This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On Moffatt’s magnetic relaxation equations

Rajendra Beekie Courant Institute of Mathematical Sciences, New York University, New York, NY 10012 beekie@cims.nyu.edu Susan Friedlander Department of Mathematics, University of Southern California, Los Angeles, CA 90089 susanfri@usc.edu  and  Vlad Vicol Courant Institute of Mathematical Sciences, New York University, New York, NY 10012 vicol@cims.nyu.edu
Abstract.

We investigate the stability properties for a family of equations introduced by Moffatt to model magnetic relaxation. These models preserve the topology of magnetic streamlines, contain a cubic nonlinearity, and yet have a favorable L2L^{2} energy structure. We consider the local and global in time well-posedness of these models and establish a difference between the behavior as tt\to\infty with respect to weak and strong norms.

1. Introduction

In the 1960s V.I. Arnold developed a new set of geometric ideas concerning the incompressible Euler equations governing the flow of an ideal fluid. In the following decades the subject of topological hydrodynamics flourished. Following Arnold’s seminal work [Arn66] there was an enormous body of literature on the subject. We refer to only a few of the many important papers including Ebin and Marsden [EM70], Holm, Marsden, Ratiu, and Weinstein [HMRW85], and Arnold and Khesin [AK98]. This geometric perspective views the incompressible Euler equations as the geodesic equations of a right-invariant metric on the infinite-dimensional group of volume preserving diffeomorphisms. Of particular importance are the fixed points of the underlying dynamical system, namely steady fluid flows, and their topological richness [EPS12, Gav19, CLV19]. Moreover, as with any dynamical system, of fundamental importance is the question of accessibility of these equilibria. In this paper, we discuss a mechanism of reaching these equilibria not through the Euler vortex dynamics itself, but via a topology preserving diffusion process, called magnetic relaxation.

The magnetic relaxation equation (MRE) considered here was introduced by Moffatt [Mof85, Mof21] to describe a topology-preserving dissipative equation, whose solutions are conjectured to converge in the infinite time limit towards ideal Euler/magnetostatic equilibria (see also Brenier [Bre14]); we recall the motivation in Section 1.1 below. We are interested in understanding the long time behavior for

tB+uB\displaystyle\partial_{t}B+u\cdot\nabla B =Bu\displaystyle=B\cdot\nabla u (1.1a)
(Δ)γu\displaystyle(-\Delta)^{\gamma}u =BB+p\displaystyle=B\cdot\nabla B+\nabla p (1.1b)
divu=divB\displaystyle\mathrm{div\,}u=\mathrm{div\,}B =0\displaystyle=0 (1.1c)

where the unknowns are the incompressible velocity vector field uu, the magnetic vector field BB, and the fluid pressure pp. We consider the problem posed on 𝕋d=[π,π]d\mathbb{T}^{d}=[-\pi,\pi]^{d} with d{2,3}d\in\{2,3\}, and uu is taken to have zero mean on 𝕋d\mathbb{T}^{d}. The parameter γ0\gamma\geq 0 is a regularization parameter of the constitutive law BuB\mapsto u: the case γ=0\gamma=0 corresponds to a Darcy-type regularization (as was done in [Mof85, Bre14, Mof21]), the case γ=1\gamma=1 corresponds to a Stokes-type regularization, while the general case γ>0\gamma>0 may be alternatively used in numerical simulations to smoothen the velocity gradients. This constitutive law may be written as

u=(Δ)γ(BB)=(Δ)γdiv(BB)\displaystyle u=(-\Delta)^{-\gamma}\mathbb{P}(B\cdot\nabla B)=(-\Delta)^{-\gamma}\mathbb{P}\mathrm{div\,}(B\otimes B) (1.2)

where \mathbb{P} is the Leray projector (onto divergence-free vector fields). We emphasize that the topology of the vector field BB is preserved under the vector transport equation (1.1a) irrespective of the regularization parameter γ\gamma in the constitutive law (1.1b). We note that if divB0=0\mathrm{div\,}B_{0}=0, then the vector transport equation (1.1a) preserves the incompressibility of BB at all later times.

From a mathematical perspective, the analysis of the MRE system (1.1) is unusually challenging. Not only is it an active vector equation, versus the more familiar active scalar equations in fluid dynamics [CMT94, CCGO09], but the nonlinearity is cubic in BB. Some of the interesting special features of MRE are discussed in the article of Brenier [Bre14]. Brenier presents a concept of dissipative weak solutions for MRE when the regularization parameter γ\gamma is set to zero. It is shown in two space dimensions that the initial value problem admits such global dissipative weak solutions, and that they are unique whenever they are smooth. However, not even the local existence of strong solutions to (1.1) is known.

Besides local well-posedness, in this paper we examine the long-time behavior of the magnetic relaxation equations (1.1), and show that although the velocity field u(,t)u(\cdot,t) converges to 0 as tt\to\infty (for a sufficiently large regularization parameter γ\gamma), there are many open questions regarding the sense in which the magnetic field B(,t)B(\cdot,t) itself converges as tt\to\infty (weak vs strong convergence; see Remark 4.2). In two dimensions, we give a specific example of asymptotic stability to a simple two dimensional steady state. In contrast, for a specific class of two-and-a-half-dimensional solutions we illustrate instability for the MRE system (1.1), by showing that the the magnetic current ×B\nabla\times B grows unboundedly as tt\to\infty. Our results are presented in Section 1.2 below.

1.1. Motivation behind magnetic relaxation

There are certain well known analogies between Euler equilibria and equilibria of incompressible magnetohydrodynamics (MHD). Recall that the ideal incompressible MHD equations are

tB+uB\displaystyle\partial_{t}B+u\cdot\nabla B =Bu\displaystyle=B\cdot\nabla u (1.3a)
tu+uu+p\displaystyle\partial_{t}u+u\cdot\nabla u+\nabla p =BB\displaystyle=B\cdot\nabla B (1.3b)
divu=divB\displaystyle\mathrm{div\,}u=\mathrm{div\,}B =0.\displaystyle=0\,. (1.3c)

The equilibrium equation for magnetostatics is obtained by setting B=B¯(x)B=\overline{B}(x) and u=0u=0 in (1.3) to give

p¯=B¯B¯,divB¯=0.\displaystyle\nabla\overline{p}=\overline{B}\cdot\nabla\overline{B}\,,\qquad\mathrm{div\,}\overline{B}=0\,. (1.4)

In comparison, the equilibrium equation for incompressible Euler steady states is obtained by setting u=u¯(x)u=\overline{u}(x) and B=0B=0 in (1.3) to give

u¯u¯+p¯=0,divu¯=0.\displaystyle\overline{u}\cdot\nabla\overline{u}+\nabla\overline{p}=0\,,\qquad\mathrm{div\,}\overline{u}=0\,. (1.5)

Clearly, any vector field B¯\overline{B} that satisfies (1.4) is also an equilibrium solving (1.5) upon changing the sign of the pressure, and vice-versa. However, this analogy between magnetic and fluid steady states (1.4)–(1.5) does not extend to the evolution of perturbations about these steady states, as governed by the ideal MHD system on the one hand, respectively the pure Euler dynamics on the other hand. For example, stability issues for Euler steady flows are not the same as the stability for magnetostatic equilibria [HMRW85, Mof86, SV93, FV95].

In [Arn74], Arnold suggested a process which demonstrates the existence of an Euler equilibrium that has the same topological structure as an arbitrary divergence free magnetic field. The idea is to use the evolution dynamics of the magnetic field to reach an Euler/magnetic equilibria which preserves Kelvin circulation. This concept was developed by Moffatt [Mof85] (see also the excellent recent overview [Mof21]). The magnetic relaxation procedure envisioned by Moffatt preserves the streamline topology of an initial divergence free three-dimensional vector B0(x)B_{0}(x), but abandons the constraint that B(x,t)B(x,t) should remain smooth as tt\to\infty. In this model, the magnetic field evolves under the frozen field equation (1.3a) via a vector field u(x,t)u(x,t) which is related to B(x,t)B(x,t) by a suitable constitutive law, which has two properties: that u(x,t)u(x,t) formally decays to 0 as time goes to infinity, and the vector fields uu and j×Bj\times B are parallel with non-negative proportionality factor (here j=×Bj=\nabla\times B is the current field). Moffatt introduced the concept of topological accessibility which is weaker than topological equivalence111Here, we say that B1B_{1} and B0B_{0} are topologically equivalent, if B1(X(α))=αX(α)B0(α)B_{1}(X(\alpha))=\nabla_{\alpha}X(\alpha)B_{0}(\alpha) for a volume preserving diffeomorphism αX(α)\alpha\mapsto X(\alpha). In contrast, to say that B1B_{1} is topologically accessible from B0B_{0} means that (see e.g. in [Mof21, Section 8.2.1]) B1=limtB(,t)B_{1}=\lim_{t\to\infty}B(\cdot,t), where BB is a solution of (1.3a) with initial datum B0B_{0} and some solenoidal vector field uu, under the additional property that 0|𝕋dB(Bu)𝑑x|𝑑t<\int_{0}^{\infty}\left|\int_{\mathbb{T}^{d}}B\cdot(B\cdot\nabla u)dx\right|dt<\infty . because it allows for the appearance of discontinuities in the magnetic field (current sheets) as tt\to\infty. As an example of a constitutive law relating uu to BB, Moffatt [Mof85, Mof21], also Brenier [Bre14], suggested

u=BB+p,\displaystyle u=B\cdot\nabla B+\nabla p\,,

which may be used in conjunction with (1.3a) to show that the magnetic energy satisfies

12ddtBL22=uL22.\displaystyle\frac{1}{2}\frac{d}{dt}\left\|B\right\|_{L^{2}}^{2}=-\left\|u\right\|_{L^{2}}^{2}\,.

Hence the energy of BB is strictly monotonically decreasing, until u0u\equiv 0. Note that B(,t)L2\|B(\cdot,t)\|_{L^{2}} is bounded from below uniformly in time, solely in terms of the initial magnetic helicity [Arn74] (see Remark 2.1).

1.2. Main Results

The main results in this paper are as follows.

In Section 2 we prove local existence for solutions of the MRE system (1.1) in Sobolev spaces HsH^{s}. This result holds for any γ0\gamma\geq 0 and dimension d2d\geq 2, for Sobolev exponents s>d/2+1s>d/2+1. Theorem 2.2 follows from the dissipative nature of (1.1), exhibited in its L2L^{2} energy estimate, by using two commutator estimates at the level of HsH^{s}.

In Section 3 we prove global existence in HsH^{s} when the regularization parameter satisfies γ>d/2+1\gamma>d/2+1. For such γ\gamma, Theorem 3.1 shows that the magnetic relaxation question is well-posed, because we can speak of a global in time solution. We recall that the natural values of γ\gamma coming from physical arguments are γ=0\gamma=0 (corresponding to a Darcy-type approximation) or γ=1\gamma=1 (corresponding to a Stokes-type approximation); unconditional global existence in this range of γ\gamma remains open.

In Section 4 we investigate the possible behavior of the solutions in Section 3 as tt\to\infty. We prove in Theorem 4.1 that the velocity field u(,t)u(\cdot,t) converges to 0, strongly, asymptotically as time diverges. The specific form of this relaxation is given by (4.1). We note, however, we do not obtain a rate for the convergence. Also, it remains open to prove that the vector B(,t)B(\cdot,t) itself converges to a steady Euler (magnetic) weak solution.

In Section 5 we consider the MRE system for d=2d=2 and γ=0\gamma=0. We study the asymptotic stability of a special magnetostatic state, B¯=e1\overline{B}=e_{1}, under Sobolev smooth perturbations. The evolution equation for the perturbations is given by (5.5), which is an active vector equation with a cubic nonlinearity. Equation (5.5) has some similarities with the equation for the perturbation of a linearly stratified density in the two dimensional incompressible porous media equation (IPM); the former being, however, an active scalar equation with a quadratic nonlinearity. In the context of IPM, Elgindi [Elg17] studied the asymptotic stability of the same special steady state and proved that solutions must converge (i.e. relax) as tt\to\infty to a stationary solution of the IPM equation; see also the work [CCL19] in the case of a bounded domain. In Theorem 5.1 we employ some of these ideas to prove asymptotic stability (relaxation) of MRE in this special two dimensional setting.

In Section 6 we turn to the three dimensional MRE system. We observe that there is an interesting class of exact solutions to (1.1) when γ=0\gamma=0, which has analogies to the well know exact solutions of the three dimensional Euler equation, which are in fact two-and-a-half dimensional, cf. Yudovich [Yud74] or DiPerna and Majda [DM87]. In the case of the Euler equation the construction of the exact solution is based on a non-constant coefficient transport equation, which produces a two-and-a-half dimensional flow whose vorticity grows unboundedly in time (linearly in time for shear flows [Yud74, Yud00], or exponentially in time for cellular flows [EM20]). In contrast, for the MRE system the construction of the exact solution is based on a non-constant coefficient heat-type equation, which has a rank 11 diffusion matrix. By choosing the spatial dependence of the initial data appropriately, in Theorem 6.1 we construct an example of a magnetic field B(,t)B(\cdot,t) which converges (relaxes) in L2L^{2} as tt\to\infty to a steady solution B¯\overline{B}, but this limiting solution is not smooth and exhibits magnetic current sheets; as such, the current j(,t)=×B(,t)j(\cdot,t)=\nabla\times B(\cdot,t) grows as t14t^{\frac{1}{4}} in L2L^{2}. Additionally, in Theorem 6.4 we show that in the presence of hyperbolic dynamics, for instance along the separatrix of a cellular flow, the current j(,t)j(\cdot,t) may even grow exponentially in time, for all time, which is a strong type of instability.

Clearly relaxation of the MRE system (1.1) is a very subtle matter. We further illustrate this in Section 7, where we discuss a number of open problems.

Acknowledgements

R.B. was supported by the NSF Graduate Fellowship Grant 1839302. S.F. was in part supported by NSF grant DMS 1613135. S.F. thanks IAS for its hospitality when she was a Member in 2020–21. V.V. was in part supported by the NSF grant CAREER DMS 1911413. V.V. thanks B. Texier and S. Shkoller for stimulating discussions.

2. Local existence in Sobolev spaces for all γ0\gamma\geq 0

The dissipative nature of (1.1), already alluded to in the introduction, is seen by inspecting the magnetic energy estimate

12ddtBL22\displaystyle\frac{1}{2}\frac{d}{dt}\left\|B\right\|_{L^{2}}^{2} =𝕋dB(Bu)=𝕋du(BB)\displaystyle=\int_{\mathbb{T}^{d}}B\cdot\left(B\cdot\nabla u\right)=-\int_{\mathbb{T}^{d}}u\cdot\left(B\cdot\nabla B\right)
=𝕋du((Δ)γup)=uH˙γ2.\displaystyle=-\int_{\mathbb{T}^{d}}u\cdot\left((-\Delta)^{\gamma}u-\nabla p\right)=-\left\|u\right\|_{\dot{H}^{\gamma}}^{2}\,. (2.1)

Integrating in time, we deduce that

sups[0,t]B(,s)L22+20tu(,s)H˙γ2𝑑sB0L22\displaystyle\sup_{s\in[0,t]}\left\|B(\cdot,s)\right\|_{L^{2}}^{2}+2\int_{0}^{t}\left\|u(\cdot,s)\right\|_{\dot{H}^{\gamma}}^{2}ds\leq\left\|B_{0}\right\|_{L^{2}}^{2} (2.2)

for all t>0t>0 such that the solution is sufficiently smooth on [0,t][0,t] to justify the integration by parts manipulations in (2.1).

Remark 2.1 (A global lower bound for the magnetic energy).

We note that no matter the level of regularization in the constitutive law BuB\mapsto u in (1.1b), the magnetic helicity

(t)=𝕋dA(x,t)B(x,t)𝑑x,\displaystyle{\mathcal{H}}(t)=\int_{\mathbb{T}^{d}}A(x,t)\cdot B(x,t)dx\,,

is still a constant function of time,222Note in contrast that the cross-helicity 𝕋duB𝑑x\int_{\mathbb{T}^{d}}u\cdot Bdx is expected to vanish as tt\to\infty since B(,t)B(\cdot,t) remains uniformly bounded in L2L^{2}, while u(,t)0u(\cdot,t)\to 0 in L2L^{2}. as long as the solutions remain sufficiently smooth. Here we have denoted by AA the zero mean vector potential for BB defined in terms of the Biot-Savart law A=(Δ)1×BA=(-\Delta)^{-1}\nabla\times B. Indeed, it is not hard to see that (1.1a) implies that ddt=2𝕋dB(u×B)𝑑x=0\frac{d}{dt}{\mathcal{H}}=2\int_{\mathbb{T}^{d}}B\cdot(u\times B)dx=0. This observation and the Poincaré inequality AL2(𝕋d)BL2(𝕋d)\|A\|_{L^{2}(\mathbb{T}^{d})}\leq\|B\|_{L^{2}(\mathbb{T}^{d})}, imply the so-called Arnold inequality [Arn74]

B(,t)L22|(0)|,\displaystyle\left\|B(\cdot,t)\right\|_{L^{2}}^{2}\geq|{\mathcal{H}}(0)|\,, (2.3)

for all t0t\geq 0. Therefore, while (2.1) shows that the magnetic energy is strictly decreasing as long as u0u\not\equiv 0, (2.3) also shows that the magnetic energy is bounded from below for all time, by a constant that depends only on the magnetic helicity of the initial datum.

Theorem 2.2 (Local existence in Sobolev spaces).

