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

Example of exponentially enhanced magnetic reconnection driven by a spatially-bounded and laminar ideal flow

Allen H Boozer and Todd Elder Columbia University, New York, NY 10027
[email protected]
Abstract

In laboratory and natural plasmas of practical interest, the spatial scale Δd\Delta_{d} at which magnetic field lines lose distinguishability differs enormously from the scale aa of magnetic reconnection across the field lines. In the solar corona, plasma resistivity gives a/Δd1012a/\Delta_{d}\sim 10^{12}, which is the magnetic Reynold number RmR_{m}. The traditional resolution of the paradox of disparate scales is for the current density jj associated with the reconnecting field BrecB_{rec} to be concentrated by a factor of RmR_{m} by the ideal evolution, so jBrec/μ0Δdj\sim B_{rec}/\mu_{0}\Delta_{d}. A second resolution is for the ideal evolution to increase the ratio of the maximum to minimum separation between pairs of arbitrarily chosen magnetic field lines, Δmax/Δmin\Delta_{max}/\Delta_{min}, when calculated at various points in time. Reconnection becomes inevitable where Δmax/ΔminRm\Delta_{max}/\Delta_{min}\sim R_{m}. A simple model of the solar corona will be used for a numerical illustration that the natural rate of increase in time is linear for the current density but exponential for Δmax/Δmin\Delta_{max}/\Delta_{min}. Reconnection occurs on a time scale and with a current density enhanced by only ln(a/Δd)\ln(a/\Delta_{d}) from the ideal evolution time and from the current density Brec/μ0aB_{rec}/\mu_{0}a. In both resolutions, once a sufficiently wide region, Δr\Delta_{r}, has undergone reconnection, the magnetic field loses static force balance and evolves on an Alfvénic time scale. The Alfvénic evolution is intrinsically ideal but expands the region in which Δmax/Δmin\Delta_{max}/\Delta_{min} is large.

I Introduction

Magnetic reconnection has an interesting early history Cargill:2015 and was defined in 1956 by Parker and Krook Parker-Krook:1956 as the “severing and reconnection of lines of force.

The ideal evolution of magnetic fields is the opposing concept to magnetic reconnection. Newcomb showed in 1958 that magnetic field lines move with a velocity u\vec{u}_{\bot} and do not break, if and only if the field obeys the ideal evolution equation Newcomb

Bt=×(u×B).\frac{\partial\vec{B}}{\partial t}=\vec{\nabla}\times(\vec{u}_{\bot}\times\vec{B}). (1)

The exact equation for the evolution of B\vec{B} is Faraday’s Law, B/t=×E\partial\vec{B}/\partial t=-\vec{\nabla}\times\vec{E}. The ideal evolution of a magnetic field that is embedded in a plasma is broken Boozer:ideal-ev when the electric field parallel to the magnetic field, E||\vec{E}_{||}, cannot be balanced by the parallel component of the gradient of a well-behaved potential, Φ\vec{\nabla}\Phi. The components E+Φ\vec{E}_{\bot}+\vec{\nabla}_{\bot}\Phi, which are perpendicular to B\vec{B}, are associated with the velocity u\vec{u}_{\bot} of the magnetic field lines, u=(E+Φ)×B/B2\vec{u}_{\bot}=(\vec{E}_{\bot}+\vec{\nabla}_{\bot}\Phi)\times\vec{B}/B^{2}.

Two effects that cause deviations from an ideal evolution are resistivity along the magnetic field η\eta, which contributes to E||E_{||} as ηj||\eta\vec{j}_{||}, and electron inertia, which contributes Spitzer to E||E_{||} as (c/ωpe)2j||/t(c/\omega_{pe})^{2}\partial j_{||}/\partial t. The electron skin depth, c/ωpec/\omega_{pe}, is the speed of light divided by the electron plasma frequency.

The Oxford English Dictionary defines a paradox as “a strongly counter-intuitive (statement), which investigation…may nevertheless prove to be well-founded or true.” Both the resistivity and the electron skin depth are so small in many plasmas of practical interest that it is paradoxical that magnetic reconnection could be of any relevance. Yet it is. Approximately 1500 papers have been written on magnetic reconnection, which demonstrate the importance of the topic to the understanding of both laboratory and naturally occurring plasmas. The speed and prevalence of magnetic reconnection are so great that they must be derivable from the properties of Equation (1) for the ideal evolution of a magnetic field Boozer:ideal-ev .

