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

On Woltjer’s force free minimizers and Moffatt’s
magnetic relaxation

R. Komendarczyk Dedicated to the memory of Sławek Kwasik (1953–2021) Department of Mathematics, Tulane University, 6823 St. Charles Ave, New Orleans, LA 70118 [email protected] dauns01.math.tulane.edu/~rako
Abstract.

In this note, we exhibit a situation where a stationary state of Moffatt’s ideal magnetic relaxation problem is different than the corresponding force-free L2L^{2} energy minimizer of Woltjer’s variational principle. Such examples have been envisioned in Moffatt’s seminal work on the subject and involve divergence free vector fields supported on collections of essentially linked magnetic tubes. Justification of Moffatt’s examples requires the strong convergence of a minimizing sequence. What is proven in the current note is that there is a gap between the global minimum (Woltjer’s minimizer) and the minimum over the weak L2L^{2} closure of the class of vector fields obtained from a topologically non-trivial field by energy-decreasing diffeomorphisms. In the context of Taylor’s conjecture, our result shows that the Woltjer’s minimizer cannot be reached during the viscous MHD relaxation in the perfectly conducting magneto-fluid if the initial field has a nontrivial topology. The result also applies beyond Moffatt’s relaxation to any other relaxation process which evolves a divergence free field by means of energy-decreasing diffeomorphisms, such processes were proposed by Vallis et.al and more recently by Nishiyama.

2010 Mathematics Subject Classification:
Primary: 35Q35; Secondary: 58D05, 76W05

1. Introduction

The Woltjer’s variational problem [28], known in the context of hydrodynamics and magnetohydrodynamics [1, 2, 3, 5, 4, 9, 26, 19, 20, 17, 25, 23], concerns the minimization of the L2L^{2}–energy E(𝐁)=Ω𝐁(𝐱)2𝑑𝐱E(\mathbf{B})=\int_{\Omega}\|\mathbf{B}(\mathbf{x})\|^{2}d\mathbf{x} of over the subspace of divergence free vector fields defined on a regular domain Ω\Omega subject to a helicity constraint. Various boundary conditions, depending on the topology of Ω\Omega, can be imposed, we refer to [15] for further details, here we consider the case of a simply connected domain Ω\Omega with smooth connected boundary. The formal analysis, presented in [15], begins with the space

𝖫curl2(Ω)={𝐁𝖫2(Ω)|𝐁=curl(𝐀),div(𝐁)=0,𝐧𝐁=0onΩ,𝐀𝖫2(Ω)},\mathsf{L}^{2}_{\operatorname{curl}}(\Omega)=\{\mathbf{B}\in\mathsf{L}^{2}(\Omega)\ |\ \mathbf{B}=\operatorname{curl}(\mathbf{A}),\ \operatorname{div}(\mathbf{B})=0,\ \mathbf{n}\cdot\mathbf{B}=0\ \text{on}\ \partial\Omega,\ \mathbf{A}\in\mathsf{L}^{2}(\Omega)\}, (1)

(where the derivatives are understood in the weak sense, [14]) and seeks minimizers of E(𝐁)E(\mathbf{B}) subject to the constraint:

(𝐁)=Ω𝐁𝐀𝑑𝐱=(𝐁,𝐀)𝖫2=c,𝐁=curl(𝐀),c=const.\mathcal{H}(\mathbf{B})=\int_{\Omega}\mathbf{B}\cdot\mathbf{A}\,d\mathbf{x}=(\mathbf{B},\mathbf{A})_{\mathsf{L}^{2}}=c_{\ast},\qquad\mathbf{B}=\operatorname{curl}(\mathbf{A}),\qquad c_{\ast}=\text{const}.

The quantity (𝐁)\mathcal{H}(\mathbf{B}) is called the helicity of the field 𝐁\mathbf{B}, [1, 2, 18, 28] and is an invariant of 𝐁\mathbf{B}, under the volume preserving deformations i.e. (𝐁)=(f𝐁)\mathcal{H}(\mathbf{B})=\mathcal{H}(f_{\ast}\mathbf{B}) for any fDiff0(Ω,d𝐱)f\in\operatorname{Diff}_{0}(\Omega,d\mathbf{x}) of class C1C^{1} (i.e. volume preserving diffeomorphisms, which are equal to the identity along the boundary of Ω\Omega). For further reference we state the Woltjer’s problem, as follows

minimize E(𝐁) over 𝖶c(Ω)={𝐁𝖫curl2(Ω)|(𝐁)=c}.\text{minimize $E(\mathbf{B})$ over }\ \mathsf{W}_{c_{\ast}}(\Omega)=\{\mathbf{B}\in\mathsf{L}^{2}_{\operatorname{curl}}(\Omega)\ |\ \mathcal{H}(\mathbf{B})=c_{\ast}\}. (2)

As shown in [15], the minimizer 𝐁¯\overline{\mathbf{B}} exists and satisfies curl(𝐁¯)=λ𝐁¯\operatorname{curl}(\overline{\mathbf{B}})=\lambda\overline{\mathbf{B}}, λ\lambda\in\mathbb{R}, i.e. 𝐁¯\overline{\mathbf{B}} is an eigenfield of the operator curl\operatorname{curl}, and therefore a smooth classical solution of the Euler equations: 𝐁𝐁=P\mathbf{B}\cdot\nabla\mathbf{B}=-\nabla P, where 𝐁=0\nabla\cdot\mathbf{B}=0, and P=12𝐁2+constP=-\tfrac{1}{2}\|\mathbf{B}\|^{2}+\operatorname{const}.