Let γ0\gamma\geq 0 and s>d/2+1s>d/2+1. Assume that B0Hs(𝕋d)B_{0}\in H^{s}(\mathbb{T}^{d}) is divergence free. Then, there exists T(CB0Hs)2T_{*}\geq(C\left\|B_{0}\right\|_{H^{s}})^{-2}, such that the active vector equation (1.1) has a unique solution BC0([0,T);Hs(𝕋d))B\in C^{0}([0,T_{*});H^{s}(\mathbb{T}^{d})), with associated velocity uC0([0,T);Hs1+2γ(𝕋d))L2((0,T);Hs+γ(𝕋d))u\in C^{0}([0,T_{*});H^{s-1+2\gamma}(\mathbb{T}^{d}))\cap L^{2}((0,T_{*});H^{s+\gamma}(\mathbb{T}^{d})). Moreover, BB satisfies the bound (2.2) and also

B(,t)H˙s2B0H˙s2exp(C0tu(,s)L+B(,s)L2ds)\displaystyle\left\|B(\cdot,t)\right\|_{\dot{H}^{s}}^{2}\leq\left\|B_{0}\right\|_{\dot{H}^{s}}^{2}\exp\left(C\int_{0}^{t}\left\|\nabla u(\cdot,s)\right\|_{L^{\infty}}+\left\|\nabla B(\cdot,s)\right\|_{L^{\infty}}^{2}ds\right) (2.4)

for t[0,T)t\in[0,T_{*}), where C>0C>0 is a constant which only depends on ss, γ\gamma, and dd.

Proof of Theorem 2.2.

We use the notation Λ=(Δ)1/2\Lambda=(-\Delta)^{1/2}, and [A,B]=ABBA\bigl{[}A,B\bigr{]}=AB-BA for the commutator of two operators. From (1.1), we then obtain

12ddtBH˙s2+uH˙s+γ2\displaystyle\frac{1}{2}\frac{d}{dt}\left\|B\right\|_{\dot{H}^{s}}^{2}+\left\|u\right\|_{\dot{H}^{s+\gamma}}^{2}
=𝕋dΛsuΛs(BB+p)+𝕋dΛsBΛs(Bu)𝕋dΛsBΛs(uB)\displaystyle=\int_{\mathbb{T}^{d}}\Lambda^{s}u\cdot\Lambda^{s}(B\cdot\nabla B+\nabla p)+\int_{\mathbb{T}^{d}}\Lambda^{s}B\cdot\Lambda^{s}(B\cdot\nabla u)-\int_{\mathbb{T}^{d}}\Lambda^{s}B\cdot\Lambda^{s}(u\cdot\nabla B)
=𝕋dΛsuΛs(BB)+𝕋dΛsB(BΛsu)+𝕋dΛsB[Λs,B]u𝕋dΛsB[Λs,u]B\displaystyle=\int_{\mathbb{T}^{d}}\Lambda^{s}u\cdot\Lambda^{s}(B\cdot\nabla B)+\int_{\mathbb{T}^{d}}\Lambda^{s}B\cdot(B\cdot\nabla\Lambda^{s}u)+\int_{\mathbb{T}^{d}}\Lambda^{s}B\cdot\bigl{[}\Lambda^{s},B\cdot\nabla\bigr{]}u-\int_{\mathbb{T}^{d}}\Lambda^{s}B\cdot\bigl{[}\Lambda^{s},u\cdot\nabla\bigr{]}B
=𝕋dΛsu[Λs,B]B+𝕋dΛsB[Λs,B]u𝕋dΛsB[Λs,u]B.\displaystyle=\int_{\mathbb{T}^{d}}\Lambda^{s}u\cdot\bigl{[}\Lambda^{s},B\cdot\nabla\bigr{]}B+\int_{\mathbb{T}^{d}}\Lambda^{s}B\cdot\bigl{[}\Lambda^{s},B\cdot\nabla\bigr{]}u-\int_{\mathbb{T}^{d}}\Lambda^{s}B\cdot\bigl{[}\Lambda^{s},u\cdot\nabla\bigr{]}B\,. (2.5)

Now, from [Li19, Corollary 5.2, equation (5.1)], by choosing p=p1=p4=2p=p_{1}=p_{4}=2 and p2=p3=p_{2}=p_{3}=\infty, this result states that for all s>0s>0 we have the following generalization of the Kato-Ponce commutator estimate:

[Λs,f]gL2ΛsfL2gL+fLΛs1gL2.\displaystyle\left\|\bigl{[}\Lambda^{s},f\bigr{]}g\right\|_{L^{2}}\lesssim\left\|\Lambda^{s}f\right\|_{L^{2}}\left\|g\right\|_{L^{\infty}}+\left\|\nabla f\right\|_{L^{\infty}}\left\|\Lambda^{s-1}g\right\|_{L^{2}}\,. (2.6)

Applying the estimate (2.6) for the pairs (f,g){(B,B),(B,u),(u,B)}(f,g)\in\{(B,\nabla B),(B,\nabla u),(u,\nabla B)\}, since [,Λs]=0\bigl{[}\nabla,\Lambda^{s}\bigr{]}=0 we obtain that

[Λs,B]BL2\displaystyle\left\|\bigl{[}\Lambda^{s},B\cdot\nabla\bigr{]}B\right\|_{L^{2}} BH˙sBL\displaystyle\lesssim\left\|B\right\|_{\dot{H}^{s}}\left\|\nabla B\right\|_{L^{\infty}} (2.7a)
[Λs,B]uL2+[Λs,u]BL2\displaystyle\left\|\bigl{[}\Lambda^{s},B\cdot\nabla\bigr{]}u\right\|_{L^{2}}+\left\|\bigl{[}\Lambda^{s},u\cdot\nabla\bigr{]}B\right\|_{L^{2}} BH˙suL+BLuH˙s.\displaystyle\lesssim\left\|B\right\|_{\dot{H}^{s}}\left\|\nabla u\right\|_{L^{\infty}}+\left\|\nabla B\right\|_{L^{\infty}}\left\|u\right\|_{\dot{H}^{s}}\,. (2.7b)

By combining (2.5) and (2.7), we arrive at

12ddtBH˙s2+uH˙s+γ2uH˙sBH˙sBL+BH˙s2uL.\displaystyle\frac{1}{2}\frac{d}{dt}\left\|B\right\|_{\dot{H}^{s}}^{2}+\left\|u\right\|_{\dot{H}^{s+\gamma}}^{2}\lesssim\left\|u\right\|_{\dot{H}^{s}}\left\|B\right\|_{\dot{H}^{s}}\left\|\nabla B\right\|_{L^{\infty}}+\left\|B\right\|_{\dot{H}^{s}}^{2}\left\|\nabla u\right\|_{L^{\infty}}\,. (2.8)

Since γ0\gamma\geq 0 and uu has zero mean on 𝕋d\mathbb{T}^{d} we have that uH˙suH˙s+γ\left\|u\right\|_{\dot{H}^{s}}\lesssim\left\|u\right\|_{\dot{H}^{s+\gamma}}, while the condition s>d/2+1s>d/2+1 implies that uLuH˙s\left\|\nabla u\right\|_{L^{\infty}}\lesssim\left\|u\right\|_{\dot{H}^{s}} and BLBH˙s\left\|\nabla B\right\|_{L^{\infty}}\lesssim\left\|B\right\|_{\dot{H}^{s}}. Thus, estimate (2.8) readily implies that there exists a constant C=C(γ,s,d)>0C=C(\gamma,s,d)>0 such that

ddtBH˙s2+uH˙s+γ2CBH˙s4.\displaystyle\frac{d}{dt}\left\|B\right\|_{\dot{H}^{s}}^{2}+\left\|u\right\|_{\dot{H}^{s+\gamma}}^{2}\leq C\left\|B\right\|_{\dot{H}^{s}}^{4}\,. (2.9)

From the a-priori estimates (2.1) and (2.9), the local existence of Ct0HxsC^{0}_{t}H^{s}_{x} solutions of (1.1) readily follows from a standard approximation procedure, and the local time of existence is at least as large as (CB0Hs)2(C\left\|B_{0}\right\|_{H^{s}})^{-2}. Note that since Hs1H^{s-1} is an algebra, we immediately obtain from (1.1b) that uCt0Hxs1+2γu\in C^{0}_{t}H^{s-1+2\gamma}_{x}, while from (2.9) we obtain that uLt2Hxs+γu\in L^{2}_{t}H^{s+\gamma}_{x}. Interestingly, when γ1\gamma\geq 1, the former information (the uniform in time one) provides more regularity in space than the latter one (the integrated in time one). The bound (2.4) is an immediate consequence of (2.8), since s>d/2+1s>d/2+1 and γ0\gamma\geq 0. ∎

3. Global existence for γ>d/2+1\gamma>d/2+1

Theorem 3.1 (Global existence for the strongly regularized system).

Let γ,s>d/2+1\gamma,s>d/2+1. Assume that B0Hs(𝕋d)B_{0}\in H^{s}(\mathbb{T}^{d}) is divergence free. Then, the local in time solution established in Theorem 2.2 is in fact global in time, meaning that T=+T_{*}=+\infty, and we have that

B(,t)H˙s2\displaystyle\left\|B(\cdot,t)\right\|_{\dot{H}^{s}}^{2} B0H˙s2exp(Ct1/2B0L2)\displaystyle\leq\left\|B_{0}\right\|_{\dot{H}^{s}}^{2}\exp\left(Ct^{1/2}\left\|B_{0}\right\|_{L^{2}}\right)
×exp(Ct(B0L2+Ct2B0L6)exp(Ct1/2B0L2))\displaystyle\quad\times\exp\left(Ct\left(\left\|\nabla B_{0}\right\|_{L^{\infty}}^{2}+Ct^{2}\left\|B_{0}\right\|_{L^{\infty}}^{6}\right)\exp\left(Ct^{1/2}\left\|B_{0}\right\|_{L^{2}}\right)\right) (3.1)

for all t0t\geq 0, where C=C(γ,s,d)>0C=C(\gamma,s,d)>0 is a constant.

Proof of Theorem 3.1.

Estimate (2.4) shows that the local in time HsH^{s} solution may be uniquely continued past TT if 0Tu(,s)L+B(,s)L2ds<\int_{0}^{T}\left\|\nabla u(\cdot,s)\right\|_{L^{\infty}}+\left\|\nabla B(\cdot,s)\right\|_{L^{\infty}}^{2}ds<\infty. Thus, the global existence of smooth solutions is established if we show that the Lipschitz norm of uu is integrable in time, and that the Lipschitz norm of BB is square integrable in time.

The condition γ>1+d/2\gamma>1+d/2 implies by the Sobolev embedding that HγLipH^{\gamma}\subset{\rm Lip}. Thus, from (2.2) we deduce

0tu(,s)L𝑑s0tu(,s)H˙γ𝑑st1/2B0L2.\displaystyle\int_{0}^{t}\left\|\nabla u(\cdot,s)\right\|_{L^{\infty}}ds\lesssim\int_{0}^{t}\left\|u(\cdot,s)\right\|_{\dot{H}^{\gamma}}ds\lesssim t^{1/2}\left\|B_{0}\right\|_{L^{2}}\,. (3.2)

Once uu satisfies (3.2), we may use the following classical fact: the solution BB of (1.1a) is given by the vector transport formula

B(X(α,t),t)=αX(α,t)B0(α)\displaystyle B(X(\alpha,t),t)=\nabla_{\alpha}X(\alpha,t)B_{0}(\alpha) (3.3)

where X(α,t)X(\alpha,t) is the 𝕋d\mathbb{T}^{d}-periodic flow of the vector field uu; that is, the solution of the ODEs

ddtX(α,t)=u(X(α,t),t),X(α,0)=α\displaystyle\frac{d}{dt}X(\alpha,t)=u(X(\alpha,t),t),\qquad X(\alpha,0)=\alpha\, (3.4)

and α𝕋d\alpha\in\mathbb{T}^{d} denotes a Lagrangian label. Differentiating (3.4) with respect to α\alpha and appealing to (3.2), we deduce that

X(,t)Lexp(0tu(,s)L𝑑s)\displaystyle\left\|\nabla X(\cdot,t)\right\|_{L^{\infty}}\leq\exp\left(\int_{0}^{t}\left\|\nabla u(\cdot,s)\right\|_{L^{\infty}}ds\right) exp(Ct1/2B0L2).\displaystyle\leq\exp\left(Ct^{1/2}\left\|B_{0}\right\|_{L^{2}}\right)\,. (3.5)

Thus, upon composing (3.3) with the back-to-labels map X1(x,t)X^{-1}(x,t), and appealing to (3.5), we obtain that

B(,t)LB0Lexp(Ct1/2B0L2)\displaystyle\left\|B(\cdot,t)\right\|_{L^{\infty}}\leq\left\|B_{0}\right\|_{L^{\infty}}\exp\left(Ct^{1/2}\left\|B_{0}\right\|_{L^{2}}\right) (3.6)

for all t>0t>0.

It remains to estimate the LL^{\infty} norm of B\nabla B. For this purpose we differentiate (1.1a) with respect to xx, and contract the resulting equation with B\nabla B to deduce

(t+u)|B|24|u||B|2+2|2u||B||B|.\displaystyle(\partial_{t}+u\cdot\nabla)|\nabla B|^{2}\leq 4|\nabla u||\nabla B|^{2}+2|\nabla^{2}u||B||\nabla B|\,.

By the maximum principle, we obtain that

B(,t)L\displaystyle\left\|\nabla B(\cdot,t)\right\|_{L^{\infty}} B0Lexp(20tu(,s)L𝑑s)\displaystyle\leq\left\|\nabla B_{0}\right\|_{L^{\infty}}\exp\left(2\int_{0}^{t}\left\|\nabla u(\cdot,s)\right\|_{L^{\infty}}ds\right)
+0t2u(,s)LB(,s)Lexp(2stu(,s)L𝑑s)𝑑s.\displaystyle\qquad+\int_{0}^{t}\left\|\nabla^{2}u(\cdot,s)\right\|_{L^{\infty}}\left\|B(\cdot,s)\right\|_{L^{\infty}}\exp\left(2\int_{s}^{t}\left\|\nabla u(\cdot,s^{\prime})\right\|_{L^{\infty}}ds^{\prime}\right)ds\,. (3.7)

Thus, we need a bound on the LL^{\infty} norm of the Hessian of uu. For this purpose we note that the condition γ>d/2+1\gamma>d/2+1 trivially implies that γ>3/2\gamma>3/2, and thus the bound

2(Δ)γdivφL\displaystyle\left\|\nabla^{2}(-\Delta)^{-\gamma}\mathbb{P}\mathrm{div\,}\varphi\right\|_{L^{\infty}} φL\displaystyle\lesssim\left\|\varphi\right\|_{L^{\infty}}

holds for every integrable 𝕋d\mathbb{T}^{d}-periodic 22-tensor φ\varphi. Combining the above estimate with the constitutive law (1.2), we obtain

2u(,t)LB(,t)B(,t)LB(,t)L2,\displaystyle\left\|\nabla^{2}u(\cdot,t)\right\|_{L^{\infty}}\lesssim\left\|B(\cdot,t)\otimes B(\cdot,t)\right\|_{L^{\infty}}\lesssim\left\|B(\cdot,t)\right\|_{L^{\infty}}^{2}\,,

for all t>0t>0. From the above display, (3.2), (3.6), and (3.7) we deduce

B(,t)L\displaystyle\left\|\nabla B(\cdot,t)\right\|_{L^{\infty}} (B0L+CtB0L3)exp(Ct1/2B0L2),\displaystyle\leq\left(\left\|\nabla B_{0}\right\|_{L^{\infty}}+Ct\left\|B_{0}\right\|_{L^{\infty}}^{3}\right)\exp\left(Ct^{1/2}\left\|B_{0}\right\|_{L^{2}}\right)\,, (3.8)

for all t>0t>0, where C>0C>0 is a sufficiently large constant which depends on γ\gamma and dd.

The bounds (3.2) and (3.8) conclude the proof of global existence of HsH^{s} solutions. Taking into account the estimate (2.4) we also obtain the bound (3.1). ∎

Remark 3.2.

The condition γ>d/2+1\gamma>d/2+1 implies that for 0<ϵ<2(γd/21)0<\epsilon<2(\gamma-d/2-1) we have the bound

(Δ)γdivφCϵ\displaystyle\left\|\nabla(-\Delta)^{-\gamma}\mathbb{P}\mathrm{div\,}\varphi\right\|_{C^{\epsilon}} φL1\displaystyle\lesssim\left\|\varphi\right\|_{L^{1}} (3.9)

holds for every integrable 𝕋d\mathbb{T}^{d}-periodic 22-tensor φ\varphi. Combining (3.9), (1.2), and (2.2) we first deduce that

u(,t)CϵB(,t)B(,t)L1B(,t)L22B0L22,\displaystyle\left\|\nabla u(\cdot,t)\right\|_{C^{\epsilon}}\lesssim\left\|B(\cdot,t)\otimes B(\cdot,t)\right\|_{L^{1}}\lesssim\left\|B(\cdot,t)\right\|_{L^{2}}^{2}\lesssim\left\|B_{0}\right\|_{L^{2}}^{2}\,, (3.10)

for all t>0t>0. We note that this bound is pointwise in time, in contrast to (3.2) which is time integrated.

4. Convergence as tt\to\infty for γ>d/2+1\gamma>d/2+1

In view of Theorem 3.1, we know that if the initial datum lies in Hs(𝕋d)H^{s}(\mathbb{T}^{d}) and the regularization parameter γ\gamma in (1.2) is sufficiently large, namely γ>d/2+1\gamma>d/2+1, then the system (1.1) has global existence of solutions. In this section we discuss the possible behavior of these solutions as tt\to\infty.

Our first result shows that as tt\to\infty, the velocity field u(,t)u(\cdot,t) converges to 0.

Theorem 4.1 (Asymptotic behavior for the velocity).

Let γ,s>d/2+1\gamma,s>d/2+1 and assume that B0Hs(𝕋d)B_{0}\in H^{s}(\mathbb{T}^{d}) is divergence free. Then the zero mean velocity field uu associated to the magnetic field BC0([0,);Hs(𝕋d))B\in C^{0}([0,\infty);H^{s}(\mathbb{T}^{d})) has the property that

limtu(,t)L=0.\displaystyle\lim_{t\to\infty}\left\|\nabla u(\cdot,t)\right\|_{L^{\infty}}=0\,. (4.1)
Proof of Theorem 4.1.

