 683–689. MR0075640,
Show rawAMSref \bib{Kuiper55}{article}{ author={Kuiper, Nicolaas H.}, title={On $C^1$-isometric imbeddings. I, II}, journal={Nederl. Akad. Wetensch. Proc. Ser. A. {\bf 58} = Indag. Math.}, volume={17}, date={1955}, pages={545--556, 683--689}, review={\MR {0075640}}, }
Reference [109]
I. Kukavica, Pressure integrability conditions for uniqueness of mild solutions of the Navier–Stokes system, J. Differential Equations 223 (2006), no. 2, 427–441, DOI 10.1016/j.jde.2005.07.005. MR2214942,
Show rawAMSref \bib{Kukavica06b}{article}{ author={Kukavica, Igor}, title={Pressure integrability conditions for uniqueness of mild solutions of the Navier--Stokes system}, journal={J. Differential Equations}, volume={223}, date={2006}, number={2}, pages={427--441}, issn={0022-0396}, review={\MR {2214942}}, doi={10.1016/j.jde.2005.07.005}, }
Reference [110]
I. Kukavica, Role of the pressure for validity of the energy equality for solutions of the Navier–Stokes equation, J. Dynam. Differential Equations 18 (2006), no. 2, 461–482, DOI 10.1007/s10884-006-9010-9. MR2229985,
Show rawAMSref \bib{Kukavica06}{article}{ author={Kukavica, Igor}, title={Role of the pressure for validity of the energy equality for solutions of the Navier--Stokes equation}, journal={J. Dynam. Differential Equations}, volume={18}, date={2006}, number={2}, pages={461--482}, issn={1040-7294}, review={\MR {2229985}}, doi={10.1007/s10884-006-9010-9}, }
Reference [111]
I. Kukavica, The fractal dimension of the singular set for solutions of the Navier–Stokes system, Nonlinearity 22 (2009), no. 12, 2889–2900, DOI 10.1088/0951-7715/22/12/005. MR2557452,
Show rawAMSref \bib{Kukavica09}{article}{ author={Kukavica, Igor}, title={The fractal dimension of the singular set for solutions of the Navier--Stokes system}, journal={Nonlinearity}, volume={22}, date={2009}, number={12}, pages={2889--2900}, issn={0951-7715}, review={\MR {2557452}}, doi={10.1088/0951-7715/22/12/005}, }
Reference [112]
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{Ladyzhenskaya67}{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 [113]
P. G. Lemarié-Rieusset, Recent developments in the Navier–Stokes problem, Chapman & Hall/CRC Research Notes in Mathematics, vol. 431, Chapman & Hall/CRC, Boca Raton, FL, 2002, DOI 10.1201/9781420035674. MR1938147,
Show rawAMSref \bib{LemarieRieusset02}{book}{ author={Lemari\'{e}-Rieusset, P. G.}, title={Recent developments in the Navier--Stokes problem}, series={Chapman \& Hall/CRC Research Notes in Mathematics}, volume={431}, publisher={Chapman \& Hall/CRC, Boca Raton, FL}, date={2002}, pages={xiv+395}, isbn={1-58488-220-4}, review={\MR {1938147}}, doi={10.1201/9781420035674}, }
Reference [114]
P. G. Lemarié-Rieusset, The Navier–Stokes problem in the 21st century, CRC Press, Boca Raton, FL, 2016, DOI 10.1201/b19556. MR3469428,
Show rawAMSref \bib{Lemarie16}{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 [115]
J. 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{Leray34}{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 [116]
T. M. Leslie and R. Shvydkoy, Conditions implying energy equality for weak solutions of the Navier–Stokes equations, SIAM J. Math. Anal. 50 (2018), no. 1, 870–890, DOI 10.1137/16M1104147. MR3759872,
Show rawAMSref \bib{LeslieShvydkoy18}{article}{ author={Leslie, Trevor M.}, author={Shvydkoy, Roman}, title={Conditions implying energy equality for weak solutions of the Navier--Stokes equations}, journal={SIAM J. Math. Anal.}, volume={50}, date={2018}, number={1}, pages={870--890}, issn={0036-1410}, review={\MR {3759872}}, doi={10.1137/16M1104147}, }
Reference [117]
J.-L. Lions, Quelques résultats d’existence dans des équations aux dérivées partielles non linéaires (French), Bull. Soc. Math. France 87 (1959), 245–273. MR115019,
Show rawAMSref \bib{Lions59}{article}{ author={Lions, Jacques-Louis}, title={Quelques r\'{e}sultats d'existence dans des \'{e}quations aux d\'{e}riv\'{e}es partielles non lin\'{e}aires}, language={French}, journal={Bull. Soc. Math. France}, volume={87}, date={1959}, pages={245--273}, issn={0037-9484}, review={\MR {115019}}, }
Reference [118]
J. L. Lions, Sur la régularité et l’unicité des solutions turbulentes des équations de Navier Stokes (French), Rend. Sem. Mat. Univ. Padova 30 (1960), 16–23. MR115018,