In [18], Moffatt considered the following evolution equations of a viscous and perfectly conductive magneto-fluid

ρ(t𝐯+𝐯𝐯)=p+curl(𝐁)×𝐁+μ2𝐯,\displaystyle\rho\bigl{(}\partial_{t}\mathbf{v}+\mathbf{v}\cdot\nabla\mathbf{v}\bigr{)}=-\nabla p+\operatorname{curl}(\mathbf{B})\times\mathbf{B}+\mu\nabla^{2}\mathbf{v}, (3)
t𝐁=curl(𝐯×𝐁),𝐯=𝐁=0,\displaystyle\partial_{t}\mathbf{B}=\operatorname{curl}(\mathbf{v}\times\mathbf{B}),\quad\nabla\cdot\mathbf{v}=\nabla\cdot\mathbf{B}=0, (4)
𝐁(𝐱,0)=𝐁0(𝐱),𝐯(𝐱,0)=0,\displaystyle\mathbf{B}(\mathbf{x},0)=\mathbf{B}_{0}(\mathbf{x}),\quad\mathbf{v}(\mathbf{x},0)=0, (5)
𝐧𝐁=0,𝐯=0,onΩ,\displaystyle\mathbf{n}\cdot\mathbf{B}=0,\quad\mathbf{v}=0,\quad\text{on}\ \partial\Omega, (6)

where ρ\rho is the fluid density (assumed uniform), μ\mu viscosity (in [18] it is assumed sufficiently large when compared with Reynolds number associated with the flow) p(𝐱,t)p(\mathbf{x},t) is the pressure field. The second equation in (4) assures that 𝐁t=𝐁(𝐱,t)\mathbf{B}_{t}=\mathbf{B}(\mathbf{x},t) is transported with the flow ϕ𝐯\phi_{\mathbf{v}} of 𝐯\mathbf{v}, i.e.

𝐁(𝐱,t)=ϕ𝐯(𝐱,t)𝐁0(ϕ𝐯(𝐱,t)),\mathbf{B}(\mathbf{x},t)=\phi_{\mathbf{v}}(\mathbf{x},t)_{\ast}\mathbf{B}_{0}(\phi_{\mathbf{v}}(\mathbf{x},-t)), (7)

where ϕ𝐯(𝐱,t)\phi_{\mathbf{v}}(\mathbf{x},t)_{\ast} denotes a pushforward of the field under the diffeomorhism ϕ𝐯\phi_{\mathbf{v}}. Moffatt further shows that as long as 𝐯0\mathbf{v}\neq 0, the L2L^{2} energy of 𝐯\mathbf{v} and 𝐁\mathbf{B} decreases as tt\to\infty, by means of the following formula

ddt(E(𝐁t)+Ωρ𝐯(𝐱,t)2𝑑𝐱)=2Ωμcurl(𝐯(𝐱,t))2𝑑𝐱.\frac{d}{dt}\bigl{(}E(\mathbf{B}_{t})+\int_{\Omega}\rho\|\mathbf{v}(\mathbf{x},t)\|^{2}d\mathbf{x}\bigr{)}=-2\int_{\Omega}\mu\|\operatorname{curl}(\mathbf{v}(\mathbf{x},t))\|^{2}d\mathbf{x}. (8)

Further, he asserts that a minimizing sequence 𝐁t\mathbf{B}_{t}, tt\to\infty (c.f. [6, 22]) should yield a stationary state 𝐁=𝐁\mathbf{B}=\mathbf{B}_{\infty} satisfying the Euler’s equations. In [22], Nishiyama observes that a rigorous justification of convergence of 𝐁t\mathbf{B}_{t} to the stationary state is problematic due to the perfect conductivity of the magnetofluid, and introduces, guided by Vallis et.al, [27], an alternative to (3)–(6) system which admits a measure-valued solution in the sense of DiPerna and Majda [7].

Since the relaxation of the field 𝐁0\mathbf{B}_{0} according to (3)\eqref{eq:MHD-equation-1}(6)\eqref{eq:bdry-conditions} decreases its energy, a general question studied by Moffatt in [18] is as follows

Question A.

Is a stationary state 𝐁\mathbf{B}_{\infty} of the problem (3)\eqref{eq:MHD-equation-1}(6)\eqref{eq:bdry-conditions} (provided it exists) the same as the corresponding minimizer in the Woltjer’s variational problem (2)?

Refer to caption𝐁h+\mathbf{B}^{+}_{h}𝒯1+\mathcal{T}^{+}_{1}𝒯2+\mathcal{T}^{+}_{2}𝐁h\mathbf{B}^{-}_{h}𝒯1\mathcal{T}^{-}_{1}𝒯2\mathcal{T}^{-}_{2}
Figure 1. Vector field 𝐁h=𝐁h++𝐁h\mathbf{B}_{h}=\mathbf{B}^{+}_{h}+\mathbf{B}^{-}_{h} modeled on two Hopf links: Lh+L^{+}_{h}, LhL^{-}_{h} with opposite linking numbers has zero helicity, supported on the tubes {𝒯1±,𝒯2±}\{\mathcal{T}^{\pm}_{1},\mathcal{T}^{\pm}_{2}\}.

As illustrated in [18], one expects the minimizers to be different. In particular, in the case of fields with zero helicity, the force free minimizer of (2) is the zero field, however a nontrivial topology of the initial field 𝐁0\mathbf{B}_{0} can be still prevent a complete energy relaxation. The easiest examples where this situation occurs are the vector fields modeled on essential links and knots in 3{\mathbb{R}}^{3} (see Appendix), Figure 1 shows an example of a field modeled on the pair of Hopf links. In [18], among other examples, Moffatt considers the field modeled on Borromean rings 𝐁Borr\mathbf{B}_{Borr} and observes that the energy E(𝐁Borr)E(\mathbf{B}_{Borr}) cannot be decreased to zero under (4), (7) thanks to the lower bound of Freedman and He in [12] (see further remarks after the proof of Theorem A). These considerations however require a strong L2L^{2} convergence of a minimizing sequence 𝐁t\mathbf{B}_{t} obtained from (8), for tt\to\infty, which is in general problematic as discussed above.