The proof is based on the bound (3.10), on the energy inequality (2.2), and on a bound for the time derivative of uu, which we claim satisfies

tu(,t)CϵB0L24\displaystyle\left\|\partial_{t}u(\cdot,t)\right\|_{C^{\epsilon}}\lesssim\left\|B_{0}\right\|_{L^{2}}^{4} (4.2)

for all t0t\geq 0. In order to prove (4.2) we apply a time derivative to (1.1b), and replace tB\partial_{t}B in the resulting formula via (1.1a), to arrive at

(Δ)γtuii(tp)\displaystyle(-\Delta)^{\gamma}\partial_{t}u_{i}-\partial_{i}(\partial_{t}p) =t(BjjBi)\displaystyle=\partial_{t}(B_{j}\partial_{j}B_{i})
=jt(BjBi)\displaystyle=\partial_{j}\partial_{t}(B_{j}B_{i})
=j(tBjBi+BjtBi)\displaystyle=\partial_{j}(\partial_{t}B_{j}B_{i}+B_{j}\partial_{t}B_{i})
=j((BkkujukkBj)Bi+Bj(BkkuiukkBi))\displaystyle=\partial_{j}\left((B_{k}\partial_{k}u_{j}-u_{k}\partial_{k}B_{j})B_{i}+B_{j}(B_{k}\partial_{k}u_{i}-u_{k}\partial_{k}B_{i})\right)
=j(BiBkkuj+BjBkkuiukBikBjukBjkBi)\displaystyle=\partial_{j}\left(B_{i}B_{k}\partial_{k}u_{j}+B_{j}B_{k}\partial_{k}u_{i}-u_{k}B_{i}\partial_{k}B_{j}-u_{k}B_{j}\partial_{k}B_{i}\right)
=j(BiBkkuj+BjBkkui)j(ukk(BiBj))\displaystyle=\partial_{j}\left(B_{i}B_{k}\partial_{k}u_{j}+B_{j}B_{k}\partial_{k}u_{i}\right)-\partial_{j}\left(u_{k}\partial_{k}(B_{i}B_{j})\right)
=j(BiBkkuj+BjBkkui)jk(ukBiBj)\displaystyle=\partial_{j}\left(B_{i}B_{k}\partial_{k}u_{j}+B_{j}B_{k}\partial_{k}u_{i}\right)-\partial_{j}\partial_{k}\left(u_{k}B_{i}B_{j}\right)

for every component i{1,,d}i\in\{1,\ldots,d\}. Therefore, we have established that

tu\displaystyle\partial_{t}u =(Δ)γdiv(B(B)u+(B)uB)(Δ)γdivdiv(BBu).\displaystyle=(-\Delta)^{-\gamma}\mathbb{P}\mathrm{div\,}\left(B\otimes(B\cdot\nabla)u+(B\cdot\nabla)u\otimes B\right)-(-\Delta)^{-\gamma}\mathbb{P}\mathrm{div\,}\mathrm{div\,}\left(B\otimes B\otimes u\right)\,.

Since γ>1+d/2\gamma>1+d/2, we may again use inequality (3.9) along with the Poincaré inequality, and deduce from the above formula for tu\partial_{t}u that

tu(,t)Cϵ\displaystyle\left\|\partial_{t}u(\cdot,t)\right\|_{C^{\epsilon}} (BBu)(,t)L1+(BBu)(,t)L1\displaystyle\lesssim\left\|(B\otimes B\otimes\nabla u)(\cdot,t)\right\|_{L^{1}}+\left\|(B\otimes B\otimes u)(\cdot,t)\right\|_{L^{1}}
B(,t)L2(𝕋d)2u(,t)W1,\displaystyle\lesssim\left\|B(\cdot,t)\right\|_{L^{2}(\mathbb{T}^{d})}^{2}\left\|u(\cdot,t)\right\|_{W^{1,\infty}}
B0L2(𝕋d)4.\displaystyle\lesssim\left\|B_{0}\right\|_{L^{2}(\mathbb{T}^{d})}^{4}\,.

In the last inequality above we have appealed to (2.2) and (3.10). Thus, we have shown that (4.2) holds.

In order to conclude the proof, we note that the energy inequality (2.2) and the Sobolev embedding HγLH^{\gamma}\subset L^{\infty} gives that

0u(,t)L2\displaystyle\int_{0}^{\infty}\left\|u(\cdot,t)\right\|_{L^{\infty}}^{2} B0L2(𝕋d)2.\displaystyle\lesssim\left\|B_{0}\right\|_{L^{2}(\mathbb{T}^{d})}^{2}\,.

Combined with (4.2), the above estimate shows that

limtu(,t)L=0.\displaystyle\lim_{t\to\infty}\left\|u(\cdot,t)\right\|_{L^{\infty}}=0\,.

The conclusion (4.1) now follows by interpolating the LL^{\infty} norm of u\nabla u between the LL^{\infty} norm of uu, which vanishes as tt\to\infty as shown above, and the CϵC^{\epsilon} norm of u\nabla u, which is uniformly bounded by (3.10). ∎

Remark 4.2 (Relaxation towards Euler steady states?).

Since (4.1) shows that u(,t)L\left\|\nabla u(\cdot,t)\right\|_{L^{\infty}} vanishes as tt\to\infty, it is tempting to conjecture (as was already done by Moffatt [Mof85]) that as tt\to\infty the magnetic field B(,t)B(\cdot,t) relaxes to a steady state B¯\overline{B} which solves (1.1b) with the left hand side equals to zero; that is, B¯\overline{B} is a stationary solution of the incompressible Euler equations. The purpose of this remark is to argue that the information provided in Theorem 4.1 does not appear to be sufficient to conclude this statement. By (2.2) and weak compactness we do have the existence of subsequences tkt_{k}\to\infty such that the associated magnetic fields Bk(x)=B(x,tk)B_{k}(x)=B(x,t_{k}) converge weakly in L2L^{2}; say BkB¯B_{k}\rightharpoonup\overline{B}, for some (weakly) incompressible vector field B¯\overline{B}. Additionally, (2.1) shows that the sequence {BkL2}k1\{\|B_{k}\|_{L^{2}}\}_{k\geq 1} is strictly decreasing and non-negative, so that there exists E¯0\overline{E}\geq 0 with limkBkL2=E¯\lim_{k\to\infty}\|B_{k}\|_{L^{2}}=\overline{E}; in fact E¯>0\overline{E}>0 in view of the Arnold inequality (2.3) as long as the initial datum is topologically nontrivial. Of course, we do not know whether B¯L2\|\overline{B}\|_{L^{2}} equals to E¯\overline{E}, or else we’d have strong L2L^{2} convergence as kk\to\infty. The additional information provided by Theorem 4.1 and (1.1b) gives that div(BkBk)0\mathbb{P}\mathrm{div\,}(B_{k}\otimes B_{k})\rightharpoonup 0. These facts do not, however, seem to imply that div(B¯B¯)=0\mathbb{P}\mathrm{div\,}(\overline{B}\otimes\overline{B})=0 in the sense of distributions, which would be the condition that B¯\overline{B} is a stationary Euler flow. Indeed, we may only deduce that

div(B¯B¯)\displaystyle\mathbb{P}\mathrm{div\,}(\overline{B}\otimes\overline{B}) =div((B¯Bk)B¯)+div(B¯(B¯Bk))\displaystyle=\mathbb{P}\mathrm{div\,}((\overline{B}-B_{k})\otimes\overline{B})+\mathbb{P}\mathrm{div\,}(\overline{B}\otimes(\overline{B}-B_{k}))
+div(BkBk)div((BkB¯)(BkB¯)).\displaystyle\qquad+\mathbb{P}\mathrm{div\,}(B_{k}\otimes B_{k})-\mathbb{P}\mathrm{div\,}((B_{k}-\overline{B})\otimes(B_{k}-\overline{B}))\,.

The first three terms on the right side of the above do converge to 0 weakly as kk\to\infty (the first two terms by the assumption that BkB¯B_{k}\rightharpoonup\overline{B}, and the third term due to Theorem 4.1). However, we do not have enough information to conclude that the fourth term converges to 0 as kk\to\infty; the information that (BkB¯)(BkB¯)(B_{k}-\overline{B})\otimes(B_{k}-\overline{B}) is a symmetric non-negative tensor, which is uniformly bounded in L1L^{1}, is not sufficient; the enemy is B¯L2<E¯\|\overline{B}\|_{L^{2}}<\overline{E}.

Remark 4.3 (Time integrability of u(,t)L\left\|\nabla u(\cdot,t)\right\|_{L^{\infty}}).

From (2.2) and the Sobolev embedding HγW1,H^{\gamma}\subset W^{1,\infty} we may deduce that u(,t)LL2(0,)\left\|\nabla u(\cdot,t)\right\|_{L^{\infty}}\in L^{2}(0,\infty). It remains however an open problem to show that the Lipschitz norm of uu decays sufficiently fast to ensure that u(,t)LL1(0,)\left\|\nabla u(\cdot,t)\right\|_{L^{\infty}}\in L^{1}(0,\infty). If this faster decay were true, then in view of the H˙1\dot{H}^{1} energy estimate for (1.1), which after exploring a few cancelations can be shown to be

12ddtBL22+uH˙γ+123BL22uL,\displaystyle\frac{1}{2}\frac{d}{dt}\left\|\nabla B\right\|_{L^{2}}^{2}+\left\|u\right\|_{\dot{H}^{\gamma+1}}^{2}\leq 3\left\|\nabla B\right\|_{L^{2}}^{2}\left\|\nabla u\right\|_{L^{\infty}}\,,

would imply that B(,t)L2\left\|\nabla B(\cdot,t)\right\|_{L^{2}} is uniformly bounded in time. In turn, such information would be sufficient to extract as tt\to\infty limit points B¯\overline{B}, which are stationary solutions of the Euler equations. We note, however, that at least when γ=0\gamma=0, in Theorem 6.1 we show, for suitable choices of initial data, the H˙1\dot{H}^{1} norm of BB does not remain uniformly bounded in time. Thus it is not possible for the Lipschitz norm of uu to be integrable in time. Whether this situation is generic remains an open problem.

5. 2D stability of the state B=e1B=e_{1} and u=0u=0

We consider the MRE system (1.1) in two space dimensions with γ=0\gamma=0. In this section we study the asymptotic stability of the steady state

B=e1andu=0,B=e_{1}\qquad\mbox{and}\qquad u=0\,,

under Sobolev smooth perturbations. We note that for the MHD system with viscosity but no resistivity on 2\mathbb{R}^{2}, i.e. for (1.3) with Δu\Delta u added to (1.3b), the stability of (u,B)=(0,e1)(u,B)=(0,e_{1}) was proved for Sobolev smooth perturbations with certain admissibility conditions for the initial data of magnetic perturbations in [LXZ15] (see also [RWXZ14] where the admissibility conditions were removed). These works make use of the fact that at the linearized level uu satisfies

t2uΔtu12u=0.\partial_{t}^{2}u-\Delta\partial_{t}u-\partial_{1}^{2}u=0\,.

For the magnetic relaxation equations (1.1), this favorable structure is no longer available since uu is completely determined from BB through (1.1b). Linearizing around (u,B)=(0,e1)(u,B)=(0,e_{1}) instead leads to the partially dissipative equation

tB=12B.\partial_{t}B=\partial_{1}^{2}B\,. (5.1)

This motivates the use of a different approach to proving global existence for the perturbations, as in [Elg17].

The perturbation of the magnetic field around the steady state e1e_{1} is written as bb, i.e. we consider

b:=Be1.\displaystyle b:=B-e_{1}\,. (5.2)

From (1.1b), with γ=0\gamma=0, we deduce that

u=1b+v\displaystyle u=\partial_{1}b+v (5.3)

where the nonlinear part of the velocity, denoted as vv, is given by

v:=bb+p,divv=0.\displaystyle v:=b\cdot\nabla b+\nabla p\,,\qquad\mathrm{div\,}v=0\,. (5.4)

Inserting the ansatz (5.2)–(5.4) into (1.1) we obtain the evolution equation for the perturbation of the magnetic field

tb+vbbv12b=b1b1bb+1v.\displaystyle\partial_{t}b+v\cdot\nabla b-b\cdot\nabla v-\partial_{1}^{2}b=b\cdot\nabla\partial_{1}b-\partial_{1}b\cdot\nabla b+\partial_{1}v\,.

Using (5.4) we arrive at the following system for the magnetic perturbation:

tb+vbbv12b\displaystyle\partial_{t}b+v\cdot\nabla b-b\cdot\nabla v-\partial_{1}^{2}b =2(b1b)\displaystyle=2\mathbb{P}(b\cdot\nabla\partial_{1}b) (5.5a)
v\displaystyle v =bb+p\displaystyle=b\cdot\nabla b+\nabla p (5.5b)
divv\displaystyle\mathrm{div\,}v =divb=0.\displaystyle=\mathrm{div\,}b=0\,. (5.5c)

Before stating our main theorem, it will be useful to introduce notation for the x1x_{1}-independent and the x1x_{1}-dependent components of bb. As such, for any function ψ:𝕋2\psi\colon\mathbb{T}^{2}\to\mathbb{R}, we define

0ψ(x2)\displaystyle\mathbb{P}_{0}\psi(x_{2}) :=𝕋ψ(x1,x2)𝑑x1\displaystyle:=\fint_{\mathbb{T}}\psi(x_{1},x_{2})dx_{1}
ψ(x1,x2)\displaystyle\mathbb{P}_{\!\perp}\psi(x_{1},x_{2}) :=ψ(x1,x2)0ψ(x2).\displaystyle:=\psi(x_{1},x_{2})-\mathbb{P}_{0}\psi(x_{2})\,.

With this notation, our main result concerning the system (5.5) is:

Theorem 5.1 (Stability and relaxation).

Let k4k\geq 4 and mk+9m\geq k+9. Choose δ(0,1)\delta\in(0,1). There exists ε0\varepsilon_{0} such that if

b0Hm=εε0,\displaystyle\|b_{0}\|_{H^{m}}=\varepsilon\leq\varepsilon_{0}\,, (5.6)

and 0b0=0\mathbb{P}_{0}b_{0}=0, then we have that (5.5) has a unique global in time smooth solution (b,v)(b,v), which satisfies b(,t)L2ε\left\|b(\cdot,t)\right\|_{L^{2}}\leq\varepsilon and

b(,t)H˙k4εe(1δ)t\displaystyle\left\|\mathbb{P}_{\!\perp}b(\cdot,t)\right\|_{\dot{H}^{k}}\leq 4\varepsilon e^{-(1-\delta)t} (5.7a)
0b1(,t)Hk+24ε\displaystyle\left\|\mathbb{P}_{0}b_{1}(\cdot,t)\right\|_{H^{k+2}}\leq 4\varepsilon (5.7b)
b(,t)H˙m24εeεt\displaystyle\left\|b(\cdot,t)\right\|_{\dot{H}^{m}}^{2}\leq 4\varepsilon e^{\varepsilon t} (5.7c)

for t[0,)t\in[0,\infty). As a consequence, the total velocity field satisfies u(,t)0u(\cdot,t)\to 0 as tt\to\infty, whereas the total magnetic field B(,t)=e1+b(,t)B(\cdot,t)=e_{1}+b(\cdot,t) relaxes to a steady state B¯\overline{B} with B¯e1Hk+24ε\|\overline{B}-e_{1}\|_{H^{k+2}}\leq 4\varepsilon, both convergences taking place with respect to strong topologies.

Remark 5.2 (Notation).

For simplicity of notation, throughout the proof of Theorem 5.1, we shall use the notation:

a\displaystyle a =a(x2,t)=0b1(x2,t)\displaystyle=a(x_{2},t)=\mathbb{P}_{0}b_{1}(x_{2},t) (5.8a)
f\displaystyle f =f(x1,x2,t)=b(x1,x2,t)\displaystyle=f(x_{1},x_{2},t)=\mathbb{P}_{\!\perp}b(x_{1},x_{2},t) (5.8b)
w\displaystyle w =w(x1,x2,t)=v(x1,x2,t).\displaystyle=w(x_{1},x_{2},t)=\mathbb{P}_{\!\perp}v(x_{1},x_{2},t)\,. (5.8c)

We do not introduce new notation for 0v1\mathbb{P}_{0}v_{1}, but note from (5.5b) and the observation that 0b2(,t)=0\mathbb{P}_{0}b_{2}(\cdot,t)=0 (which will be established in Lemma 5.3 below), we have

0v1=20(b2b1)=20(f2(f1+a))=20(f2f1).\displaystyle\mathbb{P}_{0}v_{1}=\partial_{2}\mathbb{P}_{0}(b_{2}b_{1})=\partial_{2}\mathbb{P}_{0}(f_{2}(f_{1}+a))=\partial_{2}\mathbb{P}_{0}(f_{2}f_{1})\,.

The above identity will be used in the analysis below. Note that with the notation in (5.8), we have that the stability estimates in (5.7) become

f(,t)H˙k4εe(1δ)t,a(,t)Hk+24ε,b(,t)H˙m4εeεt,\displaystyle\left\|f(\cdot,t)\right\|_{\dot{H}^{k}}\leq 4\varepsilon e^{-(1-\delta)t}\,,\qquad\left\|a(\cdot,t)\right\|_{H^{k+2}}\leq 4\varepsilon\,,\qquad\left\|b(\cdot,t)\right\|_{\dot{H}^{m}}\leq 4\varepsilon e^{\varepsilon t}\,, (5.9)

which are the bounds proven below.

5.1. The evolution equations for aa and ff

Before turning to the proof of Theorem 5.1, we need to determine the evolution equations for aa and ff. In turn, this is necessary because the 12b\partial_{1}^{2}b dissipative term present in (5.5a) may only be expected to cause decay of the part of bb which is not constant in the x1x_{1} direction, i.e. for ff. Moreover, the precise coupling between the evolution equations for aa and ff is crucial to the proof (and is also the reason why in three dimensions this stability result doesn’t hold). In this direction, for aa we have:

Lemma 5.3 (The aa evolution).

Assume that (b,v)(b,v) are smooth solutions of (5.5) and 0b0=0\mathbb{P}_{0}b_{0}=0. Then, we have

ta\displaystyle\partial_{t}a =20(2f21f1+f2w1w2f1)=:N(f,w)\displaystyle=\partial_{2}\mathbb{P}_{0}(2f_{2}\partial_{1}f_{1}+f_{2}w_{1}-w_{2}f_{1})=:N^{\prime}(f,w) (5.10)

and 0b2(,t)=0\mathbb{P}_{0}b_{2}(\cdot,t)=0. Crucially, aa does not appear on the right side of (5.10).

Proof of Lemma 5.3.