Show rawAMSref \bib{Lions60}{article}{ author={Lions, J. L.}, title={Sur la r\'{e}gularit\'{e} et l'unicit\'{e} des solutions turbulentes des \'{e}quations de Navier Stokes}, language={French}, journal={Rend. Sem. Mat. Univ. Padova}, volume={30}, date={1960}, pages={16--23}, issn={0041-8994}, review={\MR {115018}}, }
Reference [119]
P.-L. Lions, Mathematical topics in fluid mechanics. Vol. 2, Oxford Lecture Series in Mathematics and its Applications, vol. 10, The Clarendon Press, Oxford University Press, New York, 1998. Compressible models; Oxford Science Publications. MR1637634,
Show rawAMSref \bib{Lions96}{book}{ author={Lions, Pierre-Louis}, title={Mathematical topics in fluid mechanics. Vol. 2}, series={Oxford Lecture Series in Mathematics and its Applications}, volume={10}, note={Compressible models; Oxford Science Publications}, publisher={The Clarendon Press, Oxford University Press, New York}, date={1998}, pages={xiv+348}, isbn={0-19-851488-3}, review={\MR {1637634}}, }
Reference [120]
P.-L. Lions and N. Masmoudi, Uniqueness of mild solutions of the Navier–Stokes system in , Comm. Partial Differential Equations 26 (2001), no. 11-12, 2211–2226, DOI 10.1081/PDE-100107819. MR1876415,
Show rawAMSref \bib{LionsMasmoudi01}{article}{ author={Lions, P.-L.}, author={Masmoudi, N.}, title={Uniqueness of mild solutions of the Navier--Stokes system in $L^N$}, journal={Comm. Partial Differential Equations}, volume={26}, date={2001}, number={11-12}, pages={2211--2226}, issn={0360-5302}, review={\MR {1876415}}, doi={10.1081/PDE-100107819}, }
Reference [121]
T. Luo and P. Qu, Non-uniqueness of weak solutions to 2D hypoviscous Navier–Stokes equations, J. Differential Equations 269 (2020), no. 4, 2896–2919, DOI 10.1016/j.jde.2020.02.014. MR4097236,
Show rawAMSref \bib{LuoQu20}{article}{ author={Luo, Tianwen}, author={Qu, Peng}, title={Non-uniqueness of weak solutions to 2D hypoviscous Navier--Stokes equations}, journal={J. Differential Equations}, volume={269}, date={2020}, number={4}, pages={2896--2919}, issn={0022-0396}, review={\MR {4097236}}, doi={10.1016/j.jde.2020.02.014}, }
Reference [122]
T. Luo and E. S. Titi, Non-uniqueness of weak solutions to hyperviscous Navier–Stokes equations: on sharpness of J.-L. Lions exponent, Calc. Var. Partial Differential Equations 59 (2020), no. 3, Paper No. 92, 15, DOI 10.1007/s00526-020-01742-4. MR4092687,
Show rawAMSref \bib{LuoTiti18}{article}{ author={Luo, Tianwen}, author={Titi, Edriss S.}, title={Non-uniqueness of weak solutions to hyperviscous Navier--Stokes equations: on sharpness of J.-L. Lions exponent}, journal={Calc. Var. Partial Differential Equations}, volume={59}, date={2020}, number={3}, pages={Paper No. 92, 15}, issn={0944-2669}, review={\MR {4092687}}, doi={10.1007/s00526-020-01742-4}, }
Reference [123]
X. Luo, Stationary solutions and nonuniqueness of weak solutions for the Navier–Stokes equations in high dimensions, Arch. Ration. Mech. Anal. 233 (2019), no. 2, 701–747, DOI 10.1007/s00205-019-01366-9. MR3951691,
Show rawAMSref \bib{Luo18}{article}{ author={Luo, Xiaoyutao}, title={Stationary solutions and nonuniqueness of weak solutions for the Navier--Stokes equations in high dimensions}, journal={Arch. Ration. Mech. Anal.}, volume={233}, date={2019}, number={2}, pages={701--747}, issn={0003-9527}, review={\MR {3951691}}, doi={10.1007/s00205-019-01366-9}, }
Reference [124]
A. J. Majda and A. L. Bertozzi, Vorticity and incompressible flow, Cambridge Texts in Applied Mathematics, vol. 27, Cambridge University Press, Cambridge, 2002. MR1867882,
Show rawAMSref \bib{MB2002}{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 [125]
C. Marchioro and M. Pulvirenti, Mathematical theory of incompressible nonviscous fluids, Applied Mathematical Sciences, vol. 96, Springer-Verlag, New York, 1994, DOI 10.1007/978-1-4612-4284-0. MR1245492,
Show rawAMSref \bib{MP12}{book}{ author={Marchioro, Carlo}, author={Pulvirenti, Mario}, title={Mathematical theory of incompressible nonviscous fluids}, series={Applied Mathematical Sciences}, volume={96}, publisher={Springer-Verlag, New York}, date={1994}, pages={xii+283}, isbn={0-387-94044-8}, review={\MR {1245492}}, doi={10.1007/978-1-4612-4284-0}, }
Reference [126]
C. Meneveau and K. R. Sreenivasan, The multifractal nature of turbulent energy dissipation, J. Fluid Mech. 224 (1991), 429–484.