To define the reconnection paradox, let aa be a characteristic spatial scale over which uu_{\bot} varies across the magnetic field lines; the time scale for an ideal evolution is

τevau,\tau_{ev}\equiv\frac{a}{u_{\bot}}, (2)

Resistivity spatially interdiffuses magnetic field lines, which implies that lines that approach each other closer than the distance Δd=(η/μ0)τη\Delta_{d}=\sqrt{(\eta/\mu_{0})\tau_{\eta}} remain distinguishable only for a time tt less than τη\tau_{\eta}. The ratio of the resistive time scale of the overall system to the evolution time scale τev\tau_{ev} is the magnetic Reynolds number

Rm(a2η/μ0)(ua)=μ0uηa.R_{m}\equiv\left(\frac{a^{2}}{\eta/\mu_{0}}\right)\left(\frac{u_{\bot}}{a}\right)=\frac{\mu_{0}u_{\bot}}{\eta}a. (3)

Magnetic field lines that come closer than Δd=a/Rm\Delta_{d}=a/R_{m} lose their distinguishability, where Rm1012R_{m}\sim 10^{12} in the solar corona.

The electron inertia causes evolving magnetic field lines that approach each other closer than Δd=c/ωpe\Delta_{d}=c/\omega_{pe}, the electron skin depth, to become indistinguishable, Appendix C of Boozer:null-X . In the solar corona a/(c/ωpe)109a/(c/\omega_{pe})\sim 10^{9}.

What is paradoxical is that magnetic field lines are observed to reconnect and become indistinguishable over a region of width aa across the magnetic field lines on a time scale only an order of magnitude or so longer than the characteristic ideal-evolution time of the magnetic field, τev\tau_{ev}. Effects such as the resistivity would be expected to cause a loss of distinguishability of magnetic field lines in the solar corona only on a time scale of order 101210^{12} times longer than τev\tau_{ev}. This paradox was clearly described in 1988 by Schindler, Hesse, and Birn Schindler:1988 .

Schindler et al Schindler:1988 sought to resolve the paradox of reconnection occurring over a region of far greater width than Δd\Delta_{d}. In their resolution, the current density would become and remain extremely large, jBrec/(μ0Δd)j\sim B_{rec}/(\mu_{0}\Delta_{d}), in a layer of width Δd\Delta_{d}, where BrecB_{rec} is the part of the magnetic field that is reconnecting. This assumption has formed the basis of most of the reconnection literature from even before Schindler et al developed their reconnection theory. Nonetheless, it is difficult to understand how an ideal evolution would result in such a strong current. Most of the reconnection literature has not dealt with that issue but has focused instead on how such a large current density could be maintained if it were initially present. A foundational publication for recent work on the maintenance issue, with more than two hundred citations, is the 2010 paper by Uzdenksy, Louriero, and Schekochihin Loureiro:2010 on plasmoids.

The paradoxical speed of magnetic reconnection when non-ideal effects are extremely small has analogues with other phenomena, such as thermal equilibration in air. The analogy between magnetic reconnection and thermal equilibration in a room was demonstrated Boozer:rec-phys by Boozer in 2021. In both reconnection and thermal equilibration, the time scale of the relaxation of the ideal constraint is the ideal evolution time τev\tau_{ev} times a term that is logarithmically dependent on the ratio of the non-ideal time scale divided by the τev\tau_{ev}. This explains why a radiator can heat a room in tens of minutes rather than the several weeks that would be expected from thermal diffusion alone. Many may find the subtle mathematics used in Boozer:rec-phys difficult to follow. Here a simple model of magnetic field evolution driven by footpoint motion, as in the solar corona, is used to illustrate the general principles that were explained in the companion paper Boozer:rec-phys .

The equation for temperature equilibration in air moving with a divergence-free velocity v(x,t)\vec{v}(\vec{x},t) is T/t+vT=(DT)\partial T/\partial t+\vec{v}\cdot\vec{\nabla}T=\vec{\nabla}\cdot(D\vec{\nabla}T), which is the standard form for the advection-diffusion equation. In many problems of practical importance, the Péclet number, va/Dva/D, is many orders of magnitude greater than unity. The diffusion coefficient for magnetic field lines is η/μ0\eta/\mu_{0}, so the magnetic Reynolds number would be better called the magnetic Péclet number.

The classic paper on the advection-diffusion equation was written in 1984 by Hassan Aref Aref;1984 and has had over two thousand citations. This paper showed that a laminar flow can equilibrate TT on the evolution time a/va/v times a logarithm of the Péclet number as va/Dva/D\rightarrow\infty. Before Aref’s paper, it had been assumed a turbulent v\vec{v} was required to obtain fast equilibration. Section I.B in an article Aref:2017 published in the Reviews of Modern Physics discusses the merits of the use of turbulent versus laminar flows to speed the mixing of fluids. What that article did not discuss is that for a given maximum flow speed vv a laminar flow can give a faster mixing over a large region than a turbulent flow. A demonstration starts with the last sentence on p. 3 of Boozer:rec-phys . The laminar part of a flow, which is defined as an average over the small spatial scales of v(x,t)\vec{v}(\vec{x},t), generally dominates the advective part of the advection-diffusion equation. A flow v(x,t)\vec{v}(\vec{x},t), no matter how complicated and rapid the dependence on x\vec{x} and tt, cannot directly produce diffusive mixing but can exponentially enhance the speed with which diffusion can act over the spatial scale covered by the advection of the fluid.