Applying 0\mathbb{P}_{0} to (5.5a) and using that 01ψ=0\mathbb{P}_{0}\partial_{1}\psi=0 for any periodic ψ\psi, gives

t0bi=20(b2viv2bi+2b21bi).\partial_{t}\mathbb{P}_{0}b_{i}=\partial_{2}\mathbb{P}_{0}(b_{2}v_{i}-v_{2}b_{i}+2b_{2}\partial_{1}b_{i}). (5.11)

When i=2i=2, we appeal to the fact that 0(b21b2)=01(b22/2)=0\mathbb{P}_{0}(b_{2}\partial_{1}b_{2})=\mathbb{P}_{0}\partial_{1}(b_{2}^{2}/2)=0, and to the assumption 0b2,0=0\mathbb{P}_{0}b_{2,0}=0, to conclude from (5.11) that 0b2(x2,t)=0\mathbb{P}_{0}b_{2}(x_{2},t)=0 for all t0t\geq 0. In particular, this implies that f2=b2f_{2}=b_{2}. Moreover, since 𝕋2f=0\fint_{\mathbb{T}^{2}}f=0 and 0=divb=divf0=\mathrm{div\,}b=\mathrm{div\,}f, there exists a periodic stream function ϕ\phi, such that f=ϕf=\nabla^{\perp}\phi. In particular,

f2=1ϕ.f_{2}=\partial_{1}\phi\,.

Similarly, since divv=0\mathrm{div\,}v=0 and 𝕋2v=𝕋2div(bb+pI2)=0\fint_{\mathbb{T}^{2}}v=\fint_{\mathbb{T}^{2}}\mathrm{div\,}(b\otimes b+pI_{2})=0, we obtain that there exists a periodic stream function φ\varphi, such that v=φv=\nabla^{\perp}\varphi. In particular,

v2=w2=1φ.v_{2}=w_{2}=\partial_{1}\varphi\,.

With this information, we return to (5.11) and set i=1i=1. Since 1b1=1f1\partial_{1}b_{1}=\partial_{1}f_{1}, b2=f2b_{2}=f_{2}, and v2=w2v_{2}=w_{2}, we have that

0(b2v1v2b1+2b21b1)\displaystyle\mathbb{P}_{0}(b_{2}v_{1}-v_{2}b_{1}+2b_{2}\partial_{1}b_{1}) =0(f2v1w2b1+2f21f1)\displaystyle=\mathbb{P}_{0}(f_{2}v_{1}-w_{2}b_{1}+2f_{2}\partial_{1}f_{1})
=0(f2w1w2f1+2f21f1)+0(b20v1v20b1)\displaystyle=\mathbb{P}_{0}(f_{2}w_{1}-w_{2}f_{1}+2f_{2}\partial_{1}f_{1})+\mathbb{P}_{0}(\mathbb{P}_{\!\perp}b_{2}\mathbb{P}_{0}v_{1}-\mathbb{P}_{\!\perp}v_{2}\mathbb{P}_{0}b_{1})

which establishes (5.10), upon noting that 0(ψ10ψ2)=0ψ20(ψ)=0\mathbb{P}_{0}(\mathbb{P}_{\!\perp}\psi_{1}\mathbb{P}_{0}\psi_{2})=\mathbb{P}_{0}\psi_{2}\mathbb{P}_{0}(\mathbb{P}_{\!\perp}\psi)=0 for any ψ1,ψ2\psi_{1},\psi_{2}. ∎

The evolution equation for ff is more complicated, and is given by the following lemma.

Lemma 5.4 (The ff evolution).

Assume that (b,v)(b,v) are smooth solutions of (5.5) and 0b0=0\mathbb{P}_{0}b_{0}=0. Then, we have

tf=L(f)+N(f,w)\partial_{t}f=L(f)+N(f,w) (5.12)

where the linear operator LL acts on the vector field f=(f1,f2)Tf=(f_{1},f_{2})^{T} as

L(f)\displaystyle L(f) :=(1+a)212f+(1+a)1pL2a2pLe1\displaystyle:=(1+a)^{2}\partial_{1}^{2}f+(1+a)\nabla\partial_{1}p_{L}-\partial_{2}a\partial_{2}p_{L}e_{1} (5.13a)
pL\displaystyle p_{L} :=pL(a,f)=2(Δ)1(2a1f2)\displaystyle:=p_{L}(a,f)=2(-\Delta)^{-1}(\partial_{2}a\partial_{1}f_{2}) (5.13b)

where as in (5.8), a=0b1a=\mathbb{P}_{0}b_{1}. The nonlinear operator NN appearing in (5.12) is defined as

N(f,w)\displaystyle N(f,w) :=a1(ff+pN)+(fwwf)20(f1f2)1f\displaystyle:=a\partial_{1}\mathbb{P}_{\!\perp}(f\cdot\nabla f+\nabla p_{N})+\mathbb{P}_{\!\perp}(f\cdot\nabla w-w\cdot\nabla f)-\partial_{2}\mathbb{P}_{0}(f_{1}f_{2})\partial_{1}f
+2(f1f)+1pN+(220(f1f2)f22a(ff2+2pN))e1\displaystyle\quad+2\mathbb{P}_{\!\perp}(f\cdot\nabla\partial_{1}f)+\nabla\partial_{1}\mathbb{P}_{\!\perp}p_{N}+\left(\partial_{2}^{2}\mathbb{P}_{0}(f_{1}f_{2})f_{2}-\partial_{2}a\mathbb{P}_{\!\perp}(f\cdot\nabla f_{2}+\partial_{2}p_{N})\right)e_{1} (5.14a)
pN\displaystyle p_{N} :=pN(f)=2(Δ)1((1f1)2+1f22f1)).\displaystyle:=p_{N}(f)=2(-\Delta)^{-1}((\partial_{1}f_{1})^{2}+\partial_{1}f_{2}\partial_{2}f_{1})). (5.14b)
Proof of Lemma 5.4.

We apply \mathbb{P}_{\!\perp} to (5.5a) to get

tf12f=(bvvb+2b1b+1p).\partial_{t}f-\partial_{1}^{2}f=\mathbb{P}_{\!\perp}(b\cdot\nabla v-v\cdot\nabla b+2b\cdot\nabla\partial_{1}b+\nabla\partial_{1}p)\,. (5.15)

The goal is to further decompose the right side of (5.15), in order to extract from it all local and nonlocal terms which are linear in ff.

We first determine a decomposition for the pressure. Applying a divergence to (5.4) gives

Δp=div(bb)\displaystyle-\Delta p=\mathrm{div\,}(b\cdot\nabla b) =2(1b1)2+21b22b1\displaystyle=2(\partial_{1}b_{1})^{2}+2\partial_{1}b_{2}\partial_{2}b_{1}
=2((1f1)2+1f22f1)=:ΔpN+21f22a=:ΔpL\displaystyle=\underbrace{2((\partial_{1}f_{1})^{2}+\partial_{1}f_{2}\partial_{2}f_{1})}_{=:-\Delta p_{N}}+\underbrace{2\partial_{1}f_{2}\partial_{2}a}_{=:-\Delta p_{L}} (5.16)

where pNp_{N} is the pressure which is nonlinear in ff, and pLp_{L} is the pressure which is linear with respect to ff. Note that both of these pressure terms are uniquely defined once we impose that they have zero mean on 𝕋2\mathbb{T}^{2}, and that they correspond to definitions (5.13b) and (5.14b).

Next, we compute the velocity in terms of the magnetic perturbation. As noted in Remark 5.8, we may decompose the velocity field as

v1=w1+20(f1f2),v2=w2.\displaystyle v_{1}=w_{1}+\partial_{2}\mathbb{P}_{0}(f_{1}f_{2})\,,\qquad v_{2}=w_{2}\,. (5.17)

Furthermore, by applying \mathbb{P}_{\!\perp} to (5.4), and using (5.1), we obtain that

w\displaystyle w =(b11b+b22b+p)\displaystyle=\mathbb{P}_{\!\perp}(b_{1}\partial_{1}b+b_{2}\partial_{2}b+\nabla p)
=a1f+pL+2af2e1+(ff+pN).\displaystyle=a\partial_{1}f+\nabla p_{L}+\partial_{2}af_{2}e_{1}+\mathbb{P}_{\!\perp}(f\cdot\nabla f+\nabla p_{N})\,. (5.18)

With (5.1) in hand, we now compute the stretching and the advection terms present on the right side of (5.15). Indeed, from (5.17) and (5.1) for the stretching term in (5.15) we obtain

(bv1)\displaystyle\mathbb{P}_{\!\perp}(b\cdot\nabla v_{1}) =((a+f1)1w1+f22(w1+20(f1f2)))\displaystyle=\mathbb{P}_{\!\perp}\left((a+f_{1})\partial_{1}w_{1}+f_{2}\partial_{2}(w_{1}+\partial_{2}\mathbb{P}_{0}(f_{1}f_{2}))\right)
=a1w1+f2220(f1f2)+(fw1)\displaystyle=a\partial_{1}w_{1}+f_{2}\partial_{2}^{2}\mathbb{P}_{0}(f_{1}f_{2})+\mathbb{P}_{\!\perp}(f\cdot\nabla w_{1})
=a1(a1f1+2af2+1pL)+a1(ff1+1pN)+f2220(f1f2)+(fw1)\displaystyle=a\partial_{1}(a\partial_{1}f_{1}+\partial_{2}af_{2}+\partial_{1}p_{L})+a\partial_{1}\mathbb{P}_{\!\perp}(f\cdot\nabla f_{1}+\partial_{1}p_{N})+f_{2}\partial_{2}^{2}\mathbb{P}_{0}(f_{1}f_{2})+\mathbb{P}_{\!\perp}(f\cdot\nabla w_{1})
=a212f1+a2a1f2+a12pL+a1(ff1+1pN)+f2220(f1f2)+(fw1)\displaystyle=a^{2}\partial_{1}^{2}f_{1}+a\partial_{2}a\partial_{1}f_{2}+a\partial_{1}^{2}p_{L}+a\partial_{1}\mathbb{P}_{\!\perp}(f\cdot\nabla f_{1}+\partial_{1}p_{N})+f_{2}\partial_{2}^{2}\mathbb{P}_{0}(f_{1}f_{2})+\mathbb{P}_{\!\perp}(f\cdot\nabla w_{1})

and similarly,

(bv2)\displaystyle\mathbb{P}_{\!\perp}(b\cdot\nabla v_{2}) =((a+f1)1w2+f22w2)\displaystyle=\mathbb{P}_{\!\perp}((a+f_{1})\partial_{1}w_{2}+f_{2}\partial_{2}w_{2})
=a1w2+(fw2)\displaystyle=a\partial_{1}w_{2}+\mathbb{P}_{\!\perp}(f\cdot\nabla w_{2})
=a1(a1f2+2pL)+a1(ff2+2pN)+(fw2)\displaystyle=a\partial_{1}(a\partial_{1}f_{2}+\partial_{2}p_{L})+a\partial_{1}\mathbb{P}_{\!\perp}(f\cdot\nabla f_{2}+\partial_{2}p_{N})+\mathbb{P}_{\!\perp}(f\cdot\nabla w_{2})
=a212f2+a12pL+a1(ff2+2pN)+(fw2).\displaystyle=a^{2}\partial_{1}^{2}f_{2}+a\partial_{1}\partial_{2}p_{L}+a\partial_{1}\mathbb{P}_{\!\perp}(f\cdot\nabla f_{2}+\partial_{2}p_{N})+\mathbb{P}_{\!\perp}(f\cdot\nabla w_{2})\,.

On the other hand, for the transport term in (5.15) we have

(vb1)\displaystyle\mathbb{P}_{\!\perp}(v\cdot\nabla b_{1}) =((w1+20(f1f2))1f1+w22(f1+a))\displaystyle=\mathbb{P}_{\!\perp}((w_{1}+\partial_{2}\mathbb{P}_{0}(f_{1}f_{2}))\partial_{1}f_{1}+w_{2}\partial_{2}(f_{1}+a))
=(wf1)+20(f1f2)1f1+2aw2\displaystyle=\mathbb{P}_{\!\perp}(w\cdot\nabla f_{1})+\partial_{2}\mathbb{P}_{0}(f_{1}f_{2})\partial_{1}f_{1}+\partial_{2}aw_{2}
=a2a1f2+2a2pL+2a(ff2+2pN)+20(f1f2)1f1+(wf1),\displaystyle=a\partial_{2}a\partial_{1}f_{2}+\partial_{2}a\partial_{2}p_{L}+\partial_{2}a\mathbb{P}_{\!\perp}(f\cdot\nabla f_{2}+\partial_{2}p_{N})+\partial_{2}\mathbb{P}_{0}(f_{1}f_{2})\partial_{1}f_{1}+\mathbb{P}_{\!\perp}(w\cdot\nabla f_{1})\,,

and

(vb2)\displaystyle\mathbb{P}_{\!\perp}(v\cdot\nabla b_{2}) =((w1+20(f1f2))1f2+w22b2)\displaystyle=\mathbb{P}_{\!\perp}((w_{1}+\partial_{2}\mathbb{P}_{0}(f_{1}f_{2}))\partial_{1}f_{2}+w_{2}\partial_{2}b_{2})
=20(f1f2)1f2+(wf2).\displaystyle=\partial_{2}\mathbb{P}_{0}(f_{1}f_{2})\partial_{1}f_{2}+\mathbb{P}_{\!\perp}(w\cdot\nabla f_{2})\,.

For the third nonlinear term on the right side of (5.15) we have

(b1b)\displaystyle\mathbb{P}_{\!\perp}(b\cdot\nabla\partial_{1}b) =((a+f1)12f+f221f)\displaystyle=\mathbb{P}_{\!\perp}((a+f_{1})\partial_{1}^{2}f+f_{2}\partial_{2}\partial_{1}f)
=a12f+(f1f).\displaystyle=a\partial_{1}^{2}f+\mathbb{P}_{\!\perp}(f\cdot\nabla\partial_{1}f)\,.

Gathering the above five displayed equations, we obtain that

Linear terms on right side of (5.15)\displaystyle\mbox{Linear terms on right side of }\eqref{eq:pt:f:1} =a212f+a1pL2a2pLe1+2a12f,\displaystyle=a^{2}\partial_{1}^{2}f+a\partial_{1}\nabla p_{L}-\partial_{2}a\partial_{2}p_{L}e_{1}+2a\partial_{1}^{2}f\,,
Nonlinear terms on right side of (5.15)\displaystyle\mbox{Nonlinear terms on right side of }\eqref{eq:pt:f:1} =a1(ff+pN)+(fwwf)\displaystyle=a\partial_{1}\mathbb{P}_{\!\perp}(f\cdot\nabla f+\nabla p_{N})+\mathbb{P}_{\!\perp}(f\cdot\nabla w-w\cdot\nabla f)
+2(f1f)20(f1f2)1f\displaystyle\quad+2\mathbb{P}_{\!\perp}(f\cdot\nabla\partial_{1}f)-\partial_{2}\mathbb{P}_{0}(f_{1}f_{2})\partial_{1}f
+(f2220(f1f2)2a(ff2+2pN))e1.\displaystyle\quad+\left(f_{2}\partial_{2}^{2}\mathbb{P}_{0}(f_{1}f_{2})-\partial_{2}a\mathbb{P}_{\!\perp}(f\cdot\nabla f_{2}+\partial_{2}p_{N})\right)e_{1}\,.

From the above displayed equations and (5.15), the proof of (5.12) follows.∎

5.2. Properties of the linear operator L(f)L(f)

Lemma 5.5.

Suppose f(x1,x2)f(x_{1},x_{2}) is sufficiently regular such that

divf\displaystyle\mathrm{div\,}f =0\displaystyle=0 (5.19a)
0f\displaystyle\mathbb{P}_{0}f =0\displaystyle=0 (5.19b)

Then

divLf\displaystyle\mathrm{div\,}Lf =0\displaystyle=0
0(L(f))\displaystyle\mathbb{P}_{0}(L(f)) =0\displaystyle=0
Proof of Lemma 5.5.

We can write LfLf as

Lf=1((1+a)21f+(1+a)pL22a2(Δ)1(2af)e1)Lf=\partial_{1}\left((1+a)^{2}\partial_{1}f+(1+a)\nabla p_{L}-2\partial_{2}a\partial_{2}(-\Delta)^{-1}(\partial_{2}af)e_{1}\right)

Therefore, assuming ff is sufficiently regular, we conclude for each tt and x2x_{2} we have 0(L(f))(x2,t)=0\mathbb{P}_{0}(L(f))(x_{2},t)=0. Furthermore

divLf\displaystyle\mathrm{div\,}Lf =(1+a)212divf+(1+a)Δ1pL+2(1+a)212f22a12pL+2(1+a)12pL\displaystyle=(1+a)^{2}\partial_{1}^{2}\mathrm{div\,}f+(1+a)\Delta\partial_{1}p_{L}+\partial_{2}(1+a)^{2}\partial_{1}^{2}f_{2}-\partial_{2}a\partial_{1}\partial_{2}p_{L}+\partial_{2}(1+a)\partial_{1}\partial_{2}p_{L}
=2(1+a)2a12f+2(1+a)2a12f2\displaystyle=-2(1+a)\partial_{2}a\partial_{1}^{2}f+2(1+a)\partial_{2}a\partial_{1}^{2}f_{2}
=0,\displaystyle=0\,, (5.21)

which concludes the proof. ∎

Remark 5.6 (Solvability of the linear equation).

Now let us consider the evolution equation

tf=Lf,f|t=0=f0,\partial_{t}f=Lf\,,\qquad f|_{t=0}=f_{0}\,, (5.22)

where the initial data f0f_{0} satisfies (5.19), i.e. it is divergence free and its zero frequency in the x1x_{1} variable is trivial. Using Lemma 5.5 and the energy estimates done in Proposition 5.7 we can show that for sufficiently regular initial data f0f_{0}, the unique solution ff of (5.22) also satisfies (5.19).