Reference [127]
P. D. Mininni and A. Pouquet, Finite dissipation and intermittency in magnetohydrodynamics, Physical Review E 80 (2009), no. 2, 025401.
Reference [128]
S. Modena and G. Sattig, Convex integration solutions to the transport equation with full dimensional concentration, Ann. Inst. H. Poincaré Anal. Non Linéaire 37 (2020), no. 5, 1075–1108, DOI 10.1016/j.anihpc.2020.03.002. MR4138227,
Show rawAMSref \bib{ModenaSattig20}{article}{ author={Modena, Stefano}, author={Sattig, Gabriel}, title={Convex integration solutions to the transport equation with full dimensional concentration}, journal={Ann. Inst. H. Poincar\'{e} Anal. Non Lin\'{e}aire}, volume={37}, date={2020}, number={5}, pages={1075--1108}, issn={0294-1449}, review={\MR {4138227}}, doi={10.1016/j.anihpc.2020.03.002}, }
Reference [129]
S. Modena and L. Székelyhidi Jr., Non-uniqueness for the transport equation with Sobolev vector fields, Ann. PDE 4 (2018), no. 2, Paper No. 18, 38, DOI 10.1007/s40818-018-0056-x. MR3884855,
Show rawAMSref \bib{ModenaSZ17}{article}{ author={Modena, Stefano}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={Non-uniqueness for the transport equation with Sobolev vector fields}, journal={Ann. PDE}, volume={4}, date={2018}, number={2}, pages={Paper No. 18, 38}, issn={2524-5317}, review={\MR {3884855}}, doi={10.1007/s40818-018-0056-x}, }
Reference [130]
S. Modena and L. Székelyhidi Jr., Non-renormalized solutions to the continuity equation, Calculus of Variations and Partial Differential Equations 58 (2019), no. 6, 208.
Reference [131]
H. K. Moffatt, The degree of knottedness of tangled vortex lines, J. Fluid Mech. 35 (1969), no. 1, 117–129.
Reference [132]
H. K. Moffatt, Magnetic relaxation and the Taylor conjecture, Journal of Plasma Physics 81 (2015), no. 6.
Reference [133]
A. S. Monin and A. M. Yaglom, Statistical fluid mechanics: mechanics of turbulence. Vol. II, Dover Publications, Inc., Mineola, NY, 2007. Translated from the 1965 Russian original; Edited and with a preface by John L. Lumley; English edition updated, augmented and revised by the authors; Reprinted from the 1975 edition. MR2406668,
Show rawAMSref \bib{MoninYaglom13}{book}{ author={Monin, A. S.}, author={Yaglom, A. M.}, title={Statistical fluid mechanics: mechanics of turbulence. Vol. II}, note={Translated from the 1965 Russian original; Edited and with a preface by John L. Lumley; English edition updated, augmented and revised by the authors; Reprinted from the 1975 edition}, publisher={Dover Publications, Inc., Mineola, NY}, date={2007}, pages={xii+874}, isbn={978-0-486-45891-5}, isbn={0-486-45891-1}, review={\MR {2406668}}, }
Reference [134]
S. Müller and V. Šverák, Convex integration for Lipschitz mappings and counterexamples to regularity, Ann. of Math. (2) 157 (2003), no. 3, 715–742, DOI 10.4007/annals.2003.157.715. MR1983780,
Show rawAMSref \bib{MullerSverak03}{article}{ author={M\"{u}ller, S.}, author={\v {S}ver\'{a}k, V.}, title={Convex integration for Lipschitz mappings and counterexamples to regularity}, journal={Ann. of Math. (2)}, volume={157}, date={2003}, number={3}, pages={715--742}, issn={0003-486X}, review={\MR {1983780}}, doi={10.4007/annals.2003.157.715}, }
Reference [135]
J. Nash, isometric imbeddings, Ann. of Math. (2) 60 (1954), 383–396, DOI 10.2307/1969840. MR65993,
Show rawAMSref \bib{Nash54}{article}{ author={Nash, John}, title={$C^1$ isometric imbeddings}, journal={Ann. of Math. (2)}, volume={60}, date={1954}, pages={383--396}, issn={0003-486X}, review={\MR {65993}}, doi={10.2307/1969840}, }
Reference [136]
F. Nguyen, J.-P. Laval, P. Kestener, A. Cheskidov, R. Shvydkoy, and B. Dubrulle, Local estimates of Hölder exponents in turbulent vector fields, Physical Review E 99 (2019), no. 5, 053114.