The essential element in an enhanced relaxation by a laminar flow is that the flow be chaotic. Using standard terminology, a flow is deterministic but chaotic when neighboring streamlines have a separation that increases exponentially with time. Articles on the mathematics of deterministic chaos and topological mixing can easily be found on the web, but their importance to this paper is only that such effects are common. As noted in Boozer:rec-phys , although the condition that the flow be chaotic may sound restrictive, it is a non-chaotic natural flow that is essentially impossible to realize. No special effort is required to achieve enhanced mixing by stirring. Every cook knows that stirring enhances the mixing of fluids—no particular pattern of stirring or detailed computations are required.

The terms chaotic and stochastic are sometimes considered synonyms in descriptions of flows or magnetic fields. Here a distinction is made that is consistent with a distinction made in the mathematical literature. Chaos, or more precisely deterministic chaos, when applied to a flow within a bounded region of space, means the streamlines of the flow are predictable but exponentiate apart throughout a finite fraction of that space, the region in which the flow is chaotic. In contrast, a stochastic motion is not deterministic and has a random component on all time and spatial scales. Stochastic flows were used by L. F. Richardson in a paper Richardson:1926 published in 1926 to explain the enhancement of diffusion produced by atmospheric flows. His assumption was that that the velocity of air Δx/Δt\Delta\vec{x}/\Delta t has no well defined limit as Δt\Delta t goes to zero. Weak magnetic field line stochasticity makes field lines indeterminant in a way that is similar to the the effect that Richardson’s assumption has on streamlines of air. The enhancement of reconnection by weak stochasticity of the magnetic field lines was studied by A. Lazarian and E. T. Vishniac Lazarian:1999 in 1999. Here we consider only deterministic flows, which means velocities are well defined. In an ideal magnetic evolution, Equation (1), deterministic flows give deterministic magnetic field lines.

The evolution equation for a magnetic field is also of the advection-diffusion type but more subtly so than the equation for thermal equilibration. The companion paper Boozer:rec-phys derives two salient differences: (1) A flow-enhanced equilbration of the temperature requires a dependence of v\vec{v} on at least two spatial coordinates to be energetically feasible, but flow-enhanced magnetic reconnection requires a dependence of u\vec{u}_{\bot} on all three spatial coordinates for energetic feasibility. (2) In thermal relaxation, |T|\big{|}\vec{\nabla}T\big{|} increases exponentially with time but what one might think would be the analogous quantity for the magnetic field, j=×B/μ0\vec{j}=\vec{\nabla}\times\vec{B}/\mu_{0} does not. In the model developed in this paper, the increase in the current density is limited to being proportional to time. Related limits on the current density are known for resonant perturbations to toroidal plasmas Hahm-Kulsrud and for currents flowing near magnetic field lines that intercept a magnetic null Elder-Boozer .

When the ideal flow velocity of a magnetic field u\vec{u}_{\bot} is chaotic, the ratio Δmax/Δmin\Delta_{max}/\Delta_{min}, the maximum to the minimum separation between two field lines at any particular time, tends to increase exponentially during the evolution when the closest approach of the lines is small, Δmin/a0\Delta_{min}/a\rightarrow 0. Large scale reconnection occurs for magnetic field lines that satisfy ΔminΔd\Delta_{min}\lesssim\Delta_{d} and Δmaxa\Delta_{max}\approx a. As will be seen, a simple laminar flow can be followed as Δmax/Δmin\Delta_{max}/\Delta_{min} increases by more than nine orders of magnitude.

The model that will be used to illustrate features of magnetic reconnection is developed in Section II. In this model, the magnetic field evolution is driven in a bounded plasma in the same way as footpoint motion drives the magnetic field in the solar corona.

Section III will show that when the footpoint velocity is sufficiently small compared to the Alfvén speed in the plasma that important features can be determined without solving the difficult problem of determining B(x,t)\vec{B}(\vec{x},t) throughout the plasma. These features include the exponentiation in time of the separation between neighboring magnetic field lines and the current density j||j_{||} along each magnetic field line, or more precisely Kμ0j||/BK\equiv\mu_{0}j_{||}/B. A bound on the magnitude of KK is obtained, which increases linearly in time. For these features, only the determination of the streamlines of a time-dependent flow in two dimensions is required.

The simplicity of the treatment given in Section III depends on the evolution being slow compared to the Alfvén transit time. This would be violated if the plasma became kink unstable. Section IV shows that the flow example introduced in Section II ensures kink stability.

Section V shows that the rate of production of the magnetic flux associated with a particular magnetic field line by the footpoint motion is the magnetic Reynolds number RmR_{m} times larger than its rate of destruction by resistivity. The sign of the flux differs from field line to field line in kink-stable evolutions, and the rate of flux production and destruction can be brought into balance by reconnection.