Before we state our main semigroup estimate, Proposition 5.7 below, we specify the function spaces where we consider the evolution of solutions of (5.22). For kk\in\mathbb{N} we define

H˙0k:={fHk(𝕋2;2):0fj=0,j{1,2}}.\dot{H}_{0}^{k}:=\left\{f\in H^{k}(\mathbb{T}^{2};\mathbb{R}^{2})\colon\mathbb{P}_{0}f_{j}=0,j\in\{1,2\}\right\}. (5.23)

Note that H˙0k\dot{H}_{0}^{k} embeds into H˙0l\dot{H}_{0}^{l} for any lkl\leq k because Poincare’s inequality holds.

Proposition 5.7 (Linear decay estimates).

Let ff be a solution of (5.22). For any δ(0,1)\delta\in(0,1) there exists ε0>0\varepsilon_{0}>0 such that for any 0<εε00<\varepsilon\leq\varepsilon_{0} if

a(,t)Wk+1,4εt[0,T]\|a(\cdot,t)\|_{W^{k+1,\infty}}\leq 4\varepsilon\quad t\in[0,T] (5.24)

then

eLtH˙0kH˙0ke(1δ)tt[0,T]\|e^{Lt}\|_{\dot{H}_{0}^{k}\to\dot{H}_{0}^{k}}\leq e^{-(1-\delta)t}\quad t\in[0,T] (5.25)

where kk is as in Theorem 5.1.

Proof of Proposition 5.7.

Differentiating (5.13a) kk times with respect to 1\partial_{1}, multiplying by 1kf\partial_{1}^{k}f, and then integrating gives

12ddt1kfL22+(1+a)1k+1fL22=(1+a)1k+1pL,1kf2a21kpL,1kf1\displaystyle\frac{1}{2}\frac{d}{dt}\|\partial_{1}^{k}f\|_{L^{2}}^{2}+\|(1+a)\partial_{1}^{k+1}f\|_{L^{2}}^{2}=\langle(1+a)\nabla\partial_{1}^{k+1}p_{L},\partial_{1}^{k}f\rangle-\langle\partial_{2}a\partial_{2}\partial_{1}^{k}p_{L},\partial_{1}^{k}f_{1}\rangle (5.26)

where we have used that aa does not depend on x1x_{1}. Using the definition of pLp_{L} we have

(1+a)1k+1pL,1kf\displaystyle\langle(1+a)\nabla\partial_{1}^{k+1}p_{L},\partial_{1}^{k}f\rangle 1+aL1k+1pLL21kfL221+aL2aL1k+1fL22\displaystyle\leq\|1+a\|_{L^{\infty}}\|\nabla\partial_{1}^{k+1}p_{L}\|_{L^{2}}\|\partial_{1}^{k}f\|_{L^{2}}\leq 2\|1+a\|_{L^{\infty}}\|\partial_{2}a\|_{L^{\infty}}\|\partial_{1}^{k+1}f\|_{L^{2}}^{2}
2a21kpL,1kf1\displaystyle\langle\partial_{2}a\partial_{2}\partial_{1}^{k}p_{L},\partial_{1}^{k}f_{1}\rangle 2aL21kpLL2kfL222aL21k+1fL22\displaystyle\leq\|\partial_{2}a\|_{L^{\infty}}\|\partial_{2}\partial_{1}^{k}p_{L}\|_{L^{2}}\|\partial_{k}f\|_{L^{2}}\leq 2\|\partial_{2}a\|_{L^{\infty}}^{2}\|\partial_{1}^{k+1}f\|_{L^{2}}^{2} (5.27)

where we have used that Poincare’s inequality in the x1x_{1} variable holds with constant 11. For the given δ\delta, we can take ε0\varepsilon_{0} small enough such that if aWk+1,=εε0\|a\|_{W^{k+1,\infty}}=\varepsilon\leq\varepsilon_{0} then

12ddt1kfL22+(1δ2)1k+1fL220.\displaystyle\frac{1}{2}\frac{d}{dt}\|\partial_{1}^{k}f\|_{L^{2}}^{2}+\left(1-\frac{\delta}{2}\right)\|\partial_{1}^{k+1}f\|_{L^{2}}^{2}\leq 0. (5.28)

Repeating the same process with 2k\partial_{2}^{k} gives

12ddt2kfL22+(1+a)2k1fL22\displaystyle\frac{1}{2}\frac{d}{dt}\|\partial_{2}^{k}f\|_{L^{2}}^{2}+\|(1+a)\partial_{2}^{k}\partial_{1}f\|_{L^{2}}^{2}
=0<dkcd,k2d(1+a)22kd12f,2kf+0dkcd,k2d(1+a)2kd1pL,2kf\displaystyle=\sum_{0<d\leq k}c_{d,k}\langle\partial_{2}^{d}(1+a)^{2}\partial_{2}^{k-d}\partial_{1}^{2}f,\partial_{2}^{k}f\rangle+\sum_{0\leq d\leq k}c_{d,k}\langle\partial_{2}^{d}(1+a)\partial_{2}^{k-d}\nabla\partial_{1}p_{L},\partial_{2}^{k}f\rangle
0dkcd,k2d2a2kd2pL,2kf1=0<dkT1,d+0dk(T2,dT3,d)\displaystyle-\sum_{0\leq d\leq k}c_{d,k}\langle\partial_{2}^{d}\partial_{2}a\partial_{2}^{k-d}\partial_{2}p_{L},\partial_{2}^{k}f_{1}\rangle=\sum_{0<d\leq k}T_{1,d}+\sum_{0\leq d\leq k}(T_{2,d}-T_{3,d})

We now bound Ti,dT_{i,d}:

T1,d\displaystyle T_{1,d} :=cd,k2d(1+a)22kd12f,2kf\displaystyle:=c_{d,k}\langle\partial_{2}^{d}(1+a)^{2}\partial_{2}^{k-d}\partial_{1}^{2}f,\partial_{2}^{k}f\rangle
=cd,k2d(1+a)22kd1f,2k1f\displaystyle=-c_{d,k}\langle\partial_{2}^{d}(1+a)^{2}\partial_{2}^{k-d}\partial_{1}f,\partial_{2}^{k}\partial_{1}f\rangle
cd,k2d(1+a)2L2kd1fL22k1fL2\displaystyle\leq c_{d,k}\|\partial_{2}^{d}(1+a)^{2}\|_{L^{\infty}}\|\partial_{2}^{k-d}\partial_{1}f\|_{L^{2}}\|\partial_{2}^{k}\partial_{1}f\|_{L^{2}}
cd,kcj,d2j(1+a)L2dj(1+a)L2kd1fL22k1fL2\displaystyle\leq c_{d,k}c_{j,d}\|\partial_{2}^{j}(1+a)\|_{L^{\infty}}\|\partial_{2}^{d-j}(1+a)\|_{L^{\infty}}\|\partial_{2}^{k-d}\partial_{1}f\|_{L^{2}}\|\partial_{2}^{k}\partial_{1}f\|_{L^{2}} (5.29)

where 0<dk0<d\leq k and 0jd0\leq j\leq d. Since dd is never 0, this ensures either 2j(1+a)=2ja\partial_{2}^{j}(1+a)=\partial_{2}^{j}a or 2dj(1+a)=2dja\partial_{2}^{d-j}(1+a)=\partial_{2}^{d-j}a which are smaller than ε\varepsilon in LL^{\infty}. Therefore, by choosing ε0\varepsilon_{0} sufficiently small, we can take these terms as small as we want. Similarly,

T2,d\displaystyle T_{2,d} :=cd,k2d(1+a)2kd1pL,2kf\displaystyle:=c_{d,k}\langle\partial_{2}^{d}(1+a)\partial_{2}^{k-d}\nabla\partial_{1}p_{L},\partial_{2}^{k}f\rangle
2cd,k2d(1+a)L2kd(2a1f2)L22kfL2\displaystyle\leq 2c_{d,k}\|\partial_{2}^{d}(1+a)\|_{L^{\infty}}\|\partial_{2}^{k-d}(\partial_{2}a\partial_{1}f_{2})\|_{L^{2}}\|\partial_{2}^{k}f\|_{L^{2}}
2cd,kcj,kd2d(1+a)L2j+1aL2kdj1f2L22kfL2\displaystyle\leq 2c_{d,k}c_{j,k-d}\|\partial_{2}^{d}(1+a)\|_{L^{\infty}}\|\partial_{2}^{j+1}a\|_{L^{\infty}}\|\partial_{2}^{k-d-j}\partial_{1}f_{2}\|_{L^{2}}\|\partial_{2}^{k}f\|_{L^{2}}
2cd,kcj,kd2d(1+a)L2j+1aL1fH˙0k2\displaystyle\leq 2c_{d,k}c_{j,k-d}\|\partial_{2}^{d}(1+a)\|_{L^{\infty}}\|\partial_{2}^{j+1}a\|_{L^{\infty}}\|\partial_{1}f\|_{\dot{H}_{0}^{k}}^{2} (5.30)

and

T3,d\displaystyle T_{3,d} :=cd,k2d+1a2kd+1pL,2kf1\displaystyle:=c_{d,k}\langle\partial_{2}^{d+1}a\partial_{2}^{k-d+1}p_{L},\partial_{2}^{k}f_{1}\rangle
2cd,k2d+1aL2kd(2af2)L22kf1L2\displaystyle\leq 2c_{d,k}\|\partial_{2}^{d+1}a\|_{L^{\infty}}\|\partial_{2}^{k-d}(\partial_{2}af_{2})\|_{L^{2}}\|\partial_{2}^{k}f_{1}\|_{L^{2}}
2cd,kcj,kd2d+1aL2j+1aL2kdjf2L22kfL2\displaystyle\leq 2c_{d,k}c_{j,k-d}\|\partial_{2}^{d+1}a\|_{L^{\infty}}\|\partial_{2}^{j+1}a\|_{L^{\infty}}\|\partial_{2}^{k-d-j}f_{2}\|_{L^{2}}\|\partial_{2}^{k}f\|_{L^{2}}
2cd,kcj,kd2d+1aL2j+1aL1fH˙0k2.\displaystyle\leq 2c_{d,k}c_{j,k-d}\|\partial_{2}^{d+1}a\|_{L^{\infty}}\|\partial_{2}^{j+1}a\|_{L^{\infty}}\|\partial_{1}f\|_{\dot{H}_{0}^{k}}^{2}. (5.31)

Combining the estimates for (5.2), (5.2), and (5.2) gives

12ddt2kfL22+(1+a)2k1fL22Cε1fH˙0k2.\displaystyle\frac{1}{2}\frac{d}{dt}\|\partial_{2}^{k}f\|_{L^{2}}^{2}+\|(1+a)\partial_{2}^{k}\partial_{1}f\|_{L^{2}}^{2}\leq C\varepsilon\|\partial_{1}f\|_{\dot{H}_{0}^{k}}^{2}. (5.32)

Combining (5.32) with (5.28) and taking ε0\varepsilon_{0} sufficiently small we conclude

12ddtfH˙0k2(1δ)1fH˙0k(1δ)fH˙0k2\displaystyle\frac{1}{2}\frac{d}{dt}\|f\|_{\dot{H}_{0}^{k}}^{2}\leq-(1-\delta)\|\partial_{1}f\|_{\dot{H}_{0}^{k}}\leq-(1-\delta)\|f\|_{\dot{H}_{0}^{k}}^{2} (5.33)

which completes the proof. ∎

5.3. Proof of Theorem 5.1

The proof is based on the local existence result in Theorem 2.2, and a standard bootstrap argument for the bounds (5.7). Since 0b0=0\mathbb{P}_{0}b_{0}=0 we have that 𝕋2b0𝑑x1𝑑x2=0\int_{\mathbb{T}^{2}}b_{0}dx_{1}dx_{2}=0, and by appealing to (5.5a) we see that 𝕋2b(,t)𝑑x1𝑑x2=0\int_{\mathbb{T}^{2}}b(\cdot,t)dx_{1}dx_{2}=0 for all t0t\geq 0. It follows that BL22=bL22+|𝕋|2+2𝕋2b1𝑑x1𝑑x2=bL22+|𝕋|2\left\|B\right\|_{L^{2}}^{2}=\left\|b\right\|_{L^{2}}^{2}+|\mathbb{T}|^{2}+2\int_{\mathbb{T}^{2}}b_{1}dx_{1}dx_{2}=\left\|b\right\|_{L^{2}}^{2}+|\mathbb{T}|^{2}. Therefore, (2.2) and (5.6) imply that

b(,t)L2b0L2ε\displaystyle\left\|b(\cdot,t)\right\|_{L^{2}}\leq\left\|b_{0}\right\|_{L^{2}}\leq\varepsilon (5.34)

for all t0t\geq 0. Moreover, (5.6) and (2.9) show that there exists T0>0T_{0}>0 such that for all t[0,T0]t\in[0,T_{0}] we have that b(,t)H˙m2ε\left\|b(\cdot,t)\right\|_{\dot{H}^{m}}\leq 2\varepsilon. This bound may be combined with (5.34) to conclude that the bounds (5.7) hold on [0,T0][0,T_{0}], with all inequalities being strict inequalities. Due to the local existence result in Theorem 2.2 via a standard continuity argument we may thus define a maximal time T[T0,]T_{*}\in[T_{0},\infty] such that the estimates (5.7) hold on [0,T)[0,T_{*}). Our goal is to show that T=T_{*}=\infty. In order to achieve this we show that if (5.7) hold on [0,T][0,T] for some T>0T>0, then we may a posteriori show that these bounds in fact hold with constants 3ε3\varepsilon instead of 4ε4\varepsilon in (5.7a)–(5.7c); this then shows T=T_{*}=\infty. The rest of the proof is dedicated to establishing this implication, and so we fix a time interval [0,T][0,T], and we assume throughout that (5.7) hold. We recall and use the notation in (5.8).

5.3.1. Estimates for the nonlinear terms NN and NN^{\prime}

Under the standing assumptions, we estimate the nonlinear terms N(f,w)N(f,w) defined in (5.14a) and N(f,w)N^{\prime}(f,w) defined in (5.10) and claim that

N(f,w)(,t)Hkε2e32(1δ)t\|N(f,w)(\cdot,t)\|_{H^{k}}\leq\varepsilon^{2}e^{-\frac{3}{2}(1-\delta)t} (5.35)

and

N(f,w)(,t)Hk+2ε2e(1δ)t.\|N^{\prime}(f,w)(\cdot,t)\|_{H^{k+2}}\leq\varepsilon^{2}e^{-(1-\delta)t}. (5.36)

Prior to establishing (5.35) and (5.36), we claim that the pressure terms in (5.13b) and (5.14b) satisfy the bounds

pLHk+β\displaystyle\|p_{L}\|_{H^{k+\beta}} aHk+βfHkmkβ+1mkfHmβ1mk\displaystyle\lesssim\|a\|_{H^{k+\beta}}\|f\|_{H^{k}}^{\frac{m-k-\beta+1}{m-k}}\|f\|_{H^{m}}^{\frac{\beta-1}{m-k}} (5.37a)
pNHk+β\displaystyle\|p_{N}\|_{H^{k+\beta}} fHk2(mkβ+1)mkfHm2(β1)mk\displaystyle\lesssim\|f\|_{H^{k}}^{\frac{2(m-k-\beta+1)}{m-k}}\|f\|_{H^{m}}^{\frac{2(\beta-1)}{m-k}} (5.37b)

for 1β101\leq\beta\leq 10. The estimate for the linear pressure follows directly from (5.13b), the fact that HkH^{k} is an algebra, and interpolation:

pLHk+β2af2Hk+β1aHk+βf2Hk+β1aHk+βfHkmkβ+1mkfHmβ1mk.\displaystyle\|p_{L}\|_{H^{k+\beta}}\lesssim\|\partial_{2}af_{2}\|_{H^{k+\beta-1}}\lesssim\|a\|_{H^{k+\beta}}\|f_{2}\|_{H^{k+\beta-1}}\lesssim\|a\|_{H^{k+\beta}}\|f\|_{H^{k}}^{\frac{m-k-\beta+1}{m-k}}\|f\|_{H^{m}}^{\frac{\beta-1}{m-k}}\,.

Similarly, from (5.14b) we have

pNHk+β\displaystyle\|p_{N}\|_{H^{k+\beta}} (1f1)2+1f22f1Hk+β2fHk+β12fHk2(mkβ+1)mkfHm2(β1)mk.\displaystyle\lesssim\|(\partial_{1}f_{1})^{2}+\partial_{1}f_{2}\partial_{2}f_{1}\|_{H^{k+\beta-2}}\lesssim\|f\|_{H^{k+\beta-1}}^{2}\lesssim\|f\|_{H^{k}}^{\frac{2(m-k-\beta+1)}{m-k}}\|f\|_{H^{m}}^{\frac{2(\beta-1)}{m-k}}.

Next, we claim that the velocity field ww from (5.1) satisfies the estimate

wHk+β\displaystyle\|w\|_{H^{k+\beta}} aHk+β+1fHkmkβ1mkfHmβ+1mk+fHk2(mkβ1)mkfHm2(β+1)mk\displaystyle\lesssim\|a\|_{H^{k+\beta+1}}\|f\|_{H^{k}}^{\frac{m-k-\beta-1}{m-k}}\|f\|_{H^{m}}^{\frac{\beta+1}{m-k}}+\|f\|_{H^{k}}^{\frac{2(m-k-\beta-1)}{m-k}}\|f\|_{H^{m}}^{\frac{2(\beta+1)}{m-k}} (5.38)

for β0\beta\geq 0. This bound follows from (5.1), the previously established bounds (5.37b)–(5.37b), and an algebra + interpolation argument:

wHk+β\displaystyle\|w\|_{H^{k+\beta}} a1fHk+β+2af2Hk+β+ffHk+β+pLHk+β+pNHk+β\displaystyle\lesssim\|a\partial_{1}f\|_{H^{k+\beta}}+\|\partial_{2}af_{2}\|_{H^{k+\beta}}+\|f\cdot\nabla f\|_{H^{k+\beta}}+\|\nabla p_{L}\|_{H^{k+\beta}}+\|\nabla p_{N}\|_{H^{k+\beta}}
aHk+β+1fHk+β+1+fHk+β+12+pLHk+β+1+pLHk+β+1\displaystyle\lesssim\|a\|_{H^{k+\beta+1}}\|f\|_{H^{k+\beta+1}}+\|f\|_{H^{k+\beta+1}}^{2}+\|p_{L}\|_{H^{k+\beta+1}}+\|p_{L}\|_{H^{k+\beta+1}}
aHk+β+1fHkmkβ1mkfHmβ+1mk+fHk2(mkβ1)mkfHm2(β+1)mk\displaystyle\lesssim\|a\|_{H^{k+\beta+1}}\|f\|_{H^{k}}^{\frac{m-k-\beta-1}{m-k}}\|f\|_{H^{m}}^{\frac{\beta+1}{m-k}}+\|f\|_{H^{k}}^{\frac{2(m-k-\beta-1)}{m-k}}\|f\|_{H^{m}}^{\frac{2(\beta+1)}{m-k}}
+aHk+β+1fHkmkβmkfHmβmk+fHk2(mkβ)mkfHm2βmk.\displaystyle\qquad+\|a\|_{H^{k+\beta+1}}\|f\|_{H^{k}}^{\frac{m-k-\beta}{m-k}}\|f\|_{H^{m}}^{\frac{\beta}{m-k}}+\|f\|_{H^{k}}^{\frac{2(m-k-\beta)}{m-k}}\|f\|_{H^{m}}^{\frac{2\beta}{m-k}}\,.