Reference [137]
M. Novack, Nonuniqueness of weak solutions to the 3 dimensional quasi-geostrophic equations, SIAM J. Math. Anal. 52 (2020), no. 4, 3301–3349, DOI 10.1137/19M1281009. MR4126319,
Show rawAMSref \bib{Novack20}{article}{ author={Novack, Matthew}, title={Nonuniqueness of weak solutions to the 3 dimensional quasi-geostrophic equations}, journal={SIAM J. Math. Anal.}, volume={52}, date={2020}, number={4}, pages={3301--3349}, issn={0036-1410}, review={\MR {4126319}}, doi={10.1137/19M1281009}, }
Reference [138]
L. Onsager, Statistical hydrodynamics, Nuovo Cimento (9) 6 (1949), no. Suppl, Supplemento, no. 2 (Convegno Internazionale di Meccanica Statistica), 279–287. MR36116,
Show rawAMSref \bib{Onsager49}{article}{ author={Onsager, L.}, title={Statistical hydrodynamics}, journal={Nuovo Cimento (9)}, volume={6}, date={1949}, number={Suppl, Supplemento, no. 2 (Convegno Internazionale di Meccanica Statistica)}, pages={279--287}, issn={0029-6341}, review={\MR {36116}}, }
Reference [139]
S. Ortolani and D. Schnack, Magnetohydrodynamics of plasma relaxation, World Scientific, 1993.
Reference [140]
W. S. Ożański, The partial regularity theory of Caffarelli, Kohn, and Nirenberg and its sharpness, Lecture Notes in Mathematical Fluid Mechanics, Birkhäuser/Springer, Cham, 2019, DOI 10.1007/978-3-030-26661-5. MR3969978,
Show rawAMSref \bib{Ozanski19}{book}{ author={O\.{z}a\'{n}ski, Wojciech S.}, title={The partial regularity theory of Caffarelli, Kohn, and Nirenberg and its sharpness}, series={Lecture Notes in Mathematical Fluid Mechanics}, publisher={Birkh\"{a}user/Springer, Cham}, date={2019}, pages={vi+138}, isbn={978-3-030-26660-8}, isbn={978-3-030-26661-5}, review={\MR {3969978}}, doi={10.1007/978-3-030-26661-5}, }
Reference [141]
B. R. Pearson, P.-Å. Krogstad, and W. Van De Water, Measurements of the turbulent energy dissipation rate, Physics of Fluids 14 (2002), no. 3, 1288–1290.
Reference [142]
G. 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{Prodi59}{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 [143]
J. C. Robinson, J. L. Rodrigo, and W. Sadowski, The three-dimensional Navier–Stokes equations, Cambridge Studies in Advanced Mathematics, vol. 157, Cambridge University Press, Cambridge, 2016. Classical theory, DOI 10.1017/CBO9781139095143. MR3616490,
Show rawAMSref \bib{RSR16}{book}{ author={Robinson, James C.}, author={Rodrigo, Jos\'{e} L.}, author={Sadowski, Witold}, title={The three-dimensional Navier--Stokes equations}, series={Cambridge Studies in Advanced Mathematics}, volume={157}, note={Classical theory}, publisher={Cambridge University Press, Cambridge}, date={2016}, pages={xiv+471}, isbn={978-1-107-01966-9}, review={\MR {3616490}}, doi={10.1017/CBO9781139095143}, }
Reference [144]
J. C. Robinson and W. Sadowski, Decay of weak solutions and the singular set of the three-dimensional Navier–Stokes equations, Nonlinearity 20 (2007), no. 5, 1185–1191, DOI 10.1088/0951-7715/20/5/007. MR2312388,
Show rawAMSref \bib{RobinsonSadowski07}{article}{ author={Robinson, James C.}, author={Sadowski, Witold}, title={Decay of weak solutions and the singular set of the three-dimensional Navier--Stokes equations}, journal={Nonlinearity}, volume={20}, date={2007}, number={5}, pages={1185--1191}, issn={0951-7715}, review={\MR {2312388}}, doi={10.1088/0951-7715/20/5/007}, }
Reference [145]
V. Scheffer, Partial regularity of solutions to the Navier–Stokes equations, Pacific J. Math. 66 (1976), no. 2, 535–552. MR454426,
Show rawAMSref \bib{Scheffer76}{article}{ author={Scheffer, Vladimir}, title={Partial regularity of solutions to the Navier--Stokes equations}, journal={Pacific J. Math.}, volume={66}, date={1976}, number={2}, pages={535--552}, issn={0030-8730}, review={\MR {454426}}, }
Reference [146]
V. Scheffer, Hausdorff measure and the Navier–Stokes equations, Comm. Math. Phys. 55 (1977), no. 2, 97–112. MR510154,