In order to circumvent these problems, we observe that any minimizing sequence {𝐁t}\{\mathbf{B}_{t}\} obtained from (8) always has a weakly convergent subsequence. Thus we consider a variational problem associated with (3)–(6) which, asks to minimize the L2L^{2}–energy; EE over the subset

𝖬¯w(Ω,𝐁0)=the weak L2 closure of𝖬(Ω,𝐁0)={𝐁|𝐁=f𝐁0,fDiff0(Ω,d𝐱);E(𝐁)E(𝐁0)}𝖫curl2(Ω),\begin{split}\overline{\mathsf{M}}^{w}(\Omega,\mathbf{B}_{0})&=\text{the weak $L^{2}$ closure of}\\ &\mathsf{M}(\Omega,\mathbf{B}_{0})=\{\mathbf{B}\ |\ \mathbf{B}=f_{\ast}\mathbf{B}_{0},f\in\operatorname{Diff}_{0}(\Omega,d\mathbf{x});E(\mathbf{B})\leq E(\mathbf{B}_{0})\}\subset\mathsf{L}^{2}_{\operatorname{curl}}(\Omega),\end{split} (9)

of divergence free fields obtained from 𝐁0𝖫curl2(Ω)\mathbf{B}_{0}\in\mathsf{L}^{2}_{\operatorname{curl}}(\Omega) via pushforwards by volume preserving diffeomorphisms of Ω\Omega which are identity when restricted to Ω\partial\Omega (the C1C^{1} diffeomorphisms are denoted by Diff0(Ω,d𝐱)\operatorname{Diff}_{0}(\Omega,d\mathbf{x})). This is consistent with (4), since every vector field in 𝖬(Ω,𝐁0)\mathsf{M}(\Omega,\mathbf{B}_{0}) has the same topology as the initial field 𝐁0\mathbf{B}_{0}, and (4) simply defines a path111for the perfectly conducting magneto–fluid the long time existence is not known [22]. in 𝖬(Ω,𝐁0)\mathsf{M}(\Omega,\mathbf{B}_{0}). A clear difference with Woltjer’s problem is that the helicity constraint provides only a “mild” restriction on a topology of a field, whereas vector fields in 𝖬(Ω,𝐁0)\mathsf{M}(\Omega,\mathbf{B}_{0}) have equivalent topology to the initial 𝐁0\mathbf{B}_{0}.

In order to put it in the general setting, recall that a usual variational problem asks to minimize a weakly lower semicontinuous functional EE over a weakly compact class of functions WW, [8, 16]. One then considers a minimizing sequence fnWf_{n}\in W weakly convergent to fWf\in W, then E(f)liminfE(fn)E(f)\leq\lim\inf E(f_{n}). By the extreme value theorem for the weakly lower semicontinuous functions, ff is a minimizer of EE over WW. Note that 𝖶c(Ω)\mathsf{W}_{c_{\ast}}(\Omega) in (2), is weakly closed in 𝖫2(Ω)\mathsf{L}^{2}(\Omega) [15, p. 1237]. Since 𝖬(Ω,𝐁0)\mathsf{M}(\Omega,\mathbf{B}_{0}) is not weakly closed, we consider its weak 𝖫2\mathsf{L}^{2} closure 𝖬¯w(Ω,𝐁0)\overline{\mathsf{M}}^{w}(\Omega,\mathbf{B}_{0}) and ask to

minimize E(𝐁) over 𝖬¯w(Ω,𝐁0).\text{minimize $E(\mathbf{B})$ over }\ \overline{\mathsf{M}}^{w}(\Omega,\mathbf{B}_{0}). (10)

This formulation meets the requirements of the usual variational problem. Clearly, the caveat of replacing 𝖬(Ω,𝐁0)\mathsf{M}(\Omega,\mathbf{B}_{0}) by 𝖬¯w(Ω,𝐁0)\overline{\mathsf{M}}^{w}(\Omega,\mathbf{B}_{0}) is that the field line topology is no longer preserved and it is possible that

inf𝐁𝖬¯w(Ω,𝐁0)E(𝐁)<inf𝐁𝖬(Ω,𝐁0)E(𝐁),\inf_{\mathbf{B}\in\overline{\mathsf{M}}^{w}(\Omega,\mathbf{B}_{0})}E(\mathbf{B})<\inf_{\mathbf{B}\in\mathsf{M}(\Omega,\mathbf{B}_{0})}E(\mathbf{B}),

examples of paths in 𝖬(Ω,𝐁0)\mathsf{M}(\Omega,\mathbf{B}_{0}) where this possibility is realized were constructed in [16]. However, the relaxation (3)–(6), under the assumption of the long time existence of classical solutions, always takes place in 𝖬(Ω,𝐁0)\mathsf{M}(\Omega,\mathbf{B}_{0}), therefore the field line topology is preserved for “all time” except in the limit where it can be lost. Moffatt [18] defines fields topologically accessible from 𝐁0\mathbf{B}_{0} (i.e. limits of minimizing sequences 𝐁t\mathbf{B}_{t} obtaind from (4), (8)), here we simply allow such fields to be all possible weak L2L^{2} limits of fields from 𝖬(Ω,𝐁0)\mathsf{M}(\Omega,\mathbf{B}_{0}). Equivalently, one could say that under Moffatt’s assumption, topologically accessible fields are in the strong L2L^{2} closure of 𝖬(Ω,𝐁0)\mathsf{M}(\Omega,\mathbf{B}_{0}), whereas in general we may only show that they are in the weak L2L^{2} closure 𝖬¯w(Ω,𝐁0)\overline{\mathsf{M}}^{w}(\Omega,\mathbf{B}_{0}) of 𝖬(Ω,𝐁0)\mathsf{M}(\Omega,\mathbf{B}_{0}). Theorem A, presented in this work, shows that the global minimizer (Woltjer’s minimizer) is not in 𝖬¯w(Ω,𝐁0)\overline{\mathsf{M}}^{w}(\Omega,\mathbf{B}_{0}) for a field 𝐁0\mathbf{B}_{0} with a nontrivial linkage of flowlines.

Remark A.