The bound (5.38) follows by using the Poincaré inequality (recall that ff has zero mean on 𝕋2\mathbb{T}^{2}).

With (5.37b)–(5.37b) and (5.38) available, we next give the proof of (5.35). The right side of (5.14a) contains ten terms, and as such we estimate

N(f,w)HkN1++N10,\displaystyle\|N(f,w)\|_{H^{k}}\leq N_{1}+\ldots+N_{10}\,,

where

N1\displaystyle N_{1} :=a1(ff)HkaHkfHk+22aHkfHk24mkfHm4mk\displaystyle:=\|a\partial_{1}\mathbb{P}_{\!\perp}(f\cdot\nabla f)\|_{H^{k}}\lesssim\|a\|_{H^{k}}\|f\|_{H^{k+2}}^{2}\lesssim\|a\|_{H^{k}}\|f\|_{H^{k}}^{2-\frac{4}{m-k}}\|f\|_{H^{m}}^{\frac{4}{m-k}}
N2\displaystyle N_{2} :=a1pNHkaHkpNHk+2aHkfHk22mkfHm2mk\displaystyle:=\|a\partial_{1}\mathbb{P}_{\!\perp}\nabla p_{N}\|_{H^{k}}\lesssim\|a\|_{H^{k}}\|p_{N}\|_{H^{k+2}}\lesssim\|a\|_{H^{k}}\|f\|_{H^{k}}^{2-\frac{2}{m-k}}\|f\|_{H^{m}}^{\frac{2}{m-k}}
N3\displaystyle N_{3} :=(fw)HkfHkwHk+1aHk+2fHk22mkfHm2mk+fHk34mkfHm4mk\displaystyle:=\|\mathbb{P}_{\!\perp}(f\cdot\nabla w)\|_{H^{k}}\lesssim\|f\|_{H^{k}}\|w\|_{H^{k+1}}\lesssim\|a\|_{H^{k+2}}\|f\|_{H^{k}}^{2-\frac{2}{m-k}}\|f\|_{H^{m}}^{\frac{2}{m-k}}+\|f\|_{H^{k}}^{3-\frac{4}{m-k}}\|f\|_{H^{m}}^{\frac{4}{m-k}}
N4\displaystyle N_{4} :=(wf)HkwHkfHk+1aHk+1fHk22mkfHm2mk+fHk33mkfHm3mk\displaystyle:=\|\mathbb{P}_{\!\perp}(w\cdot\nabla f)\|_{H^{k}}\lesssim\|w\|_{H^{k}}\|f\|_{H^{k+1}}\lesssim\|a\|_{H^{k+1}}\|f\|_{H^{k}}^{2-\frac{2}{m-k}}\|f\|_{H^{m}}^{\frac{2}{m-k}}+\|f\|_{H^{k}}^{3-\frac{3}{m-k}}\|f\|_{H^{m}}^{\frac{3}{m-k}}
N5\displaystyle N_{5} :=20(f1f2)1fHkfHk+13fHk33mkfHm3mk\displaystyle:=\|\partial_{2}\mathbb{P}_{0}(f_{1}f_{2})\partial_{1}f\|_{H^{k}}\lesssim\|f\|_{H^{k+1}}^{3}\lesssim\|f\|_{H^{k}}^{3-\frac{3}{m-k}}\|f\|_{H^{m}}^{\frac{3}{m-k}}
N6\displaystyle N_{6} :=2(f1f)HkfHk+22fHk24mkfHm4mk\displaystyle:=2\|\mathbb{P}_{\!\perp}(f\cdot\nabla\partial_{1}f)\|_{H^{k}}\lesssim\|f\|_{H^{k+2}}^{2}\lesssim\|f\|_{H^{k}}^{2-\frac{4}{m-k}}\|f\|_{H^{m}}^{\frac{4}{m-k}}
N7\displaystyle N_{7} :=1pNHkpNHk+2fHk22mkfHm2mk\displaystyle:=\|\nabla\partial_{1}\mathbb{P}_{\!\perp}p_{N}\|_{H^{k}}\lesssim\|p_{N}\|_{H^{k+2}}\lesssim\|f\|_{H^{k}}^{2-\frac{2}{m-k}}\|f\|_{H^{m}}^{\frac{2}{m-k}}
N8\displaystyle N_{8} :=220(f1f2)f2HkfHk+23fHk36mkfHm6mk\displaystyle:=\|\partial_{2}^{2}\mathbb{P}_{0}(f_{1}f_{2})f_{2}\|_{H^{k}}\lesssim\|f\|_{H^{k+2}}^{3}\lesssim\|f\|_{H^{k}}^{3-\frac{6}{m-k}}\|f\|_{H^{m}}^{\frac{6}{m-k}}
N9\displaystyle N_{9} :=2a(ff2)HkaHk+1fHk+12aHk+1fHk22mkfHm2mk\displaystyle:=\|\partial_{2}a\mathbb{P}_{\!\perp}(f\cdot\nabla f_{2})\|_{H^{k}}\lesssim\|a\|_{H^{k+1}}\|f\|_{H^{k+1}}^{2}\lesssim\|a\|_{H^{k+1}}\|f\|_{H^{k}}^{2-\frac{2}{m-k}}\|f\|_{H^{m}}^{\frac{2}{m-k}}
N10\displaystyle N_{10} :=2a2pNHkaHk+1pNHk+1aHk+1fHk2,\displaystyle:=\|\partial_{2}a\partial_{2}\mathbb{P}_{\!\perp}p_{N}\|_{H^{k}}\lesssim\|a\|_{H^{k+1}}\|p_{N}\|_{H^{k+1}}\lesssim\|a\|_{H^{k+1}}\|f\|_{H^{k}}^{2}\,,

where the implicit constants only depend on mm and kk. At this point we use the assumption that mk9m-k\geq 9, the standing assumption (5.9), and take ε\varepsilon sufficiently small depending on δ\delta to obtain that (5.35) holds.

In a similar fashion, we estimate the nonlinear term in (5.10) as

N(f,w)Hk+2N1+N2+N3\displaystyle\|N^{\prime}(f,w)\|_{H^{k+2}}\leq N_{1}^{\prime}+N_{2}^{\prime}+N_{3}^{\prime}

with

N1\displaystyle N_{1}^{\prime} :=220(f21f1)Hk+2fHk+42fHk28mkfHm8mk\displaystyle:=2\|\partial_{2}\mathbb{P}_{0}(f_{2}\partial_{1}f_{1})\|_{H^{k+2}}\lesssim\|f\|_{H^{k+4}}^{2}\lesssim\|f\|_{H^{k}}^{2-\frac{8}{m-k}}\|f\|_{H^{m}}^{\frac{8}{m-k}}
N2\displaystyle N_{2}^{\prime} :=20(f2w1)Hk+2fHk+3wHk+3aHk+4fHk27mkfHm7mk+fHk311mkfHm11mk\displaystyle:=\|\partial_{2}\mathbb{P}_{0}(f_{2}w_{1})\|_{H^{k+2}}\lesssim\|f\|_{H^{k+3}}\|w\|_{H^{k+3}}\lesssim\|a\|_{H^{k+4}}\|f\|_{H^{k}}^{2-\frac{7}{m-k}}\|f\|_{H^{m}}^{\frac{7}{m-k}}+\|f\|_{H^{k}}^{3-\frac{11}{m-k}}\|f\|_{H^{m}}^{\frac{11}{m-k}}
N3\displaystyle N_{3}^{\prime} :=20(w2f1)Hk+2wHk+3fHk+3aHk+4fHk27mkfHm7mk+fHk311mkfHm11mk,\displaystyle:=\|\partial_{2}\mathbb{P}_{0}(w_{2}f_{1})\|_{H^{k+2}}\lesssim\left\|w\right\|_{H^{k+3}}\left\|f\right\|_{H^{k+3}}\lesssim\|a\|_{H^{k+4}}\|f\|_{H^{k}}^{2-\frac{7}{m-k}}\|f\|_{H^{m}}^{\frac{7}{m-k}}+\|f\|_{H^{k}}^{3-\frac{11}{m-k}}\|f\|_{H^{m}}^{\frac{11}{m-k}}\,,

where the implicit constant depends only on mm and kk. Once again, using mk9m-k\geq 9, the standing assumption (5.9), and the bound a(t)Hk+4a(t)Hmeεt\|a(t)\|_{H^{k+4}}\lesssim\|a(t)\|_{H^{m}}\lesssim e^{\varepsilon t}, after taking ε\varepsilon sufficiently small depending on δ\delta we have that (5.36) holds.

5.3.2. Closing the (5.7a) bootstrap

This argument is based on the Duhamel formula, which allows us via (5.12) to write

f(t)=eLtf0+0teL(ts)N(f,w)(s)𝑑s.f(t)=e^{Lt}f_{0}+\int_{0}^{t}e^{L(t-s)}N(f,w)(s)ds\,. (5.39)

Next, we note that aa is a function defined on 𝕋\mathbb{T}, and thus by the Sobolev embedding we have Hk+2Wk+1,H^{k+2}\subset W^{k+1,\infty}, which shows that (5.9) implies (5.24); thus, we may apply Proposition 5.7. Applying the H˙k\dot{H}^{k} norm to (5.39), and appealing to (5.35), we arrive at

f(t)H˙k\displaystyle\|f(t)\|_{\dot{H}^{k}} εe(1δ)t+0te(1δ)(ts)N(f,w)(s)H˙k𝑑s\displaystyle\leq\varepsilon e^{-(1-\delta)t}+\int_{0}^{t}e^{-(1-\delta)(t-s)}\|N(f,w)(s)\|_{\dot{H}^{k}}ds
εe(1δ)t+ε20te(1δ)(ts)e32(1δ)s𝑑s\displaystyle\leq\varepsilon e^{-(1-\delta)t}+\varepsilon^{2}\int_{0}^{t}e^{-(1-\delta)(t-s)}e^{-\frac{3}{2}(1-\delta)s}ds
εe(1δ)t(1+ε0te12(1δ)s𝑑s)\displaystyle\leq\varepsilon e^{-(1-\delta)t}\left(1+\varepsilon\int_{0}^{t}e^{-\frac{1}{2}(1-\delta)s}ds\right)
2εe(1δ)t,\displaystyle\leq 2\varepsilon e^{-(1-\delta)t}\,, (5.40)

once ε\varepsilon is chosen sufficiently small with respect to δ\delta. This bound improves on (5.7a), as desired.

5.3.3. Closing the (5.7b) bootstrap

By assumption, we have that a|t=0=0a|_{t=0}=0. Integrating the evolution equation for aa in (5.10), and appealing to (5.36), we thus deduce that

a(,t)Hk+20tN(f,w)(,s)Hk+2𝑑sε20te(1δ)s𝑑sε\displaystyle\left\|a(\cdot,t)\right\|_{H^{k+2}}\leq\int_{0}^{t}\left\|N^{\prime}(f,w)(\cdot,s)\right\|_{H^{k+2}}ds\leq\varepsilon^{2}\int_{0}^{t}e^{-(1-\delta)s}ds\leq\varepsilon

for all t[0,T]t\in[0,T], upon choosing ε\varepsilon to be sufficiently small in terms of δ\delta. This improves (5.7b) by a constant factor, as desired.

5.3.4. Closing the (5.7c) bootstrap

From (2.4), (5.3), (5.6), and the fact that BH˙s=bH˙s\left\|B\right\|_{\dot{H}^{s}}=\left\|b\right\|_{\dot{H}^{s}}, we deduce that for a constant CmC_{m} which only depends on mm, we have

b(,t)H˙m2\displaystyle\left\|b(\cdot,t)\right\|_{\dot{H}^{m}}^{2} ε2exp(Cm0tv(,s)L+1b(,s)L+b(,s)L2ds).\displaystyle\leq\varepsilon^{2}\exp\left(C_{m}\int_{0}^{t}\left\|\nabla v(\cdot,s)\right\|_{L^{\infty}}+\left\|\nabla\partial_{1}b(\cdot,s)\right\|_{L^{\infty}}+\left\|\nabla b(\cdot,s)\right\|_{L^{\infty}}^{2}ds\right)\,. (5.41)

From (5.9), (5.17), (5.38), and the fact that k>d/2+1=2k>d/2+1=2, we note that

0tv(,s)L𝑑s\displaystyle\int_{0}^{t}\left\|\nabla v(\cdot,s)\right\|_{L^{\infty}}ds 0tw(,s)Hk+f(,s)Hk+12ds\displaystyle\lesssim\int_{0}^{t}\left\|w(\cdot,s)\right\|_{H^{k}}+\left\|f(\cdot,s)\right\|_{H^{k+1}}^{2}ds
0ta(,s)Hk+1fHk11mkfHm1mk+fHk22mkfHm2mkds\displaystyle\lesssim\int_{0}^{t}\left\|a(\cdot,s)\right\|_{H^{k+1}}\left\|f\right\|_{H^{k}}^{1-\frac{1}{m-k}}\left\|f\right\|_{H^{m}}^{\frac{1}{m-k}}+\left\|f\right\|_{H^{k}}^{2-\frac{2}{m-k}}\left\|f\right\|_{H^{m}}^{\frac{2}{m-k}}ds
ε20te89(1δ)s+29εs𝑑sε,\displaystyle\lesssim\varepsilon^{2}\int_{0}^{t}e^{-\frac{8}{9}(1-\delta)s+\frac{2}{9}\varepsilon s}ds\lesssim\varepsilon\,, (5.42)

if ε\varepsilon is sufficiently small with respect to δ\delta. Returning to (5.41), we see that 1b=1f\nabla\partial_{1}b=\nabla\partial_{1}f, and since k>2+d/2=3k>2+d/2=3 and ff has zero mean on 𝕋2\mathbb{T}^{2}, we have that (5.9) implies

0t1b(,s)L𝑑s0tf(,s)Hk𝑑sε0te(1δ)s𝑑sε12.\displaystyle\int_{0}^{t}\left\|\nabla\partial_{1}b(\cdot,s)\right\|_{L^{\infty}}ds\lesssim\int_{0}^{t}\left\|f(\cdot,s)\right\|_{H^{k}}ds\lesssim\varepsilon\int_{0}^{t}e^{-(1-\delta)s}ds\lesssim\varepsilon^{\frac{1}{2}}\,. (5.43)

Lastly, for the third term in (5.41) we similarly note that

0tb(,s)L2𝑑s0ta(,s)Hk2+f(,s)Hk2dsε2t+ε,\displaystyle\int_{0}^{t}\left\|\nabla b(\cdot,s)\right\|_{L^{\infty}}^{2}ds\lesssim\int_{0}^{t}\left\|a(\cdot,s)\right\|_{H^{k}}^{2}+\left\|f(\cdot,s)\right\|_{H^{k}}^{2}ds\lesssim\varepsilon^{2}t+\varepsilon\,, (5.44)

where the implicit constant only depends on mm and kk. By combining (5.41)–(5.44) we thus obtain that there exists a constant Cm,k>0C_{m,k}>0, which only depends on mm and kk, such that

b(,t)H˙m2ε2exp(Cm,k(ε2t+ε12))εexp(εt)\displaystyle\left\|b(\cdot,t)\right\|_{\dot{H}^{m}}^{2}\leq\varepsilon^{2}\exp\left(C_{m,k}(\varepsilon^{2}t+\varepsilon^{\frac{1}{2}})\right)\leq\varepsilon\exp(\varepsilon t)

upon taking ε\varepsilon to be sufficiently small, solely in terms of mm and kk. This bound improves on (5.7c), as desired.

5.3.5. Relaxation in the infinite time limit

We recall that the total velocity field has zero mean on 𝕋2\mathbb{T}^{2} and is given from (5.3)–(5.4) as u=1b+vu=\partial_{1}b+v, and vv is computed from ff and ww via (5.17). Since 1b=1f\partial_{1}b=\partial_{1}f, the fact that u(,t)Hk0\left\|u(\cdot,t)\right\|_{H^{k}}\to 0 as tt\to\infty, exponentially fast, now follows from (5.9) and (5.38):

u(,t)Hk1\displaystyle\left\|u(\cdot,t)\right\|_{H^{k-1}} =u(,t)H˙k1fH˙k+wH˙k+f1f2H˙k\displaystyle=\left\|u(\cdot,t)\right\|_{\dot{H}^{k-1}}\lesssim\left\|f\right\|_{\dot{H}^{k}}+\left\|w\right\|_{\dot{H}^{k}}+\left\|f_{1}f_{2}\right\|_{\dot{H}^{k}}
εe(1δ)t+ε3e(1δ)(mk1)εmkt+ε2e2(1δ)(mk1)2εmkt+ε2e2(1δ)t\displaystyle\lesssim\varepsilon e^{-(1-\delta)t}+\varepsilon^{3}e^{-\frac{(1-\delta)(m-k-1)-\varepsilon}{m-k}t}+\varepsilon^{2}e^{-\frac{2(1-\delta)(m-k-1)-2\varepsilon}{m-k}t}+\varepsilon^{2}e^{-2(1-\delta)t}
εe1δ2t.\displaystyle\lesssim\varepsilon e^{-\frac{1-\delta}{2}t}\,.