Show rawAMSref \bib{Scheffer77}{article}{ author={Scheffer, Vladimir}, title={Hausdorff measure and the Navier--Stokes equations}, journal={Comm. Math. Phys.}, volume={55}, date={1977}, number={2}, pages={97--112}, issn={0010-3616}, review={\MR {510154}}, }
Reference [147]
V. Scheffer, An inviscid flow with compact support in space-time, J. Geom. Anal. 3 (1993), no. 4, 343–401, DOI 10.1007/BF02921318. MR1231007,
Show rawAMSref \bib{Scheffer93}{article}{ author={Scheffer, Vladimir}, title={An inviscid flow with compact support in space-time}, journal={J. Geom. Anal.}, volume={3}, date={1993}, number={4}, pages={343--401}, issn={1050-6926}, review={\MR {1231007}}, doi={10.1007/BF02921318}, }
Reference [148]
M. Sermange and R. Temam, Some mathematical questions related to the MHD equations, Comm. Pure Appl. Math. 36 (1983), no. 5, 635–664, DOI 10.1002/cpa.3160360506. MR716200,
Show rawAMSref \bib{SermangeTemam83}{article}{ author={Sermange, Michel}, author={Temam, Roger}, title={Some mathematical questions related to the MHD equations}, journal={Comm. Pure Appl. Math.}, volume={36}, date={1983}, number={5}, pages={635--664}, issn={0010-3640}, review={\MR {716200}}, doi={10.1002/cpa.3160360506}, }
Reference [149]
J. 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{Serrin62}{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 [150]
J. Serrin, The initial value problem for the Navier–Stokes equations, Nonlinear Problems (Proc. Sympos., Madison, Wis., 1962), Univ. of Wisconsin Press, Madison, Wis., 1963, pp. 69–98. MR0150444,
Show rawAMSref \bib{Serrin63}{article}{ author={Serrin, James}, title={The initial value problem for the Navier--Stokes equations}, conference={ title={Nonlinear Problems}, address={Proc. Sympos., Madison, Wis.}, date={1962}, }, book={ publisher={Univ. of Wisconsin Press, Madison, Wis.}, }, date={1963}, pages={69--98}, review={\MR {0150444}}, }
Reference [151]
M. Shinbrot, The energy equation for the Navier–Stokes system, SIAM J. Math. Anal. 5 (1974), 948–954, DOI 10.1137/0505092. MR435629,
Show rawAMSref \bib{Shinbrot74}{article}{ author={Shinbrot, Marvin}, title={The energy equation for the Navier--Stokes system}, journal={SIAM J. Math. Anal.}, volume={5}, date={1974}, pages={948--954}, issn={0036-1410}, review={\MR {435629}}, doi={10.1137/0505092}, }
Reference [152]
A. Shnirelman, Weak solutions with decreasing energy of incompressible Euler equations, Comm. Math. Phys. 210 (2000), no. 3, 541–603, DOI 10.1007/s002200050791. MR1777341,
Show rawAMSref \bib{Shnirelman00}{article}{ author={Shnirelman, A.}, title={Weak solutions with decreasing energy of incompressible Euler equations}, journal={Comm. Math. Phys.}, volume={210}, date={2000}, number={3}, pages={541--603}, issn={0010-3616}, review={\MR {1777341}}, doi={10.1007/s002200050791}, }
Reference [153]
A. Shnirelman, On the nonuniqueness of weak solution of the Euler equation, Comm. Pure Appl. Math. 50 (1997), no. 12, 1261–1286, DOI 10.1002/(SICI)1097-0312(199712)50:12<1261::AID-CPA3>3.3.CO;2-4. MR1476315,
Show rawAMSref \bib{Shnirelman97}{article}{ author={Shnirelman, A.}, title={On the nonuniqueness of weak solution of the Euler equation}, journal={Comm. Pure Appl. Math.}, volume={50}, date={1997}, number={12}, pages={1261--1286}, issn={0010-3640}, review={\MR {1476315}}, doi={10.1002/(SICI)1097-0312(199712)50:12<1261::AID-CPA3>3.3.CO;2-4}, }
Reference [154]
R. Shvydkoy, Lectures on the Onsager conjecture, Discrete Contin. Dyn. Syst. Ser. S 3 (2010), no. 3, 473–496, DOI 10.3934/dcdss.2010.3.473. MR2660721,
Show rawAMSref \bib{Shvydkoy10}{article}{ author={Shvydkoy, Roman}, title={Lectures on the Onsager conjecture}, journal={Discrete Contin. Dyn. Syst. Ser. S}, volume={3}, date={2010}, number={3}, pages={473--496}, issn={1937-1632}, review={\MR {2660721}}, doi={10.3934/dcdss.2010.3.473}, }
Reference [155]