Section VI shows that even if the plasma becomes kink unstable, the magnetic helicity contained in the bounded plasma will increase without limit when Rm>>1R_{m}>>1 unless the footpoint motion obeys a constraint. If the plasma were unbounded, an ever increasing helicity would presumptively result in a plasma eruption.

Section VII derives the power that must be supplied to maintain the specified footpoint motion. For a steady-state flow, the required power increases linearly with KK. Section VIII discusses the results and their implications.

II Model of the solar corona

Features of magnetic reconnection will be illustrated using a simplified model of magnetic loops in the solar corona. The evolution of coronal loops is driven by the motion of their foopoints by photospheric flows.

The description of the model has three parts: (1) Section II.1 defines a spatially bounded perfectly-conducting cylinder that encloses the loop as well as the flow vt\vec{v}_{t} in the top surface of the cylinder, which represents the photospheric motion. (2) Section II.2 explains the specific form chosen for vt\vec{v}_{t}. (3) Section II.3 explains how the chaotic properties of vt\vec{v}_{t} are quantified.

The numerical results for the streamlines of the flow vt\vec{v}_{t} are given in Section II.4.

Refer to caption

Figure 1: A perfectly conducting cylinder of radius aa and height LL in zz encloses an ideal pressureless plasma. All the sides of the cylinder are fixed except the top, which flows with a specified velocity vt\vec{v}_{t}. Initially, B=B0z^\vec{B}=B_{0}\hat{z}.

II.1 Definition of cylindrical region

The model that will be used to illustrate important features of magnetic reconnection consists of an ideal pressureless plasma enclosed by a perfectly conducting cylinder of radius aa and height LL, Figure 1. An ideal plasma means Equation (1) for an ideal magnetic evolution holds exactly. The conducting surfaces of the cylinder are rigid except for a flowing top-surface, which provides the footpoint motion that drives the evolution.

The flow of the top surface of the cylinder must be divergence free to avoid compressing the initial magnetic field B0=B0z^\vec{B}_{0}=B_{0}\hat{z} with B0B_{0} a constant. The implication is the flow must have the form

vt=z^×ht(x,y,t),\vec{v}_{t}=\hat{z}\times\vec{\nabla}h_{t}(x,y,t), (4)

where the stream function hth_{t} is chosen to represent a photospheric-like motion at one interception of a coronal magnetic loop. For simplicity, the flow at the other photospheric interception is taken to be zero. The effect of the flow is to produce a magnetic field orthogonal to B0\vec{B}_{0}.

The model illustrated in Figure 1 is closely related to the well-known Parker Problem Parker:problem , which was recently reviewed by Pontin and Hornig Review:ParkerProb . It is also a simplified version of reconnection models of the corona published by Boozer Boozer:prevalence and independently by Reid et al Reid:2018 in 2018. Although these two models are similar, the mathematical cause for reconnection emphasized in these papers is different. Reid et al Reid:2018 used an anomalous resistivity to ensure “that resistivity, as opposed to a numerical diffusion, is responsible for any magnetic reconnection.” Boozer’s paper and his more recent work Boozer:ideal-ev ; Boozer:part.acc ; Boozer:null-X ; Boozer:rec-phys emphasized that an imposed chaotic flow vt\vec{v}_{t} would make magnetic reconnection exponentially sensitive to any departures from an ideal evolution—even effects due to numerics rather than physics.

II.2 Choice of vt\vec{v}_{t}

The divergence-free velocity vt\vec{v}_{t} is specified by the stream function ht(x,y,t)h_{t}(x,y,t), Equation (4). Using Cartesian coordinates, the stream function is the Hamiltonian for the streamlines of the flow on the top of the cylinder dx/dt=ht/ydx/dt=-\partial h_{t}/\partial y and dy/dt=ht/xdy/dt=\partial h_{t}/\partial x.

The flow should satisfy certain conditions, and these are easier to impose in cylindrical coordinates r,θ,zr,\theta,z where x=rcosθx=r\cos\theta and y=rsinθy=r\sin\theta. The vertical part of the cylindrical shell is at r=ar=a. The steam function hth_{t} should be chosen so hth_{t}, dr/dtdr/dt, dθ/dtd\theta/dt, and the radial gradient of dθ/dtd\theta/dt are all zero at r=ar=a. This ensures that the streamlines can never strike the r=ar=a boundary of the cylinder, vt=0\vec{v}_{t}=0 at r=ar=a, and extremely large currents do not form in the plasma near r=ar=a.