In order to conclude the proof of the theorem, we note that in view of (1.3a) and the fact that k4k\geq 4, we have

tbHk2=tBHk2BuHk1BHk1uHk1(1+bHk)uH˙k1\displaystyle\left\|\partial_{t}b\right\|_{H^{k-2}}=\left\|\partial_{t}B\right\|_{H^{k-2}}\lesssim\left\|B\otimes u\right\|_{H^{k-1}}\lesssim\left\|B\right\|_{H^{k-1}}\left\|u\right\|_{H^{k-1}}\lesssim(1+\left\|b\right\|_{H^{k}})\left\|u\right\|_{\dot{H}^{k-1}}

and thus in view of (5.7a)–(5.7b) we obtain

tbHk2εe1δ2t.\displaystyle\left\|\partial_{t}b\right\|_{H^{k-2}}\lesssim\varepsilon e^{-\frac{1-\delta}{2}t}\,.

The strong convergence limtb(,t)=b¯\lim_{t\to\infty}b(\cdot,t)=\overline{b}, with respect to the Hk2H^{k-2} norm, for an incompressible vector field b¯\overline{b} which has norm 4ε\leq 4\varepsilon, now follows from the above estimate and the fundamental theorem of calculus in time. The corresponding limiting relaxation state for the total magnetic field is then B¯=e1+b¯\overline{B}=e_{1}+\overline{b}.

6. Nonlinear instabilities in 3D

In this section we consider a class of two-and-a-half dimensional exact solutions of the three-dimensional MRE system (1.1), when γ=0\gamma=0, and show that for suitable choices of initial conditions, these solutions exhibit infinite time growth of gradients. These examples draw on an analogy with the 3D Euler equation, for which Yudovich [Yud74, Yud00] has constructed similar solutions. Theorem 6.1 below gives an example in which magnetic relaxation holds with respect to the L2L^{2} norm, but fails with respect to the H1H^{1} norm. Furthermore, as in the work of Elgindi and Masmoudi [EM20] we construct examples where the magnetic current grows exponentially in time, for all time.

To fix notation, for any vector x3x\in\mathbb{R}^{3}, we denote by xHx_{H} its first two horizontal components, i.e. xH=(x1,x2)x_{H}=(x_{1},x_{2}). We also write divH=H\mathrm{div\,}_{H}=\nabla_{H}\cdot where H=(1,2)\nabla_{H}=(\partial_{1},\partial_{2}).

6.1. Euler examples

We recall from [Yud74, Yud00] the following two-and-a-half dimensional solution of 3D Euler, which exhibits infinite time growth of the vorticity.

The setting is as follows. Consider any stationary state v=v(xH)v=v(x_{H}) of the 2D Euler equations on 𝕋2\mathbb{T}^{2}. These stationary states may be written as v=Hϕv=\nabla_{H}^{\perp}\phi, where the periodic stream function ϕ:𝕋2\phi\colon\mathbb{T}^{2}\to\mathbb{R} satisfies ΔHϕ=F(ϕ)\Delta_{H}\phi=F(\phi) for a sufficiently smooth FF. Then, an exact solution of the 3D Euler system is given by

u(x,t)=(v(xH),g(xH,t))\displaystyle u(x,t)=(v(x_{H}),g(x_{H},t)) (6.1)

where the function g:𝕋2×+g\colon\mathbb{T}^{2}\times\mathbb{R}_{+}\to\mathbb{R} satisfies the transport equation

tg+(vH)g=0.\displaystyle\partial_{t}g+(v\cdot\nabla_{H})g=0\,. (6.2)

Indeed, one may verify that tu+uu=(vHv,tg+vHg)=(Hp,0)=p\partial_{t}u+u\cdot\nabla u=(v\cdot\nabla_{H}v,\partial_{t}g+v\cdot\nabla_{H}g)=(-\nabla_{H}p,0)=-\nabla p, where p=p(xH)p=p(x_{H}) is the pressure associated to the steady solution vv.

Shear flow.   When vv is a shear flow, such as

v(xH)=(V(x2),0)v(x_{H})=\left(V(x_{2}),0\right)

for a smooth function V:𝕋V\colon\mathbb{T}\to\mathbb{R}, the solution of (6.3) is explicit in terms of its initial datum g0g_{0}:

g(xH,t)=g0(x1tV(x2),x2).\displaystyle g(x_{H},t)=g_{0}(x_{1}-tV(x_{2}),x_{2})\,. (6.3)

Thus, combining (6.1) with (6.3), we are lead to the following exact solution of 3D Euler:

u(x,t)=(V(x2),0,g0(x1tV(x2),x2)).\displaystyle u(x,t)=\left(V(x_{2}),0,g_{0}(x_{1}-tV(x_{2}),x_{2})\right)\,. (6.4)

Even though u(,t)u(\cdot,t) remains bounded with respect to the LL^{\infty} norm, the vorticity ω=×u\omega=\nabla\times u has the property that its first component is given by

ω1(x,t)=(2u33u2)(x,t)=tV(x2)(1g0)(x1tV(x2),x2)+(2g0)(x1tV(x2),x2).\omega_{1}(x,t)=(\partial_{2}u_{3}-\partial_{3}u_{2})(x,t)=-tV^{\prime}(x_{2})(\partial_{1}g_{0})(x_{1}-tV(x_{2}),x_{2})+(\partial_{2}g_{0})(x_{1}-tV(x_{2}),x_{2})\,.

It is clear that for suitable choice of the initial datum g0g_{0}, and for VV\not\equiv constant, we have that ω1(,t)Lt\left\|\omega_{1}(\cdot,t)\right\|_{L^{\infty}}\gtrsim t as tt\to\infty. As such, in [Yud00], Yudovich calls the solution given in (6.4) as weakly nonlinearly unstable.

Hyperbolic flow.   In analogy with the above example, Elginidi and Masmoudi [EM20] consider the stationary solution v=v(xH)v=v(x_{H}) appearing in (6.1) to be an eigenfunction of the Laplacian which displays hyperbolic dynamics near the separatrix; more precisely, they consider

v=H(sinx1sinx2)=(sinx1cosx2cosx1sinx2).v=\nabla_{H}^{\perp}\left(\sin x_{1}\sin x_{2}\right)=\begin{pmatrix}-\sin x_{1}\cos x_{2}\\ \cos x_{1}\sin x_{2}\end{pmatrix}. (6.5)

In this case, the solution gg of the transport equation (6.3) is again bounded, but its derivative in the x1x_{1} direction, restricted to the separatrix {x2=0}\{x_{2}=0\} satisfies the equation

t(1g)|x2=0sinx11(1g)|x2=0=cosx1(1g)|x2=0,\displaystyle\partial_{t}(\partial_{1}g)|_{x_{2}=0}-\sin x_{1}\partial_{1}(\partial_{1}g)|_{x_{2}=0}=\cos x_{1}(\partial_{1}g)|_{x_{2}=0}\,,

and as such we have that (1g)(0,0,t)=(1g)(0,0,0)et(\partial_{1}g)(0,0,t)=(\partial_{1}g)(0,0,0)e^{t}. Thus, in this situation the solution uu of 3D Euler given by (6.1) exhibits exponential growth with respect to time of the first component of the vorticity

ω1(,t)Let|ω1(0,0,0)|.\displaystyle\left\|\omega_{1}(\cdot,t)\right\|_{L^{\infty}}\geq e^{t}|\omega_{1}(0,0,0)|\,.

6.2. MRE examples

The example of Yudovich outlined above, has a direct correspondent for the 3D MRE system. The main observation is that an exact solution of the 3D MRE equations (1.1) with γ=0\gamma=0 is given by

B=(v,g),u=(0,0,(vH)g)\displaystyle B=\left(v,g\right),\qquad u=\left(0,0,(v\cdot\nabla_{H})g\right) (6.6)

where g=g(xH,t):𝕋2×+g=g(x_{H},t)\colon\mathbb{T}^{2}\times\mathbb{R}_{+}\to\mathbb{R} satisfies the rank 11 diffusion equation

tg=(vH)2g.\displaystyle\partial_{t}g=(v\cdot\nabla_{H})^{2}g\,. (6.7)

In order to verify this, we start from the ansatz (6.6) and the fact that gg is independent of x3x_{3}, to immediately see that BB and uu are divergence free, so that (1.1c) holds. When γ=0\gamma=0, and with the ansatz (6.6), the first two components of (1.1b) become 0=vHv+Hp0=v\cdot\nabla_{H}v+\nabla_{H}p; this identity holds because vv was chosen to be an exact stationary state of the 2D Euler equations. On the other hand, the third component of (1.1b) reads u3=vHgu_{3}=v\cdot\nabla_{H}g, which justifies the definition of u3u_{3} in (6.6). Lastly, in view of (6.6) we have uB=0u\cdot\nabla B=0, since v,gv,g are independent of x3x_{3}, and Bu=e3(vH)(vH)gB\cdot\nabla u=e_{3}(v\cdot\nabla_{H})(v\cdot\nabla_{H})g. Therefore, (6.7) ensures that (1.1a) holds, as claimed.

Comparing the MRE evolution of gg in (6.7) to the Euler evolution of gg in (6.3), we see that the main difference is that gg does not solve a transport equation, but rather a rank 11 diffusion equation. Nonetheless, the gradient of gg may still exhibit infinite time growth, which is what we show next.

Shear flow.   The evolution equation (6.7) is particularly easy to solve if the 2D Euler steady state vv is chosen to be a shear flow. As such, consider

v(xH)=(V(x2),0)\displaystyle v(x_{H})=\left(V(x_{2}),0\right) (6.8)

for a smooth 𝕋\mathbb{T}-periodic scalar function VV. With (6.7), the evolution (6.7) becomes

tg=V2(x2)11g\displaystyle\partial_{t}g=V^{2}(x_{2})\partial_{11}g (6.9)

which is a heat equation in the (x1,t)(x_{1},t) variables, with viscosity coefficients that depend on x2x_{2}. In particular, if we choose an initial datum g0g_{0} which is just a function of x1x_{1} and such that its mean-free part is an eigenfunction of 11\partial_{11}, i.e.

11g0(x1)=λ2(g0(x1)𝕋g0)\displaystyle-\partial_{11}g_{0}(x_{1})=\lambda^{2}\left(g_{0}(x_{1})-\fint_{\mathbb{T}}g_{0}\right) (6.10)

for some λ>0\lambda>0, we have that the solution of equation (6.9) is given by

g(x1,x2,t)=𝕋g0+exp(λ2V2(x2)t)(g0(x1)𝕋g0).\displaystyle g(x_{1},x_{2},t)=\fint_{\mathbb{T}}g_{0}+\exp\left(-\lambda^{2}V^{2}(x_{2})t\right)\left(g_{0}(x_{1})-\fint_{\mathbb{T}}g_{0}\right)\,. (6.11)

We have thus shown that if the 2D Euler steady state is given by the shear flow in (6.8), and if the initial datum for the third component of the magnetic field is a function that satisfies (6.10), then for gg given by (6.11) the functions

B(x,t)=(V(x2),0,g(x1,x2,t)),u(x,t)=(0,0,V(x2)1g(x1,x2,t))\displaystyle B(x,t)=\left(V(x_{2}),0,g(x_{1},x_{2},t)\right),\qquad u(x,t)=\left(0,0,V(x_{2})\partial_{1}g(x_{1},x_{2},t)\right) (6.12)

are exact solutions of the 3D MRE equations (1.1) with γ=0\gamma=0.

In particular, the above example shows that solutions of (1.1) with γ=0\gamma=0 and d=3d=3 exhibit infinite time growth of gradients, even if the initial datum is a small perturbation of the steady state B=e3B=e_{3} and u=0u=0:

Theorem 6.1 (Example of solution with infinite time growth).

There exists an incompressible initial condition (B0,u0)(B_{0},u_{0}) such that B0e3=𝒪(ε)B_{0}-e_{3}=\mathcal{O}(\varepsilon) and u0=𝒪(ε2)u_{0}=\mathcal{O}(\varepsilon^{2}) in arbitrary strong topologies (e.g. real-analytic), and such that the unique solution BB of the 3D MRE equation (1.1) with this initial datum and γ=0\gamma=0 satisfies B(,t)L2=𝒪(ε3/2t1/4)\left\|\nabla B(\cdot,t)\right\|_{L^{2}}=\mathcal{O}(\varepsilon^{3/2}t^{1/4}) as tt\to\infty.

Proof of Theorem 6.1.

In (6.12), take V(x2)=εsin(x2)V(x_{2})=\varepsilon\sin(x_{2}), and g0(x1)=1+εcos(x1)g_{0}(x_{1})=1+\varepsilon\cos(x_{1}). This corresponds to the initial conditions

B0=e3+ε(sin(x2),0,cos(x1))andu0=ε2(0,0,sin(x2)sin(x1))\displaystyle B_{0}=e_{3}+\varepsilon(\sin(x_{2}),0,\cos(x_{1}))\qquad\mbox{and}\qquad u_{0}=-\varepsilon^{2}(0,0,\sin(x_{2})\sin(x_{1}))

which are clearly divergence free, and smooth. Due to the local existence and uniqueness theorem, we know that the solution is given by (6.12), where by (6.11) we have

g(x1,x2,t)=1+εcos(x1)exp(ε2sin2(x2)t).\displaystyle g(x_{1},x_{2},t)=1+\varepsilon\cos(x_{1})\exp\left(-\varepsilon^{2}\sin^{2}(x_{2})t\right)\,.

In particular, by (6.12) and the above formula, we have that

2B3(x,t)=2tε3sin(x2)cos(x2)cos(x1)exp(ε2sin2(x2)t)\displaystyle\partial_{2}B_{3}(x,t)=-2t\varepsilon^{3}\sin(x_{2})\cos(x_{2})\cos(x_{1})\exp\left(-\varepsilon^{2}\sin^{2}(x_{2})t\right)

and so we may explicitly compute

limt1C1ε3/2t1/42B3(,t)Lx1,x22\displaystyle\lim_{t\to\infty}\frac{1}{C_{1}\varepsilon^{3/2}t^{1/4}}\left\|\partial_{2}B_{3}(\cdot,t)\right\|_{L^{2}_{x_{1},x_{2}}} =1\displaystyle=1 (6.13a)
limt1C2ε2t1/22B3(,t)Lx12Lx2\displaystyle\lim_{t\to\infty}\frac{1}{C_{2}\varepsilon^{2}t^{1/2}}\left\|\partial_{2}B_{3}(\cdot,t)\right\|_{L_{x_{1}}^{2}L^{\infty}_{x_{2}}} =1\displaystyle=1 (6.13b)

where C1=(2π3)1/4C_{1}=(2\pi^{3})^{1/4}, C2=(2π2/e)1/2C_{2}=(2\pi^{2}/e)^{1/2}. Therefore, we have B(,t)L2=𝒪(ε3/2t1/4)\left\|\nabla B(\cdot,t)\right\|_{L^{2}}=\mathcal{O}(\varepsilon^{3/2}t^{1/4}) as tt\to\infty. ∎

Remark 6.2 (Asymptotic behavior with respect to weak topologies).

While the H˙1\dot{H}^{1} norm of the solution BB defined by (6.12) is growing without bound as time goes to infinity, we emphasize that its L2L^{2} norm remains uniformly bounded. In fact, for the solution in (6.12) we have the following asymptotic pointwise description

limtB(x,t)B¯(x):={(V(x2),0,𝕋g0),if V(x2)0(V(x2),0,g0(x1)),if V(x2)0\displaystyle\lim_{t\to\infty}B(x,t)\to\overline{B}(x):=\begin{cases}\left(V(x_{2}),0,\fint_{\mathbb{T}}g_{0}\right)\,,&\mbox{if }V(x_{2})\neq 0\\ \left(V(x_{2}),0,g_{0}(x_{1})\right)\,,&\mbox{if }V(x_{2})\neq 0\end{cases} (6.14)

and limtu(x,t)=0\lim_{t\to\infty}u(x,t)=0. This is thus an example of magnetic relaxation: BB converges to a magnetostatic equilibrium B¯\overline{B}, while uu converges to 0, as tt\to\infty. However, this relaxation holds with respect to weak topologies only (e.g. L2L^{2}), and weak nonlinear instability takes place in stronger topologies (e.g. H1H^{1}).

Remark 6.3 (The emergence of current sheets in the infinite time limit).

We note that even though the initial datum in the example of Remark 6.2 is smooth, namely B0=(V(x2),0,g0(x1))B_{0}=(V(x_{2}),0,g_{0}(x_{1})), the (weak) limiting magnetostatic equilibrium B¯\overline{B} may contain discontinuities in the vertical direction. For instance, take V(x2)=𝟏x2[π/2,π/2]cos2(x2)V(x_{2})={\bf 1}_{x_{2}\in[-\pi/2,\pi/2]}\cos^{2}(x_{2}), and g0(x1)=sin(x1)g_{0}(x_{1})=\sin(x_{1}). Then we have that the B¯\overline{B} vector field defined in (6.14) is given by

B¯(x)={(cos2(x2),0,0),if |x2|π/20(0,0,sin(x1)),if |x2|>π/20,\displaystyle\overline{B}(x)=\begin{cases}\left(\cos^{2}(x_{2}),0,0\right)\,,&\mbox{if }|x_{2}|\leq\pi/2\neq 0\\ \left(0,0,\sin(x_{1})\right)\,,&\mbox{if }|x_{2}|>\pi/2\neq 0\end{cases}\,,

which clearly contains a discontinuity along the planes {x𝕋3:x2=±π/2}\{x\in\mathbb{T}^{3}\colon x_{2}=\pm\pi/2\}. The associated current field j¯=×B¯\overline{j}=\nabla\times\overline{B} is given by the sum of a bounded piece 𝟏|x2|<π/2(0,0,sin(2x2))+𝟏|x2|>π/2(0,cos(x1),0){\bf 1}_{|x_{2}|<\pi/2}(0,0,\sin(2x_{2}))+{\bf 1}_{|x_{2}|>\pi/2}(0,-\cos(x_{1}),0), and a singular part which a Dirac mass supported on the planes x2=±π/2x_{2}=\pm\pi/2 and has amplitude (±sin(x1),0,0)(\pm\sin(x_{1}),0,0).