R. Shvydkoy, Convex integration for a class of active scalar equations, J. Amer. Math. Soc. 24 (2011), no. 4, 1159–1174, DOI 10.1090/S0894-0347-2011-00705-4. MR2813340,
Show rawAMSref \bib{Shvydkoy11}{article}{ author={Shvydkoy, R.}, title={Convex integration for a class of active scalar equations}, journal={J. Amer. Math. Soc.}, volume={24}, date={2011}, number={4}, pages={1159--1174}, issn={0894-0347}, review={\MR {2813340}}, doi={10.1090/S0894-0347-2011-00705-4}, }
Reference [156]
R. Shvydkoy, Homogeneous solutions to the 3D Euler system, Trans. Amer. Math. Soc. 370 (2018), no. 4, 2517–2535, DOI 10.1090/tran/7022. MR3748575,
Show rawAMSref \bib{Shvydkoy18}{article}{ author={Shvydkoy, Roman}, title={Homogeneous solutions to the 3D Euler system}, journal={Trans. Amer. Math. Soc.}, volume={370}, date={2018}, number={4}, pages={2517--2535}, issn={0002-9947}, review={\MR {3748575}}, doi={10.1090/tran/7022}, }
Reference [157]
K. R. Sreenivasan, On the fine-scale intermittency of turbulence, Journal of Fluid Mechanics 151 (1985), 81–103.
Reference [158]
K. R. Sreenivasan, An update on the energy dissipation rate in isotropic turbulence, Phys. Fluids 10 (1998), no. 2, 528–529, DOI 10.1063/1.869575. MR1604680,
Show rawAMSref \bib{Sreenivasan98}{article}{ author={Sreenivasan, Katepalli R.}, title={An update on the energy dissipation rate in isotropic turbulence}, journal={Phys. Fluids}, volume={10}, date={1998}, number={2}, pages={528--529}, issn={1070-6631}, review={\MR {1604680}}, doi={10.1063/1.869575}, }
Reference [159]
K. R. Sreenivasan, S. I. Vainshtein, R. Bhiladvala, I. San Gil, S. Chen, and N. Cao, Asymmetry of Velocity Increments in Fully Developed Turbulence and the Scaling of Low-Order Moments, Phys. Rev. Lett. 77 (1996), no. 8, 1488–1491.
Reference [160]
P.-L. Sulem and U. Frisch, Bounds on energy flux for finite energy turbulence, J. Fluid Mech. 72 (1975), no. 3, 417–423.
Reference [161]
L. Székelyhidi, Weak solutions to the incompressible Euler equations with vortex sheet initial data (English, with English and French summaries), C. R. Math. Acad. Sci. Paris 349 (2011), no. 19-20, 1063–1066, DOI 10.1016/j.crma.2011.09.009. MR2842999,
Show rawAMSref \bib{Szekelyhidi11}{article}{ author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}}, title={Weak solutions to the incompressible Euler equations with vortex sheet initial data}, language={English, with English and French summaries}, journal={C. R. Math. Acad. Sci. Paris}, volume={349}, date={2011}, number={19-20}, pages={1063--1066}, issn={1631-073X}, review={\MR {2842999}}, doi={10.1016/j.crma.2011.09.009}, }
Reference [162]
L. Székelyhidi Jr., From isometric embeddings to turbulence, HCDTE Lecture Notes. Part II. Nonlinear hyperbolic PDEs, dispersive and transport equations, AIMS Ser. Appl. Math., vol. 7, Am. Inst. Math. Sci. (AIMS), Springfield, MO, 2013, pp. 63. MR3340997,
Show rawAMSref \bib{Szekelyhidi12}{article}{ author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={From isometric embeddings to turbulence}, conference={ title={HCDTE Lecture Notes. Part II. Nonlinear hyperbolic PDEs, dispersive and transport equations}, }, book={ series={AIMS Ser. Appl. Math.}, volume={7}, publisher={Am. Inst. Math. Sci. (AIMS), Springfield, MO}, }, date={2013}, pages={63}, review={\MR {3340997}}, }
Reference [163]
L. Székelyhidi Jr., Relaxation of the incompressible porous media equation (English, with English and French summaries), Ann. Sci. Éc. Norm. Supér. (4) 45 (2012), no. 3, 491–509, DOI 10.24033/asens.2171. MR3014484,
Show rawAMSref \bib{Szekelyhidi12b}{article}{ author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={Relaxation of the incompressible porous media equation}, language={English, with English and French summaries}, journal={Ann. Sci. \'{E}c. Norm. Sup\'{e}r. (4)}, volume={45}, date={2012}, number={3}, pages={491--509}, issn={0012-9593}, review={\MR {3014484}}, doi={10.24033/asens.2171}, }
Reference [164]