We also note that in [11] a rotational magnetic field 𝐁Z\mathbf{B}_{Z} in the round ball Ω=B33\Omega=B^{3}\subset{\mathbb{R}}^{3} (Zeldovich’s neutron star) is considered and a path 𝐁t\mathbf{B}_{t} in 𝖬(Ω,𝐁Z)\mathsf{M}(\Omega,\mathbf{B}_{Z}) constructed such that 𝐁t0\mathbf{B}_{t}\to 0, as tt\to\infty in 𝖫2(Ω)\mathsf{L}^{2}(\Omega), however 𝐁t↛0\mathbf{B}_{t}\not\to 0 in 𝖫(Ω)\mathsf{L}^{\infty}(\Omega), which demonstrates that the minimizers may be highly irregular (see also [2]).

Our notation for the function spaces in the next section is as follows: 𝖢0,div(Ω)\mathsf{C}^{\infty}_{0,\operatorname{div}}(\Omega) smooth (test) divergence free compactly supported vector fields on Ω\Omega, 𝖫2(Ω)={X|XL2(Ω)×L2(Ω)×L2(Ω)}\mathsf{L}^{2}(\Omega)=\{X\ |\ X\in L^{2}(\Omega)\times L^{2}(\Omega)\times L^{2}(\Omega)\}, the square integrable vector fields, 𝖧1(Ω)={X|XH1(Ω)×H1(Ω)×H1(Ω)}\mathsf{H}^{1}(\Omega)=\{X\ |\ X\in H^{1}(\Omega)\times H^{1}(\Omega)\times H^{1}(\Omega)\} the Sobolev space of L2L^{2} vector fields with L2L^{2} weak derivatives.

2. Statement of the result

In a simply connected domain Ω3\Omega\subset{\mathbb{R}}^{3} with smooth connected boundary, let us consider as an initial vector field 𝐁h\mathbf{B}_{h} (i.e. 𝐁0=𝐁h\mathbf{B}_{0}=\mathbf{B}_{h}) modeled on a pair of Hopf links L=Lh+LhL=L^{+}_{h}\cup L^{-}_{h} (see Appendix A) as shown in Figure 1 and supported on the tubes {𝒯1±,𝒯2±}\{\mathcal{T}^{\pm}_{1},\mathcal{T}^{\pm}_{2}\}. It follows from the well known flux helicity formula [18], that the total helicity of 𝐁h\mathbf{B}_{h} is zero (we review this in the proof below) and therefore the force free minimizer of (2) is zero, [15].

Theorem A.

For the initial field 𝐁0=𝐁h\mathbf{B}_{0}=\mathbf{B}_{h}, a minimizer of the problem (10) is a nonzero field in 𝖫2(Ω)\mathsf{L}^{2}(\Omega).

Before presenting the proof, let us look closer at the construction of the divergence vector field 𝐁h=\mathbf{B}_{h}= with zero total helicity but nonzero subhelicites, i.e. (𝐁h)=0\mathcal{H}(\mathbf{B}_{h})=0 and (𝐁h+)=1\mathcal{H}(\mathbf{B}^{+}_{h})=1, (𝐁h)=1\mathcal{H}(\mathbf{B}^{-}_{h})=-1. Figure 1 illustrates the field supported on the tubes {𝒯1±,𝒯2±}\{\mathcal{T}^{\pm}_{1},\mathcal{T}^{\pm}_{2}\} about a 44–component link Lh={Lh+,Lh}L_{h}=\{L^{+}_{h},L^{-}_{h}\}, which is a disjoint union of two Hopf links Lh+=(Lh1,+,Lh2,+)L^{+}_{h}=(L^{1,+}_{h},L^{2,+}_{h}) and Lh=(Lh1,,Lh2,)L^{-}_{h}=(L^{1,-}_{h},L^{2,-}_{h}) with opposite linking numbers i.e.

lk(Lh+)=lk(Lh1,+,Lh2,+)=1,andlk(Lh)=lk(Lh1,,Lh2,)=1.\operatorname{lk}(L^{+}_{h})=\operatorname{lk}(L^{1,+}_{h},L^{2,+}_{h})=1,\qquad\text{and}\qquad\operatorname{lk}(L^{-}_{h})=\operatorname{lk}(L^{1,-}_{h},L^{2,-}_{h})=-1. (11)

Let Ω\Omega be a simply connected domain in 3{\mathbb{R}}^{3} with smooth boundary Ω\partial\Omega. We set 𝐁h\mathbf{B}_{h} to be the divergence free vector field modeled on LhL_{h} as defined in Appendix A supported on the disjoint tubes around the link LhL_{h}. Restricting 𝐁h\mathbf{B}_{h} to each individual tube we obtain

𝐁h=𝐁h++𝐁h=(𝐁h1,++𝐁h2,+)+(𝐁h1,+𝐁h2,).\mathbf{B}_{h}=\mathbf{B}^{+}_{h}+\mathbf{B}^{-}_{h}=(\mathbf{B}^{1,+}_{h}+\mathbf{B}^{2,+}_{h})+(\mathbf{B}^{1,-}_{h}+\mathbf{B}^{2,-}_{h}). (12)

We may assume that the fields 𝐁h+\mathbf{B}^{+}_{h} and 𝐁h+\mathbf{B}^{+}_{h} and the supporting tubes {𝒯1±,𝒯2±}\{\mathcal{T}^{\pm}_{1},\mathcal{T}^{\pm}_{2}\} are isometric images of each other in Ω3\Omega\subset{\mathbb{R}}^{3} as well as (the isometry needs to reverse the orientation in one of the tubes to obtain (11)).