Hyperbolic flow.   While Theorem 6.1 exhibits solutions whose magnetic current grows algebraically in time as tt\to\infty, following [EM20] we may show that if vv is chosen to be the cellular flow

v=H(sinx1sinx2)=(sinx1cosx2cosx1sinx2),v=\nabla_{H}^{\perp}\left(\sin x_{1}\sin x_{2}\right)=\begin{pmatrix}-\sin x_{1}\cos x_{2}\\ \cos x_{1}\sin x_{2}\end{pmatrix}\,, (6.15)

then we can in fact find solutions gg of (6.7), and hence of the MRE equations, which exhibit exponential growth of their gradients. This is the worst growth that they can sustain, given that (6.7) is an equation linear in gg.

Theorem 6.4 (Example of solution with exponential growth).

Let vv be as in (6.15). For any initial data g0Hkg_{0}\in H^{k}, with k3k\geq 3 an integer, which satisfyies Hg0(0,0)0\nabla_{H}g_{0}(0,0)\neq 0, the unique solution of the 3D MRE equation (1.1) with initial data B0=(v,g0)B_{0}=(v,g_{0}) and u0=(0,0,vHg0)u_{0}=(0,0,v\cdot\nabla_{H}g_{0}), satisfies

|Hg0(0,0)|etB(,t)LCB0HkeCt,\left|\nabla_{H}g_{0}(0,0)\right|e^{t}\leq\|\nabla B(\cdot,t)\|_{L^{\infty}}\leq C\left\|B_{0}\right\|_{H^{k}}e^{Ct}\,, (6.16)

where C>0C>0 is a constant which only depends on kk.

We note that while the lower bound on the Lipschitz norm of BB given by (6.16) behaves as ete^{t}, in (3.8) we have obtained an upper bound which behaves as eCt1/2e^{Ct^{1/2}} as tt\to\infty; this difference stems from the fact that (6.16) holds for γ=0\gamma=0, while (3.8) holds for γ>d/2+1=5/2\gamma>d/2+1=5/2. This indicates a different behavior between the MRE equation (γ=0\gamma=0 in (1.1)), and the regularized MRE equation (γ>0\gamma>0 is large).

Proof of Theorem 6.4.

With vv as defined in (6.15), the gg equation (6.7) becomes

tg\displaystyle\partial_{t}g =sin2x1cos2x212g+cos2x1sin2x222g12sin(2x1)sin(2x2)12g\displaystyle=\sin^{2}x_{1}\cos^{2}x_{2}\partial_{1}^{2}g+\cos^{2}x_{1}\sin^{2}x_{2}\partial_{2}^{2}g-\tfrac{1}{2}\sin(2x_{1})\sin(2x_{2})\partial_{1}\partial_{2}g
+12sin(2x1)1g+12sin(2x2)2g.\displaystyle\qquad+\tfrac{1}{2}\sin(2x_{1})\partial_{1}g+\tfrac{1}{2}\sin(2x_{2})\partial_{2}g. (6.17)

The upper bound in (6.16) follows from the energy identity

12ddtgL22+(v)gL22=0,\frac{1}{2}\frac{d}{dt}\|g\|_{L^{2}}^{2}+\|(v\cdot\nabla)g\|_{L^{2}}^{2}=0\,,

the H˙k\dot{H}^{k} estimate for (6.7)

12ddtgH˙k2+(v)kgL22vHkvHk+1gHk2gHk2,\frac{1}{2}\frac{d}{dt}\|g\|_{\dot{H}^{k}}^{2}+\|(v\cdot\nabla)\nabla^{k}g\|_{L^{2}}^{2}\lesssim\|v\|_{H^{k}}\|v\|_{H^{k+1}}\|g\|_{H^{k}}^{2}\lesssim\|g\|_{H^{k}}^{2}\,,

and the fact that the condition k>2k>2 implies by the 2D Sobolev embedding that HkLipH^{k}\subset{\rm Lip}.

In order to obtain the lower bound in (6.16), we assume without loss of generality that 1g0(0,0)0\partial_{1}g_{0}(0,0)\neq 0 (the case 2g0(0,0)0\partial_{2}g_{0}(0,0)\neq 0 is treated in the same way). Then, we differentiate (6.2) with respect to x1x_{1} and arrive at the equation

t(1g)(0,0,t)=(1g)(0,0,t),\partial_{t}(\partial_{1}g)(0,0,t)=(\partial_{1}g)(0,0,t)\,, (6.18)

The exponential growth for the gradient of gg, and hence of BB in view of (6.6), with respect to the supremum norm now directly follows. ∎

7. Open problems

We conclude the paper by highlighting a number of interesting open problems for the magnetic relaxation equation.

7.1. Global well-posedness vs finite time blowup

We have shown in Theorem 2.2 that local existence and uniqueness of strong solutions for the MRE equations (1.1) holds irrespective of the regularization parameter γ0\gamma\geq 0. However, we were only able to prove that these solutions remain smooth for all time (necessary in order to consider the relaxation in the infinite time limit), when γ\gamma was sufficiently large, namely γ>d/2+1\gamma>d/2+1. Naturally, one is left to consider:

  • Q1.

    For γ[0,d/2+1]\gamma\in[0,d/2+1] can the local smooth solutions to the active vector equation (1.1) be extended to global ones, or do finite time singularities arise?

We emphasize that for d=2d=2 and for initial magnetic field B0B_{0} of zero mean, we may identify a zero mean scalar magnetic stream function ϕ=Δ1B\phi=\Delta^{-1}\nabla^{\perp}\cdot B, so that B=ϕB=\nabla^{\perp}\cdot\phi. Then, the evolution equation (1.1a) becomes the active scalar equation

tϕ+uϕ=0\displaystyle\partial_{t}\phi+u\cdot\nabla\phi=0 (7.1)

where the constitutive law ϕu\phi\mapsto u is given by

u=(Δ)γdiv(ϕϕ).\displaystyle u=(-\Delta)^{-\gamma}\mathbb{P}\mathrm{div\,}\big{(}\nabla^{\perp}\phi\otimes\nabla^{\perp}\phi\big{)}\,. (7.2)

The active scalar equation (7.1)–(7.2) has a quadratic in ϕ\phi constitutive law, and thus a cubic nonlinearity, making the analysis of the 2d MRE equation more cumbersome when compared to classical models in the canon of active scalar equations, such as SQG [CMT94] or IPM [CCGO09].

7.2. Magnetic relaxation

Assuming that the answer to Q1 is positive, i.e., that the smooth solutions to (1.1) are global in time (and thus the evolution (1.1a) truly is topology preserving), the fundamental question is whether as tt\to\infty the magnetic field B(,t)B(\cdot,t) relaxes to an MHD/Euler equilibrium B¯\overline{B} satisfying (1.4). We have discussed in Remark 4.2 the fact that the convergence of u(,t)0u(\cdot,t)\to 0 as tt\to\infty, even with respect to very strong norms, is in general not sufficient to guarantee that weak L2L^{2} subsequential limits of B(,t)B(\cdot,t) are magnetostatic equilibria, i.e., that they solve (1.4). Also, in view of Remark 4.3 and of the results in Section 6 we have shown that generically we cannot expect magnetic relaxation with respect to strong norms, such as H1H^{1} or Lip{\rm Lip}. Thus, we are naturally lead to:

  • Q2.

    Given a global in time solution B(,t)B(\cdot,t) to (1.1), does there exist a weak solution B¯L2(𝕋d)\overline{B}\in L^{2}(\mathbb{T}^{d}) of (1.4) and a subsequence tkt_{k}\to\infty such that we have the weak convergence B(,tk)B¯B(\cdot,t_{k})\rightharpoonup\overline{B}? Furthermore, does the answer change from d=2d=2 to d=3d=3, or for various values of γ\gamma?

Furthermore, in [Mof21, Section 8, Question (vii)] Moffatt poses the question:

  • Q3.

    For the magnetic relaxation problem (1.1), when the initial field is chaotic, what is the asymptotic structure of the relaxed field? Equivalently, what is the function space within which this relaxed field B¯\overline{B} must reside?

The examples given in Section 6 show that for the two-and-a-half dimensional solutions constructed via (6.6)–(6.7) the answer to Q2 is positive (see Remark 6.2), but that the magnetostatic equilibria B¯\overline{B} may contain current sheets (see Remark 6.3), so that they are not smooth. In fact, by the maximum principle it follows that all examples of steady states B¯\overline{B} arising as infinite time limits from the ansatz (6.6)–(6.7) will lie in LL^{\infty} due to the maximum principle, but our examples show that B¯\overline{B} cannot in general be expected to lie in C0C^{0}. We note also that for generic initial data it remains an open problem to show that the MRE evolution (1.1) is such that B(,t)Lp\left\|B(\cdot,t)\right\|_{L^{p}} remains uniformly bounded in time, for any p>2p>2. Thus, in general we do not know if the answer to question Q3 is better than B¯L2\overline{B}\in L^{2}.

7.3. Global weak solutions

In the absence of a positive answer to Q1, one may wonder whether the MRE system (1.1) at least possesses global weak solutions for generic initial data. Note that in order to define weak solutions a minimal requirement is that B(,t)Lx2B(\cdot,t)\in L^{2}_{x}, in order to properly define a distribution uu via (1.2). However, defining weak solutions to (1.1a) additionally requires that (BuuB)Lloc,t,x1(B\otimes u-u\otimes B)\in L^{1}_{{\rm loc},t,x}, which is for instance true when B(,t)L2B(\cdot,t)\in L^{2} if we also know that u(,t)Lx2u(\cdot,t)\in L^{2}_{x}. Note that when γ\gamma is not large (e.g. for γ=0\gamma=0), we cannot deduce from the square integrability of BB and (1.2) that uu is also square integrable (by the Sobolev embedding this would require γ>d/4+1/2\gamma>d/4+1/2). The energy inequality (2.2) comes to the rescue, providing for any γ0\gamma\geq 0 the required square integrability in space, locally in time, for the velocity field. A natural question thus is:

  • Q4.

    Given B0L2B_{0}\in L^{2} and γ0\gamma\geq 0 does there exist a global weak solution BLtLx2B\in L^{\infty}_{t}L^{2}_{x} of (1.1) which satisfies the energy inequality (2.2)? Alternatively, for γ(d/4+1/2,d/2+1]\gamma\in(d/4+1/2,d/2+1], does there exist a global weak solution BLtLx2B\in L^{\infty}_{t}L^{2}_{x} of (1.1) which does not satisfy energy inequality (2.2)?

The first part of question Q4 is nontrivial because the dissipative term in (2.2) does not yield robust compactness properties for the vector field BB. Note, however, that Brenier [Bre14], in the two dimensional case and with γ=0\gamma=0, managed to obtain global in time measure-valued solutions, a notion of solution which is weaker than the one of a weak solution, but which retains a weak-strong uniqueness property. Concerning the second part of question Q4, we note that the cubic nature of the nonlinear term in (1.1a) and the geometric properties of the constitutive law (1.1b), prevent the immediate application of convex integration techniques to the MRE system. Indeed, both the LL^{\infty}-based convex integration techniques of De Lellis-Szekelyhidi [DLS12] and the L2L^{2}-based intermittent convex integration developed by Buckmaster and the third author [BV21], do not seem to be directly applicable to the evolution equation (1.1), so that a potentially different convex integration method would need to be developed for the MRE system.

7.4. Other models

A number of other topology preserving diffusion equations have been proposed in the literature, which all have the property that the steady states are incompressible Euler equilibria. We mention for instance the models of Vallis-Carnevale-Young [VCY89] and Bloch-Marsden-Ratiu [BKMR96]. Other types of coercive damping mechanisms, which, however, do not preserve the topology of the streamlines, were considered in [Nis01, Nis03, Pas20]. Most if not all of the questions considered in this paper (global existence of solutions, relaxation towards Euler steady states as tt\to\infty) could be asked about those models. It would be interesting to compare (analytically or numerically) the long-time properties of solutions to the models in [VCY89] or [BKMR96], with those for the MRE equation. Is there any one model better suited for magnetic relaxation?

References

  • [AK98] V.I. Arnold and B.A. Khesin. Topological methods in hydrodynamics, volume 125 of Applied Mathematical Sciences. Springer-Verlag, New York, 1998.
  • [Arn66] V.I. Arnold. Sur la géométrie différentielle des groupes de Lie de dimension infinie et ses applications à l’hydrodynamique des fluides parfaits. Ann. Inst. Fourier (Grenoble), 16(fasc. 1):319–361, 1966.
  • [Arn74] V.I. Arnold. The asymptotic Hopf invariant and it applications. Proc. Summer School in Differential Equations, Erevan. Armenian SSR Acad. Sci., English translation, 561, 1974.
  • [BKMR96] A. Bloch, P.S. Krishnaprasad, J.E. Marsden, and T.S. Ratiu. The Euler-Poincaré equations and double bracket dissipation. Comm. Math. Phys., 175(1):1–42, 1996.
  • [Bre14] Y. Brenier. Topology-preserving diffusion of divergence-free vector fields and magnetic relaxation. Comm. Math. Phys., 330(2):757–770, 2014.
  • [BV21] T. Buckmaster and V. Vicol. Convex integration constructions in hydrodynamics. Bulletin of the American Mathematical Society, 58(1):1–44, 2021.
  • [CCGO09] Á. Castro, D. Córdoba, F. Gancedo, and R. Orive. Incompressible flow in porous media with fractional diffusion. Nonlinearity, 22(8):1791–1815, 2009.
  • [CCL19] Á. Castro, D. Córdoba, and D. Lear. Global existence of quasi-stratified solutions for the confined IPM equation. Arch. Ration. Mech. Anal., 232(1):437–471, 2019.
  • [CLV19] P. Constantin, J. La, and V. Vicol. Remarks on a paper by Gavrilov: Grad-Shafranov equations, steady solutions of the three dimensional incompressible Euler equations with compactly supported velocities, and applications. Geom. Funct. Anal., 29(6):1773–1793, 2019.
  • [CMT94] P. Constantin, A.J. Majda, and E. Tabak. Formation of strong fronts in the 22-D quasigeostrophic thermal active scalar. Nonlinearity, 7(6):1495–1533, 1994.
  • [DLS12] C. De Lellis and L. Székelyhidi, Jr. The hh-principle and the equations of fluid dynamics. Bull. Amer. Math. Soc. (N.S.), 49(3):347–375, 2012.
  • [DM87] R.J. DiPerna and A.J. Majda. Oscillations and concentrations in weak solutions of the incompressible fluid equations. Comm. Math. Phys., 108(4):667–689, 1987.
  • [Elg17] T.M. Elgindi. On the asymptotic stability of stationary solutions of the inviscid incompressible porous medium equation. Arch. Ration. Mech. Anal., 225(2):573–599, 2017.
  • [EM70] D.G. Ebin and J.E. Marsden. Groups of diffeomorphisms and the motion of an incompressible fluid. Ann. of Math. (2), 92:102–163, 1970.
  • [EM20] T.M. Elgindi and N. Masmoudi. LL^{\infty} ill-posedness for a class of equations arising in hydrodynamics. Arch. Ration. Mech. Anal., 235(3):1979–2025, 2020.
  • [EPS12] A. Enciso and D. Peralta-Salas. Knots and links in steady solutions of the Euler equation. Ann. of Math. (2), 175(1):345–367, 2012.
  • [FV95] S. Friedlander and M. Vishik. On stability and instability criteria for magnetohydrodynamics. Chaos, 5(2):416–423, 1995.
  • [Gav19] A.V. Gavrilov. A steady Euler flow with compact support. Geom. Funct. Anal., 29(1):190–197, 2019.
  • [HMRW85] D.D. Holm, J.E. Marsden, T. Ratiu, and A. Weinstein. Nonlinear stability of fluid and plasma equilibria. Phys. Rep., 123(1-2):116, 1985.
  • [Li19] D. Li. On Kato-Ponce and fractional Leibniz. Rev. Mat. Iberoam., 35(1):23–100, 2019.
  • [LXZ15] F. Lin, L. Xu, and P. Zhang. Global small solutions of 2-D incompressible MHD system. J. Differential Equations, 259(10):5440–5485, 2015.
  • [Mof85] H.K. Moffatt. Magnetostatic equilibria and analogous Euler flows of arbitrarily complex topology. I. Fundamentals. J. Fluid Mech., 159:359–378, 1985.
  • [Mof86] H.K. Moffatt. Magnetostatic equilibria and analogous Euler flows of arbitrarily complex topology. Pt. 2: stability considerations. Journal of Fluid Mechanics, 166:359–78, 1986.
  • [Mof21] H. K. Moffatt. Some topological aspects of fluid dynamics. J. Fluid Mech., 914:Paper No. P1, 2021.
  • [Nis01] T. Nishiyama. Pseudo-advection method for the two-dimensional stationary Euler equations. Proceedings of the American Mathematical Society, pages 429–432, 2001.
  • [Nis03] T. Nishiyama. Construction of three–dimensional stationary Euler flows from pseudo–advected vorticity equations. Proceedings of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences, 459(2038):2393–2398, 2003.
  • [Pas20] F. Pasqualotto. Nonlinear Waves in General Relativity and Fluid Dynamics. PhD thesis, Princeton University, 2020. Thesis (Ph.D.)–Princeton University.
  • [RWXZ14] X. Ren, J. Wu, Z. Xiang, and Z. Zhang. Global existence and decay of smooth solution for the 2-D MHD equations without magnetic diffusion. J. Funct. Anal., 267(2):503–541, 2014.
  • [SV93] L. Sadun and M. Vishik. The spectrum of the second variation of the energy for an ideal incompressible fluid. Phys. Lett. A, 182(4-6):394–398, 1993.
  • [VCY89] G.K. Vallis, G.F. Carnevale, and W.R. Young. Extremal energy properties and construction of stable solutions of the Euler equations. Journal of Fluid Mechanics, 207:133–152, 1989.
  • [Yud74] V.I. Yudovich. On the loss of smoothness of the solutions of euler’s equations with time. Dinamika Sploshnoi Sredy (Dynamics of Continuous Media) (Novosibirsk, Russia), 16:71–78, 1974.
  • [Yud00] V.I. Yudovich. On the loss of smoothness of the solutions of the Euler equations and the inherent instability of flows of an ideal fluid. Chaos, 10(3):705–719, 2000.