T. Tao, 255b, Notes 2: Onsager’s conjecture.
Reference [165]
L. Tartar, Compensated compactness and applications to partial differential equations, Nonlinear analysis and mechanics: Heriot-Watt Symposium, Vol. IV, Res. Notes in Math., vol. 39, Pitman, Boston, Mass.-London, 1979, pp. 136–212. MR584398,
Show rawAMSref \bib{tartar_compensated_1979}{article}{ author={Tartar, L.}, title={Compensated compactness and applications to partial differential equations}, conference={ title={Nonlinear analysis and mechanics: Heriot-Watt Symposium, Vol. IV}, }, book={ series={Res. Notes in Math.}, volume={39}, publisher={Pitman, Boston, Mass.-London}, }, date={1979}, pages={136--212}, review={\MR {584398}}, }
Reference [166]
L. Tartar, The compensated compactness method applied to systems of conservation laws, Systems of nonlinear partial differential equations (Oxford, 1982), NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci., vol. 111, Reidel, Dordrecht, 1983, pp. 263–285. MR725524,
Show rawAMSref \bib{tartar_compensated_1983}{article}{ author={Tartar, Luc}, title={The compensated compactness method applied to systems of conservation laws}, conference={ title={Systems of nonlinear partial differential equations}, address={Oxford}, date={1982}, }, book={ series={NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci.}, volume={111}, publisher={Reidel, Dordrecht}, }, date={1983}, pages={263--285}, review={\MR {725524}}, }
Reference [167]
L. Tartar, The compensated compactness method applied to systems of conservation laws, Systems of nonlinear partial differential equations (Oxford, 1982), NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci., vol. 111, Reidel, Dordrecht, 1983, pp. 263–285. MR725524,
Show rawAMSref \bib{Tartar83}{article}{ author={Tartar, Luc}, title={The compensated compactness method applied to systems of conservation laws}, conference={ title={Systems of nonlinear partial differential equations}, address={Oxford}, date={1982}, }, book={ series={NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci.}, volume={111}, publisher={Reidel, Dordrecht}, }, date={1983}, pages={263--285}, review={\MR {725524}}, }
Reference [168]
J. B. Taylor, Relaxation of toroidal plasma and generation of reverse magnetic fields, Physical Review Letters 33 (1974), no. 19, 1139.
Reference [169]
J. B. Taylor, Relaxation and magnetic reconnection in plasmas, Reviews of Modern Physics 58 (1986), no. 3, 741.
Reference [170]
R. Temam, Navier–Stokes equations and nonlinear functional analysis, 2nd ed., CBMS-NSF Regional Conference Series in Applied Mathematics, vol. 66, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 1995, DOI 10.1137/1.9781611970050. MR1318914,
Show rawAMSref \bib{Temam95}{book}{ author={Temam, Roger}, title={Navier--Stokes equations and nonlinear functional analysis}, series={CBMS-NSF Regional Conference Series in Applied Mathematics}, volume={66}, edition={2}, publisher={Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA}, date={1995}, pages={xiv+141}, isbn={0-89871-340-4}, review={\MR {1318914}}, doi={10.1137/1.9781611970050}, }
Reference [171]
T.-P. Tsai, Lectures on Navier–Stokes equations, Graduate Studies in Mathematics, vol. 192, American Mathematical Society, Providence, RI, 2018, DOI 10.1090/gsm/192. MR3822765,
Show rawAMSref \bib{Tsai18}{book}{ author={Tsai, Tai-Peng}, title={Lectures on Navier--Stokes equations}, series={Graduate Studies in Mathematics}, volume={192}, publisher={American Mathematical Society, Providence, RI}, date={2018}, pages={xii+224}, isbn={978-1-4704-3096-2}, review={\MR {3822765}}, doi={10.1090/gsm/192}, }
Reference [172]
J. C. Vassilicos, Dissipation in turbulent flows, Annual review of fluid mechanics. Vol. 47, Annu. Rev. Fluid Mech., vol. 47, Annual Reviews, Palo Alto, CA, 2015, pp. 95–114. MR3727163,
Show rawAMSref \bib{Vassilicos15}{article}{ author={Vassilicos, J. Christos}, title={Dissipation in turbulent flows}, conference={ title={Annual review of fluid mechanics. Vol. 47}, }, book={ series={Annu. Rev. Fluid Mech.}, volume={47}, publisher={Annual Reviews, Palo Alto, CA}, }, date={2015}, pages={95--114}, review={\MR {3727163}}, }
Reference [173]
S. M. Vishik and F. V. Dolzhanskii, Analogs of the euler–lagrange equations and magnetohydrodynamics equations related to lie groups, Sov. Math. Doklady, vol. 19, 1978, pp. 149–153.