A form for hth_{t} that satisfies these conditions is

ht(r,θ,t)\displaystyle h_{t}(r,\theta,t) =\displaystyle= h~(x,y,t)(1r2a2)3eλ2r2/a2,\displaystyle\tilde{h}(x,y,t)\left(1-\frac{r^{2}}{a^{2}}\right)^{3}e^{-\lambda^{2}r^{2}/a^{2}}, (5)

where λ2\lambda^{2} is a constant. A large λ2\lambda^{2} restricts the evolution-driven region to be far from the confining cylindrical walls. In the studies reported here, λ2=0\lambda^{2}=0. The stream function hth_{t} is fully specified when h~(x,y,t)\tilde{h}(x,y,t) is given.

Flows that carry footpoints over a scale comparable to aa are the most effective at producing a rapid reconnection Boozer:rec-phys . These are slowly varying terms in xx and yy, which in simple forms are

h~\displaystyle\tilde{h} =\displaystyle= a2τ[c0cos(ω0tτ)+c1xacos(ω1tτ)\displaystyle\frac{a^{2}}{\tau}\Big{[}c_{0}\cos\left(\omega_{0}\frac{t}{\tau}\right)+c_{1}\frac{x}{a}\cos\left(\omega_{1}\frac{t}{\tau}\right) (6)
+c2yasin(ω2tτ)}+c3xya2cos(ω3tτ)],\displaystyle\hskip 0.72229pt+c_{2}\frac{y}{a}\sin\left(\omega_{2}\frac{t}{\tau}\right)\Big{\}}+c_{3}\frac{xy}{a^{2}}\cos\left(\omega_{3}\frac{t}{\tau}\right)\Big{]},\hskip 14.45377pt

where ω0\omega_{0}, ω2\omega_{2}, and ω3\omega_{3} are three frequencies, which are generally incommensurate, and c0c_{0}, c1c_{1}, c2c_{2}, and c3c_{3} are dimensionless amplitudes. The amplitudes and frequencies used in the calculations are c0=0c_{0}=0, c1=c2=c3=1/4c_{1}=c_{2}=c_{3}=1/4, ω1=6π\omega_{1}=6\pi, ω2=4π\omega_{2}=4\pi, and ω3=0\omega_{3}=0.

The c0c_{0} term by itself would produce streamlines that lie on circles, which produce a large parallel current density with a long correlation distance. Such current densities tend to be ideal-kink unstable, Section IV, which would violate the assumption of a slow evolution compared to Alfvénic. As will be shown in Section III the analysis is greatly simplified when the evolution is slow compared to Alfvénic. Consequently, the coefficient c0c_{0} was chosen to be zero. A shown by Reid et al Reid:2020 , the kinks produced by pure circular motions lead to a large scale exponential separation of neighboring magnetic field lines. The only reason for choosing c0=0c_{0}=0 is to be able to study a rigorously correct but simple example.

For determining the chaotic region associated with a particular initial condition x0,y0x_{0},y_{0}, it is advantageous for the frequencies to be commensurate because then a Poincaré plot can be constructed using the time-periodic points. When the frequencies are commensurate, τ\tau is the periodicity or transit time of h~\tilde{h}. The exponentiation, which as explained in Section II.3 is measured by the Frobenius norm, can be calculated whether the frequencies are commensurate or not.

The actual value of τ\tau is arbitrary. It is effectively the unit of time. Similarly the distance LL can be chosen arbitrarily. Only dimensionless ratios are important. The Alfvén speed VAV_{A} only enters through the constraint, L/τ<<VAL/\tau<<V_{A}.

A circular conducting-cylinder is easier to discuss and does not complicate the computations of this paper, but a cylinder with a square cross section simplifies more complete simulations. In a square cylinder, the factors of 1r2/a21-r^{2}/a^{2} in Equation (5) are replaced by (1x2/a2)(1y2/a2)(1-x^{2}/a^{2})(1-y^{2}/a^{2}).

II.3 Quantification of chaos

The stream function hth_{t} is the Hamiltonian for the stream lines of the flow vt\vec{v}_{t} with (x,y)(x,y) the canonical variables. The most important question about a specific flow is whether its streamlines are chaotic.

Neighboring chaotic streamlines have a separation δ\vec{\delta} that depends exponentially on time. Neighboring means separated by an infinitesimal distance. The exponentiation of neighboring streamlines can be defined streamline by streamline. The t=0t=0 location of a streamline (x0,y0)(x_{0},y_{0}) defines that streamline, which at time tt has the position x=x(x0,y0,t)x^+y(x0,y0,t)y^\vec{x}=x(x_{0},y_{0},t)\hat{x}+y(x_{0},y_{0},t)\hat{y}, where (x/t)x0y0dx/dt=vt(\partial\vec{x}/\partial t)_{x_{0}y_{0}}\equiv d\vec{x}/dt=\vec{v}_{t}.