The above construction yields the following helicity and cross-helicity identities

(𝐁h1,±)=(𝐁h2,±)=0,(𝐁h1,±,𝐁h2,±)=lk(Lh±)Φ(𝐁h,1±)Φ(𝐁h,2±),(𝐁h±,1,𝐁h2,)=0.\mathcal{H}(\mathbf{B}^{1,\pm}_{h})=\mathcal{H}(\mathbf{B}^{2,\pm}_{h})=0,\quad\mathcal{H}(\mathbf{B}^{1,\pm}_{h},\mathbf{B}^{2,\pm}_{h})=\operatorname{lk}(L^{\pm}_{h})\Phi(\mathbf{B}^{\pm}_{h,1})\Phi(\mathbf{B}^{\pm}_{h,2}),\quad\mathcal{H}(\mathbf{B}^{\pm,1}_{h},\mathbf{B}^{2,\mp}_{h})=0. (13)

Also, without loss of generality, we may scale the fields to obtain the unit fluxes i.e. Φ(𝐁h,±)=1\Phi(\mathbf{B}^{\pm}_{h,\ast})=1 and (𝐁h,1±,𝐁h,2±)=±1\mathcal{H}(\mathbf{B}^{\pm}_{h,1},\mathbf{B}^{\pm}_{h,2})=\pm 1. Further, \mathcal{H} is a symmetric bilinear, thus the above identities yield,

(𝐁h)=(𝐁h+)+(𝐁h)=0.\mathcal{H}(\mathbf{B}_{h})=\mathcal{H}(\mathbf{B}^{+}_{h})+\mathcal{H}(\mathbf{B}^{-}_{h})=0. (14)
Remark B.

Recall that the cross–helicity of two fields 𝐁1\mathbf{B}_{1} and 𝐁2\mathbf{B}_{2} in 𝖫curl2(Ω)\mathsf{L}^{2}_{\operatorname{curl}}(\Omega) is defined by

(𝐁1,𝐁2)=(𝐁1,𝐀2)𝖫2=Ω𝐁1(𝐱)𝐀2(𝐱)𝑑𝐱,𝐁2=curl(𝐀2)\mathcal{H}(\mathbf{B}_{1},\mathbf{B}_{2})=(\mathbf{B}_{1},\mathbf{A}_{2})_{\mathsf{L}^{2}}=\int_{\Omega}\mathbf{B}_{1}(\mathbf{x})\cdot\mathbf{A}_{2}(\mathbf{x})d\mathbf{x},\quad\mathbf{B}_{2}=\operatorname{curl}(\mathbf{A}_{2}) (15)

and is a symmetric bilinear form on 𝖫curl2(Ω)\mathsf{L}^{2}_{\operatorname{curl}}(\Omega). The single field helicity (𝐁)\mathcal{H}(\mathbf{B}) equals (𝐁,𝐁)\mathcal{H}(\mathbf{B},\mathbf{B}), i.e. the associated quadratic form.

Proof of Theorem A.

The set 𝖬(Ω,𝐁h)\mathsf{M}(\Omega,\mathbf{B}_{h}) is bounded in L2L^{2} norm, and therefore 𝖬¯w(Ω,𝐁h)\overline{\mathsf{M}}^{w}(\Omega,\mathbf{B}_{h}) is weakly compact. Since E()E(\,\cdot\,) is weakly lower semicontinuous, the extreme value theorem tells us that a minimizer 𝐁¯h\overline{\mathbf{B}}_{h} of EE exists over 𝖬¯w(Ω,𝐁h)\overline{\mathsf{M}}^{w}(\Omega,\mathbf{B}_{h}), i.e.

E(𝐁¯h)=min𝐁𝖬¯w(Ω,𝐁h)E(𝐁).E(\overline{\mathbf{B}}_{h})=\min_{\mathbf{B}\in\overline{\mathsf{M}}^{w}(\Omega,\mathbf{B}_{h})}E(\mathbf{B}).

By the Eberlein–Smulian Theorem [24], there is a sequence {𝐁h,n}𝖬(Ω,𝐁h)\{\mathbf{B}_{h,n}\}\subset\mathsf{M}(\Omega,\mathbf{B}_{h}), 𝐁h,n=fn,𝐁h\mathbf{B}_{h,n}=f_{n,\ast}\mathbf{B}_{h}, fnDiff0(Ω,d𝐱)f_{n}\in\operatorname{Diff}_{0}(\Omega,d\mathbf{x}), weakly convergent to 𝐁¯h\overline{\mathbf{B}}_{h}. The vector field push–forward ff_{\ast} is linear, so the decomposition (12) hold for every nn:

𝐁h,n=𝐁h,n++𝐁h,n,𝐁h,n±=fn,𝐁h±.\mathbf{B}_{h,n}=\mathbf{B}^{+}_{h,n}+\mathbf{B}^{-}_{h,n}\,,\qquad\mathbf{B}^{\pm}_{h,n}=f_{n,\ast}\mathbf{B}^{\pm}_{h}\,.

Since the supports of 𝐁h,n+\mathbf{B}^{+}_{h,n} and 𝐁h,n\mathbf{B}^{-}_{h,n} are disjoint

𝐁h,n±𝖫2𝐁h,n𝖫2𝐁h𝖫2,\|\mathbf{B}^{\pm}_{h,n}\|_{\mathsf{L}^{2}}\leq\|\mathbf{B}_{h,n}\|_{\mathsf{L}^{2}}\leq\|\mathbf{B}_{h}\|_{\mathsf{L}^{2}}, (16)

i.e. sequences {𝐁h,n+}\{\mathbf{B}^{+}_{h,n}\}, {𝐁h,n}\{\mathbf{B}^{-}_{h,n}\} are bounded and therefore weakly convergent (after passing to a subsequence, if necessary), let 𝐁h,n±𝐁¯h±\mathbf{B}^{\pm}_{h,n}\xrightharpoonup{\quad}\overline{\mathbf{B}}^{\pm}_{h} clearly, 𝐁¯h=𝐁¯h++𝐁¯h\overline{\mathbf{B}}_{h}=\overline{\mathbf{B}}^{+}_{h}+\overline{\mathbf{B}}^{-}_{h}. In the next step, we follow the analysis in [15, p. 1244]: for each 𝐁h,n±\mathbf{B}^{\pm}_{h,n}, we may choose a potential field 𝐀h,n±\mathbf{A}^{\pm}_{h,n} in 𝖧1(Ω)\mathsf{H}^{1}(\Omega), such that

curl(𝐀h,n±)=𝐁h,n±,\operatorname{curl}(\mathbf{A}^{\pm}_{h,n})=\mathbf{B}^{\pm}_{h,n},