Reference [174]
E. Wiedemann, Existence of weak solutions for the incompressible Euler equations, Ann. Inst. H. Poincaré Anal. Non Linéaire 28 (2011), no. 5, 727–730, DOI 10.1016/j.anihpc.2011.05.002. MR2838398,
Show rawAMSref \bib{Wiedemann11}{article}{ author={Wiedemann, Emil}, title={Existence of weak solutions for the incompressible Euler equations}, journal={Ann. Inst. H. Poincar\'{e} Anal. Non Lin\'{e}aire}, volume={28}, date={2011}, number={5}, pages={727--730}, issn={0294-1449}, review={\MR {2838398}}, doi={10.1016/j.anihpc.2011.05.002}, }
Reference [175]
E. Wiedemann, Inviscid symmetry breaking with non-increasing energy (English, with English and French summaries), C. R. Math. Acad. Sci. Paris 351 (2013), no. 23-24, 907–910, DOI 10.1016/j.crma.2013.10.021. MR3133602,
Show rawAMSref \bib{Wiedemann13}{article}{ author={Wiedemann, Emil}, title={Inviscid symmetry breaking with non-increasing energy}, language={English, with English and French summaries}, journal={C. R. Math. Acad. Sci. Paris}, volume={351}, date={2013}, number={23-24}, pages={907--910}, issn={1631-073X}, review={\MR {3133602}}, doi={10.1016/j.crma.2013.10.021}, }
Reference [176]
E. Wiedemann, Weak-strong uniqueness in fluid dynamics, Partial differential equations in fluid mechanics, London Math. Soc. Lecture Note Ser., vol. 452, Cambridge Univ. Press, Cambridge, 2018, pp. 289–326. MR3838055,
Show rawAMSref \bib{Wiedemann17}{article}{ author={Wiedemann, Emil}, title={Weak-strong uniqueness in fluid dynamics}, conference={ title={Partial differential equations in fluid mechanics}, }, book={ series={London Math. Soc. Lecture Note Ser.}, volume={452}, publisher={Cambridge Univ. Press, Cambridge}, }, date={2018}, pages={289--326}, review={\MR {3838055}}, }
Reference [177]
L. Woltjer, On hydromagnetic equilibrium, Proc. Nat. Acad. Sci. U.S.A. 44 (1958), 833–841, DOI 10.1073/pnas.44.9.833. MR96485,
Show rawAMSref \bib{Woltjer58a}{article}{ author={Woltjer, L.}, title={On hydromagnetic equilibrium}, journal={Proc. Nat. Acad. Sci. U.S.A.}, volume={44}, date={1958}, pages={833--841}, issn={0027-8424}, review={\MR {96485}}, doi={10.1073/pnas.44.9.833}, }
Reference [178]
L. Woltjer, A theorem on force-free magnetic fields, Proc. Nat. Acad. Sci. U.S.A. 44 (1958), no. 6, 489.
Reference [179]
K. Yamazaki, Remarks on the non-uniqueness in law of the Navier-Stokes equations up to the J.-L. Lions’ exponent, arXiv:2006.11861 (2020).

Article Information

MSC 2020
Primary: 35Q35 (PDEs in connection with fluid mechanics)
Author Information
Tristan Buckmaster
Department of Mathematics, Princeton University, Princeton, New Jersey
buckmaster@math.princeton.edu
ORCID
MathSciNet
Vlad Vicol
Courant Institute for Mathematical Sciences, New York University, New York, New York
vicol@cims.nyu.edu
ORCID
MathSciNet
Additional Notes

The first author was supported by the NSF grant DMS-1900149 and a Simons Foundation Mathematical and Physical Sciences Collaborative Grant.

The second author was supported by the NSF grant CAREER DMS–1911413.

Journal Information
Bulletin of the American Mathematical Society, Volume 58, Issue 1, ISSN 1088-9485, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2020 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/bull/1713
  • MathSciNet Review: 4188806
  • Show rawAMSref \bib{4188806}{article}{ author={Buckmaster, Tristan}, author={Vicol, Vlad}, title={Convex integration constructions in hydrodynamics}, journal={Bull. Amer. Math. Soc.}, volume={58}, number={1}, date={2021-01}, pages={1-44}, issn={0273-0979}, review={4188806}, doi={10.1090/bull/1713}, }

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.