To determine not only the streamline that is at (x0,y0)(x_{0},y_{0}) at t=0t=0 but also all the streamlines in its neighborhood two vector equations should be integrated simultaneously:

dxdt\displaystyle\frac{d\vec{x}}{dt} =\displaystyle= vt, where\displaystyle\vec{v_{t}},\mbox{ where } (7)
vt\displaystyle\vec{v_{t}} =\displaystyle= htyx^+htxy^, and\displaystyle-\frac{\partial h_{t}}{\partial y}\hat{x}+\frac{\partial h_{t}}{\partial x}\hat{y},\mbox{ and } (8)
dδdt\displaystyle\frac{d\vec{\delta}}{dt} =\displaystyle= δvt.\displaystyle\vec{\delta}\cdot\vec{\nabla}\vec{v_{t}}. (9)

Equation (7) is to be solved with the initial condition x(0)=x0x^+y0y^\vec{x}(0)=x_{0}\hat{x}+y_{0}\hat{y}, so solving that equation means solving two coupled equations, one for dx/dtdx/dt and one for dy/dtdy/dt.

Equation (9) for δ\vec{\delta} is obtained from the equation for neighboring magnetic field lines, which solves the exact equation d(x+δ)/dt=vt(x+δ,t)d(\vec{x}+\vec{\delta})/dt=\vec{v}_{t}(\vec{x}+\vec{\delta},t), by taking the limit as |δ|0\big{|}\vec{\delta}\big{|}\rightarrow 0. Equation (9) should be solved for two different initial conditions. The first solve is for δx=δxxx^+δxyy^\vec{\delta}_{x}=\delta_{xx}\hat{x}+\delta_{xy}\hat{y} with the initial condition δxx=1\delta_{xx}=1 and δxy=0\delta_{xy}=0. The second solve is for δy=δyxx^+δyyy^\vec{\delta}_{y}=\delta_{yx}\hat{x}+\delta_{yy}\hat{y} with the initial condition δyx=0\delta_{yx}=0 and δyy=1\delta_{yy}=1. Since Equation (9) for the evolution of the separation δ\vec{\delta} is linear, the initial separation can be taken to be unity without loss of generality.

The Jacobian matrix for the starting point (x0,y0)(x_{0},y_{0}), which is defined by

xx0\displaystyle\frac{\partial\vec{x}}{\partial\vec{x}_{0}} \displaystyle\equiv (xx0xy0yx0yy0) is then\displaystyle\left(\begin{array}[]{cc}\frac{\partial x}{\partial x_{0}}&\frac{\partial x}{\partial y_{0}}\\ \frac{\partial y}{\partial x_{0}}&\frac{\partial y}{\partial y_{0}}\end{array}\right)\mbox{ is then } (12)
=\displaystyle= (δxxδxyδyxδyy).\displaystyle\left(\begin{array}[]{cc}\delta_{xx}&\delta_{xy}\\ \ \delta_{yx}&\delta_{yy}\end{array}\right). (15)

The determinant of the Jacobian matrix, δxxδyyδyxδxy\delta_{xx}\delta_{yy}-\delta_{yx}\delta_{xy}, called the Jacobian, would be unity if there were no numerical errors. This follows from Liouville’s theorem of Hamiltonian mechanics.

The Frobenius norm of a matrix is the square root of the sum of the squares of the matrix elements and is also equal to the square root of the sum of the squares of the singular values of a Singular Value Decomposition (SVD) of the matrix. The Jacobian of a matrix is the product of its singular values, and a 2×22\times 2 matrix has two singular values Λu\Lambda_{u} and Λs\Lambda_{s}; by definition ΛuΛs\Lambda_{u}\geq\Lambda_{s}.

Consequently, the Frobenius norm of the Jacobian matrix, x/x0\|\partial\vec{x}/\partial\vec{x}_{0}\|, gives the large singular value, Λu\Lambda_{u}, of a Singular Value Decomposition (SVD) of the matrix x/x0\partial\vec{x}/\partial\vec{x}_{0}:

xx0\displaystyle\left\|\frac{\partial\vec{x}}{\partial\vec{x}_{0}}\right\| \displaystyle\equiv δxx2+δxy2+δyx2+δyy2;\displaystyle\sqrt{\delta_{xx}^{2}+\delta_{xy}^{2}+\delta_{yx}^{2}+\delta_{yy}^{2}}; (16)
=\displaystyle= Λu2+1/Λu2\displaystyle\sqrt{\Lambda_{u}^{2}+1/\Lambda_{u}^{2}}

since Λs=1/Λu\Lambda_{s}=1/\Lambda_{u}.

When the flow is chaotic, neighboring streamlines separate exponentially, and Λu\Lambda_{u} becomes exponentially large, which means Λu\Lambda_{u} is essentially equal to the Frobenius norm of the Jacobian matrix, x/x0\|\partial\vec{x}/\partial\vec{x}_{0}\|. A full SVD analysis gives additional information, the directions in both x0,y0x_{0},y_{0} space and in x,yx,y space in which trajectories exponentiate apart and exponentiate together.