in the weak sense (i.e. for any X𝖢0,div(Ω)X\in\mathsf{C}^{\infty}_{0,\operatorname{div}}(\Omega): (𝐀h,n±,curl(X))𝖫2=(𝐁h,n±,X)𝖫2(\mathbf{A}^{\pm}_{h,n},\operatorname{curl}(X))_{\mathsf{L}^{2}}=(\mathbf{B}^{\pm}_{h,n},X)_{\mathsf{L}^{2}}). The potential fields can be also chosen to satisfy

div(𝐀h,n±)=0,𝐀h,n±×𝐧=0,alongΩ,\operatorname{div}(\mathbf{A}^{\pm}_{h,n})=0,\quad\mathbf{A}^{\pm}_{h,n}\times\mathbf{n}=0,\ \text{along}\ \partial\Omega, (17)

where 𝐧\mathbf{n} is the unit normal along Ω\partial\Omega (these identities are in the weak and trace sense). By Friedrichs inequality [13], if 𝐀\mathbf{A} satisfies conditions in (17), then

𝐀𝖧1(Ω)2c1(Ω)curl(𝐀)𝖫22.\|\mathbf{A}\|^{2}_{\mathsf{H}^{1}(\Omega)}\leq c_{1}(\Omega)\|\operatorname{curl}(\mathbf{A})\|^{2}_{\mathsf{L}^{2}}.

From (16), sequences {𝐀h,n±}\{\mathbf{A}^{\pm}_{h,n}\}, {𝐀h,n}\{\mathbf{A}_{h,n}\} are bounded in 𝖧1(Ω)\mathsf{H}^{1}(\Omega), thus the Rellich compactness theorem [14] implies the following convergences (after passing to a subsequence if necessary)

𝐀h,n±,𝐀h,n𝐀¯h±,𝐀¯hstrongly in𝖫2,𝐁h,n±,𝐁h,n𝐁¯h±,𝐁¯hweakly in𝖫2.\begin{split}\mathbf{A}^{\pm}_{h,n},\mathbf{A}_{h,n}&\longrightarrow\overline{\mathbf{A}}^{\pm}_{h},\overline{\mathbf{A}}_{h}\qquad\text{strongly in}\ \mathsf{L}^{2},\\ \mathbf{B}^{\pm}_{h,n},\mathbf{B}_{h,n}&\xrightharpoonup{\quad}\overline{\mathbf{B}}^{\pm}_{h},\overline{\mathbf{B}}_{h}\qquad\text{weakly in}\ \mathsf{L}^{2}.\end{split} (18)

Suppose that, contrary to the statement of Theorem A, the minimizer of (10) is the zero field, i.e. 𝐁¯h=0\overline{\mathbf{B}}_{h}=0 in 𝖫2(Ω)\mathsf{L}^{2}(\Omega). Using the Hodge decomposition of [5, p. 879], on the simply connected Ω\Omega, for any X𝖢0,div(Ω)X\in\mathsf{C}^{\infty}_{0,\operatorname{div}}(\Omega), we have Y𝖫2(Ω)Y\in\mathsf{L}^{2}(\Omega), such that X=curl(Y)X=\operatorname{curl}(Y). By the weak convergence in (18)

(𝐀h,n,X)𝖫2=(curl(𝐀h,n),Y)𝖫2=(𝐁h,n,Y)𝖫20,asn,(\mathbf{A}_{h,n},X)_{\mathsf{L}^{2}}=(\operatorname{curl}(\mathbf{A}_{h,n}),Y)_{\mathsf{L}^{2}}=(\mathbf{B}_{h,n},Y)_{\mathsf{L}^{2}}\longrightarrow 0,\qquad\text{as}\quad n\to\infty,

where in the first identity we used222Ωcurl(W),V𝑑𝐱=±ΩW×V,𝐧𝑑σ+ΩW,curl(V)𝑑𝐱\int_{\Omega}\langle\operatorname{curl}(W),V\rangle d\mathbf{x}=\pm\int_{\partial\Omega}\langle W\times V,\mathbf{n}\rangle d\sigma+\int_{\Omega}\langle W,\operatorname{curl}(V)\rangle d\mathbf{x} the boundary condition (17).

Since the weak limit of 𝐀h,n\mathbf{A}_{h,n} is the zero field, the strong limit is also the zero field, i.e. 𝐀¯h=𝐀¯h++𝐀¯h=0\overline{\mathbf{A}}_{h}=\overline{\mathbf{A}}^{+}_{h}+\overline{\mathbf{A}}^{-}_{h}=0. From the computations in (13) and the helicity invariance under Diff0(Ω,d𝐱)\operatorname{Diff}_{0}(\Omega,d\mathbf{x}), we obtain

(𝐁h,n)=(𝐁h,n,𝐀h,n)𝖫2(𝐁¯h,𝐀¯h)𝖫2=(𝐁¯h)=0(𝐁h,n±)=(𝐁h,n±,𝐀h,n±)𝖫2(𝐁¯h±,𝐀¯h±)𝖫2=(𝐁¯h±)=±1,\begin{split}\mathcal{H}(\mathbf{B}_{h,n})&=(\mathbf{B}_{h,n},\mathbf{A}_{h,n})_{\mathsf{L}^{2}}\longrightarrow(\overline{\mathbf{B}}_{h},\overline{\mathbf{A}}_{h})_{\mathsf{L}^{2}}=\mathcal{H}(\overline{\mathbf{B}}_{h})=0\\ \mathcal{H}(\mathbf{B}^{\pm}_{h,n})&=(\mathbf{B}^{\pm}_{h,n},\mathbf{A}^{\pm}_{h,n})_{\mathsf{L}^{2}}\longrightarrow(\overline{\mathbf{B}}^{\pm}_{h},\overline{\mathbf{A}}^{\pm}_{h})_{\mathsf{L}^{2}}=\mathcal{H}(\overline{\mathbf{B}}^{\pm}_{h})=\pm 1,\end{split}

(because the inner product of the strongly convergent and weakly convergent sequences is convergent in {\mathbb{R}}.) The strong convergence: 𝐀h,n0\mathbf{A}_{h,n}\longrightarrow 0, implies333analogously for 𝐁h,n\mathbf{B}^{-}_{h,n} (𝐁h,n+,𝐀h,n)𝖫20(\mathbf{B}^{+}_{h,n},\mathbf{A}_{h,n})_{\mathsf{L}^{2}}\longrightarrow 0, but from (13)

(𝐁h,n+,𝐀h,n)𝖫2=(𝐁h,n+,𝐀h,n+)𝖫2+(𝐁h,n+,𝐀h,n)𝖫2=(𝐁h,n+)1,(\mathbf{B}^{+}_{h,n},\mathbf{A}_{h,n})_{\mathsf{L}^{2}}=(\mathbf{B}^{+}_{h,n},\mathbf{A}^{+}_{h,n})_{\mathsf{L}^{2}}+(\mathbf{B}^{+}_{h,n},\mathbf{A}^{-}_{h,n})_{\mathsf{L}^{2}}=\mathcal{H}(\mathbf{B}^{+}_{h,n})\longrightarrow 1,

since (𝐁h,n+,𝐀h,n)𝖫2=0(\mathbf{B}^{+}_{h,n},\mathbf{A}^{-}_{h,n})_{\mathsf{L}^{2}}=0 for every nn. Thus a contradiction to the assumption 𝐁¯h=0\overline{\mathbf{B}}_{h}=0. ∎

If one could assume a strong L2L^{2} convergence 𝐁h,n±𝐁¯h±\mathbf{B}^{\pm}_{h,n}\longrightarrow\overline{\mathbf{B}}^{\pm}_{h} in the proof of Theorem A then the classical energy–helicity estimate [1]: c1(Ω)|(𝐁)|E(𝐁)c_{1}(\Omega)|\mathcal{H}(\mathbf{B})|\leq E(\mathbf{B}), can be used to argue that 𝐁¯h0\overline{\mathbf{B}}_{h}\neq 0. Alternatively, one can use the asymptotic crossing number estimate of [10]. Therefore, these classical energy estimates suffice to show that there is a gap between the global minimum (the zero field, in this case) and the minimum over the strong L2L^{2} closure of 𝖬(Ω,𝐁0)\mathsf{M}(\Omega,\mathbf{B}_{0}). Theorem A shows a stronger fact i.e. that there is a gap between the global minimum and the minimum over the weak L2L^{2} closure 𝖬¯w(Ω,𝐁0)\overline{\mathsf{M}}^{w}(\Omega,\mathbf{B}_{0}). Clearly, this result applies beyond the Moffatt’s relaxation to any other relaxation process which evolves a divergence free field by means of energy-decreasing diffeomorphisms, such as Vallis [27] and Nishiyama [22]. In particular, Theorem A answers positively the question posed in [22, p. 417]. In the context of Taylor’s conjecture [26] it imples that the Woltjer’s minimizer cannot be reached during the perfectly conductive relaxation phase, if the inital field has a nontrivial topology. In the follow–up work we address this problem in the case of vector fields modeled on links exhibiting a higher order linkage (such as Borromean rings).

Appendix A Vector fields modeled on a link.

We begin by reviewing a definition of the divergence free vector field modeled on a link (c.f. [10]). Recall, that an nn–component link in 3\mathbb{R}^{3} is a smooth embedding444we often identify a link LL with is image in 3{\mathbb{R}}^{3}

L:k=1nSk13,Lk=L|Sk1n1,L:\bigsqcup^{n}_{k=1}S^{1}_{k}\longrightarrow{\mathbb{R}}^{3},\quad L_{k}=L\bigl{|}_{S^{1}_{k}}\qquad n\geq 1,

LL is called a trivial link if each component LkL_{k} is a boundary of an embedded disk, and the disks are disjoint from the link LL itself, otherwise the link is called nontrivial or essential. A divergence free vector field 𝐕=𝐕L\mathbf{V}=\mathbf{V}_{L} is said to be modeled on a link LL, [10], whenever there is a smooth volume preserving embedding

eL:k=1nDk2×Sk13,e_{L}:\bigsqcup^{n}_{k=1}D^{2}_{k}\times S^{1}_{k}\longrightarrow{\mathbb{R}}^{3},

of solid tori (tubes) 𝒯k=eL(Dk2×Sk1)\mathcal{T}_{k}=e_{L}(D^{2}_{k}\times S^{1}_{k}) into 3{\mathbb{R}}^{3} such that eL|{0}×Sk1=Lke_{L}|_{\{0\}\times S^{1}_{k}}=L_{k}, i.e. the cores of the tubes are mapped to the link LL. Further 𝐕L\mathbf{V}_{L} restricted to each 𝒯k\mathcal{T}_{k} is given by

𝐕L|𝒯k=(eL)(ϕk(𝐱)t),𝒯k=eL(Dk2×Sk1),\mathbf{V}_{L}\bigl{|}_{\mathcal{T}_{k}}=(e_{L})_{\ast}(\phi_{k}(\mathbf{x})\frac{\partial}{\partial t}),\qquad\mathcal{T}_{k}=e_{L}(D^{2}_{k}\times S^{1}_{k}),