The numerical accuracy of the calculations, which are based on Runge-Kutta integrations, can be checked not only by the deviation of the Jacobian from unity, but also by simultaneously integrating one additional equation, dht/dt=ht/tdh_{t}/dt=\partial h_{t}/\partial t, and finding the deviation of hth_{t} resulting from the integration from the actual hth_{t}.

The Frobenius norm of the Jacobian matrix is used here to define the magnitude of the exponentiation. This norm involves a sum of positive numbers and is less numerically demanding than calculating the SVD or the Jacobian, which is the difference between two numbers, each of order the Frobenius norm squared. The largest Frobenius norm in this paper is approximately 10810^{8}. The Jacobian can be the difference between two terms each of order 101610^{16}. The maximum error in the Jacobian is 15%. A more representative number is the standard deviation of the Jacobian from unity, which is 1.1%.

II.4 Numerical results for the streamlines of vt\vec{v}_{t}

Refer to caption

Figure 2: Streamline properties are illustrated for the stream function of Equation (6) with c0=0c_{0}=0, c1=c2=c3=1/4c_{1}=c_{2}=c_{3}=1/4, ω1=6π\omega_{1}=6\pi, ω2=4π\omega_{2}=4\pi, ω3=0\omega_{3}=0, and λ2=0\lambda^{2}=0. Figure 2a is plot of a hundred streamlines started on the perimeter of the small black circle, which has a radius of a/100a/100 and is centered at an arbitrary point, x/a=0.17x/a=0.17 and y/a=0.45y/a=-0.45. The red dots are the locations of the hundred streamlines after nine transits. The locations are widely scattered within the region r<ar<a in which the derivatives of the stream function are non-zero. Figure 2b shows the evolution of the Frobenius norm, Equation (16) for these hundred streamlines. The Frobenius norm is a precise measure of the separation of neighboring trajectories. As expected in a chaotic flow, the Frobenius norm tends to increase exponentially with time. Figure 2c shows the evolution of the force-free current Kμ0j||/BK\equiv\mu_{0}j_{||}/B times the length of the cylinder LL given by Equation (41) for each of the hundred field lines that initially intercepted the top surface on the perimeter of the black circle.

Refer to caption

Figure 3: Data was gathered from a thousand starting points on the perimeter of a circle of radius δ0\delta_{0}. Figure 3a shows the frequency of occurrence of different values of Frobenius norm when δ0=a/100\delta_{0}=a/100. The same value as in Figure 2. Figure 3b shows the frequency of occurrence of different values of the current KK, positive and negative, when δ0=a/100\delta_{0}=a/100. Figure 3c shows the logarithmic scaling of the number of transits required for points started on the perimeter of a circle of radius δ0\delta_{0} to reach the scale 3a/43a/4. Figures 3a and 3b used data from ten transits.

Refer to caption

Figure 4: Ten thousand starting points were uniformly spread over r<ar<a. Figure 4a is the frequency with which streamlines have various values of the Frobenius norm. Figure 4b is the distribution of the current, both positive and negative, relative to the maximum |KL|=41.8\big{|}KL\big{|}=41.8. Figure 4c relates the median value of the natural logarithm of the Frobenius norm (solid line), the median value the positive KLKL magnitudes, and the median value of the negative KLKL magnitudes.

Figure 2a illustrates the effect of a simple vt\vec{v}_{t} that is chaotic, as almost all choices of hth_{t} that have non-trivial xx, yy, and tt dependencies are. One hundred streamlines are started on the perimeter of the small black circle, but as the red dots illustrate, these hundred streamlines spread over most of the region r<ar<a after only five periods τ\tau of the flow. The perimeter of the small circle defines a tube in time with fixed area. The cross section of this tube becomes convoluted in the extreme; all the red dots must lie on the perimeter of the tube. Figure 2b illustrates the evolution of the Frobenius norms of these hundred streamlines. Figure 3a gives the frequency distribution of the Frobenius norms for a thousand streamlines started on the small black circle. The smallest Frobenius norm for these thousand streamlines is 148 and the largest is 3×1073\times 10^{7}. The frequency distribution is peaked near their geometric mean (148)×(3×107)6.7×104\sqrt{(148)\times(3\times 10^{7})}\approx 6.7\times 10^{4}.

A thousand streamlines were launched from a circle of radius δ0\delta_{0} and the number of transits, t/τt/\tau, was recorded for the streamlines to become separated by a distance 3a/43a/4. Figure 3c shows that the required number of transits is proportional to the logarithm of the chosen δ0\delta_{0}.