where (𝐱,t)(\mathbf{x},t) are coordinates on Dk2×Sk1D^{2}_{k}\times S^{1}_{k} and ϕk:Dk2[0,1]\phi_{k}:D^{2}_{k}\longrightarrow[0,1] is a unit mass bump function vanishing in some neighborhood of Dk2\partial D^{2}_{k}. Observe that in each tube 𝒯k\mathcal{T}_{k} the vector field 𝐕L\mathbf{V}_{L} is the pushforward of Xk(𝐱,t)=ϕk(𝐱)tX_{k}(\mathbf{x},t)=\phi_{k}(\mathbf{x})\frac{\partial}{\partial t} and the circular orbits {𝐱}×Sk1\{\mathbf{x}\}\times S^{1}_{k} of XkX_{k} are mapped to the circular orbits γk(𝐱,t)\gamma_{k}(\mathbf{x},t) of 𝐕L\mathbf{V}_{L} in 𝒯k\mathcal{T}_{k}. Extending 𝐕L\mathbf{V}_{L} by zero to the entire domain we obtain a smooth vector field vanishing at 𝒯\partial\mathcal{T} (𝒯=k𝒯k\mathcal{T}=\bigcup_{k}\mathcal{T}_{k}), such that 𝐕L=k=1n𝐕k\mathbf{V}_{L}=\sum^{n}_{k=1}\mathbf{V}_{k}, where 𝐕k=𝐕L|𝒯k\mathbf{V}_{k}=\mathbf{V}_{L}\bigl{|}_{\mathcal{T}_{k}}. As observed in [10], the Moser’s result [21] can be used to make the embedding eLe_{L} volume preserving and thus 𝐕L\mathbf{V}_{L} a divergence free field (as XkX_{k} is itself divergence free). Further, eLe_{L} can be chosen such that lk(γk(𝐱,t),γk(𝐲,t))\operatorname{lk}(\gamma_{k}(\mathbf{x},t),\gamma_{k}(\mathbf{y},t)), 𝐱𝐲\mathbf{x}\neq\mathbf{y}, i.e. the pairwise linking of orbits of 𝐕L\mathbf{V}_{L} within each tube 𝒯k\mathcal{T}_{k} is zero, such 𝐕L\mathbf{V}_{L} then satisfies

(𝐕k)=0,𝐕k=𝐕L|𝒯k.\mathcal{H}(\mathbf{V}_{k})=0,\qquad\mathbf{V}_{k}=\mathbf{V}_{L}\bigl{|}_{\mathcal{T}_{k}}. (19)

References

  • [1] V. I. Arnold. The asymptotic Hopf invariant and its applications. Selecta Math. Soviet., 5(4):327–345, 1986. Selected translations.
  • [2] V. I. Arnold and B. A. Khesin. Topological methods in hydrodynamics, volume 125 of Applied Mathematical Sciences. Springer-Verlag, New York, 1998.
  • [3] M. A. Berger and G. B. Field. The topological properties of magnetic helicity. J. Fluid Mech., 147:133–148, 1984.
  • [4] T. Z. Boulmezaoud and T. Amari. On the existence of non-linear force-free fields in three-dimensional domains. Z. Angew. Math. Phys., 51(6):942–967, 2000.
  • [5] J. Cantarella, D. DeTurck, H. Gluck, and M. Teytel. The spectrum of the curl operator on spherically symmetric domains. Phys. Plasmas, 7(7):2766–2775, 2000.
  • [6] S. Childress. Topological fluid dynamics for fluid dynamicists. Online Notes, 2004.
  • [7] 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.
  • [8] I. Ekeland and R. Témam. Convex analysis and variational problems, volume 28 of Classics in Applied Mathematics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, english edition, 1999. Translated from the French.
  • [9] 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.
  • [10] M. H. Freedman. A note on topology and magnetic energy in incompressible perfectly conducting fluids. J. Fluid Mech., 194:549–551, 1988.
  • [11] M. H. Freedman. Zeldovich’s neutron star and the prediction of magnetic froth. In The Arnoldfest (Toronto, ON, 1997), volume 24 of Fields Inst. Commun., pages 165–172. Amer. Math. Soc., Providence, RI, 1999.
  • [12] M. H. Freedman and Z.-X. He. Divergence-free fields: energy and asymptotic crossing number. Ann. of Math. (2), 134(1):189–229, 1991.
  • [13] K. O. Friedrichs. Differential forms on Riemannian manifolds. Comm. Pure Appl. Math., 8:551–590, 1955.
  • [14] D. Gilbarg and N. S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1998 edition.
  • [15] P. Laurence and M. Avellaneda. On Woltjer’s variational principle for force-free fields. J. Math. Phys., 32(5):1240–1253, 1991.
  • [16] P. Laurence and E. Stredulinsky. Examples of topology breaking in 3d mhd. (report)., 2000.
  • [17] Francesca Maggioni and Renzo L. Ricca. On the groundstate energy of tight knots. Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci., 465(2109):2761–2783, 2009.
  • [18] H. K. Moffatt. Magnetostatic equilibria and analogous Euler flows of arbitrarily complex topology. I. Fundamentals. J. Fluid Mech., 159:359–378, 1985.
  • [19] H. K. Moffatt. The energy spectrum of knots and links. Nature, 347(6291):367–369, September 1990.
  • [20] H. K. Moffatt. Magnetic relaxation and the taylor conjecture. Journal of Plasma Physics, 81(6):905810608, 2015.
  • [21] J. Moser. On the volume elements on a manifold. Trans. Amer. Math. Soc., 120:286–294, 1965.
  • [22] T. Nishiyama. Magnetohydrodynamic approach to solvability of the three-dimensional stationary Euler equations. Glasg. Math. J., 44(3):411–418, 2002.
  • [23] C. Oberti and R. L. Ricca. Energy and helicity of magnetic torus knots and braids. Fluid Dyn. Res., 50(1):011413, 13, 2018.
  • [24] W. Rudin. Functional analysis. International Series in Pure and Applied Mathematics. McGraw-Hill, Inc., New York, second edition, 1991.
  • [25] A. J. B. Russell, A. R. Yeates, G. Hornig, and A. L. Wilmot-Smith. Evolution of field line helicity during magnetic reconnection. Physics of Plasmas, 22(3):032106, 2015.
  • [26] J. B. Taylor. Relaxation and magnetic reconnection in plasmas. Rev. Mod. Phys., 58:741–763, Jul 1986.
  • [27] G. K. Vallis, G. F. Carnevale, and W. R. Young. Extremal energy properties and construction of stable solutions of the Euler equations. J. Fluid Mech., 207:133–152, 1989.
  • [28] L. Woltjer. A theorem on force-free magnetic fields. Proc. Nat. Acad. Sci. U.S.A., 44:489–491, 1958.