The fraction of the total r<ar<a region that has chaotic streamlines is assessed by starting a thousand streamlines at uniformly spread points over the r<ar<a region and following them for ten transits. The frequency with which various Frobenius norms arose is shown in Figure 4a. The smallest Frobenius norm is 2\sqrt{2}, which is the smallest value that is mathematically allowed. The peak at small Frobenius norms implies that non-chaotic regions exist, which are separated by Lagrangian coherent structures from the chaotic regions Borgogno:2011 . In non-chaotic regions, the separation between neighboring streamlines typically increases in proportion to time. The frequency distribution of Frobenius norms is peaked near the geometric mean, 1.1×1041.1\times 10^{4}, of the largest, 8×1078\times 10^{7}, and the smallest Frobenius norm.

The fractional distribution of exponentiations in the separation of magnetic field lines was calculated in 2014 for a related problem by Huang et al Huang:2014 with far fewer e-folds; their distribution was also peaked.

III Evolution equations

The magnetic field evolution in the model described in Section II becomes remarkably simple when the height of the cylinder is far greater than its radius, L/aL/a\rightarrow\infty and the time scale of ideal evolution is very long compared the the Alfvén transit time. For simplicity the plasma pressure is assumed to be zero. The derivation is essentially a simplified version of the derivation of Reduced MHD Kadomtsev ; Strauss .

This section consists of four subsections. The first two subsections, Section III.1 and III.2, derive two differential equations that relate the Lagrangian derivatives of time and distance along the magnetic field lines \ell of the distribution of parallel current Kμ0j||/BK\equiv\mu_{0}j_{||}/B and the vorticity Ωz^×u\Omega\equiv\hat{z}\cdot\vec{\nabla}\times\vec{u}_{\bot} of the magnetic field line flow. The third subsection, Section III.3, discusses the implications of these two equations and is the most important part of the paper. The fourth subsection, Section III.4, gives a heuristic argument that the magnitude of the current distribution KK should typically scale as the strength of the exponential separation of the magnetic field lines, their Frobenius norm. This relation is shown to hold accurately for the ensemble of many field lines but not for each line. Indeed, the correlation between large values of KK and large values of the Frobenius norm is found to be remarkably weak.

III.1 Simplification of the magnetic evolution

When L/aL/a\rightarrow\infty, the magnetic field consists of a constant field B0B_{0} in the z^\hat{z} direction plus an orthogonal field produced by the velocity in the top surface. B=0\vec{\nabla}\cdot\vec{B}=0 implies the magnetic field and the vector potential have the forms

B\displaystyle\vec{B} =\displaystyle= B0(z^+z^×H);\displaystyle B_{0}\Big{(}\hat{z}+\hat{z}\times\vec{\nabla}H\Big{)}; (17)
A\displaystyle\vec{A} =\displaystyle= B0(z^×x2Hz^),\displaystyle B_{0}\Big{(}\frac{\hat{z}\times\vec{x}}{2}-H\hat{z}\Big{)}, (18)

where B=×A\vec{B}=\vec{\nabla}\times\vec{A}.

Equation (1) for the ideal evolution of the magnetic field implies the vector potential evolves as

At=u×BB0h,\displaystyle\frac{\partial\vec{A}}{\partial t}=\vec{u}_{\bot}\times\vec{B}-B_{0}\vec{\nabla}h, (19)

where B0=z^BB_{0}=\hat{z}\cdot\vec{B} is a constant and hh represents the freedom of gauge. The constraint that the z^\hat{z}-directed field does not change is

z^Bt=(z^×At)=0.\displaystyle\hat{z}\cdot\frac{\partial\vec{B}}{\partial t}=-\vec{\nabla}\cdot\Big{(}\hat{z}\times\frac{\partial\vec{A}}{\partial t}\Big{)}=0. (20)

A vector identity implies z^×(u×B)=(z^B)u(z^u)B\hat{z}\times(\vec{u}_{\bot}\times\vec{B})=(\hat{z}\cdot\vec{B})\vec{u}_{\bot}-(\hat{z}\cdot\vec{u}_{\bot})\vec{B}. Consequently, the constraint on the constancy of B0=z^BB_{0}=\hat{z}\cdot\vec{B} is that the velocity of the magnetic field lines have the form

u\displaystyle\vec{u}_{\bot} =\displaystyle= z^×h.\displaystyle\hat{z}\times\vec{\nabla}h. (21)

The curl of the magnetic field of Equation (17) and the curl of the magnetic field line velocity u\vec{u}_{\bot} of Equation (21) give the current and the vorticity along B\vec{B}:

2H=K, where Kμ0j||B, and\displaystyle\nabla_{\bot}^{2}H=K,\mbox{ where }K\equiv\frac{\mu_{0}j_{||}}{B},\mbox{ and } (22)
2h=Ω, where Ωz^×u.\displaystyle\nabla_{\bot}^{2}h=\Omega,\mbox{ where }\Omega\equiv\hat{z}\cdot\vec{\nabla}\times\vec{u}_{\bot}. (23)

Equations (18) and (19) give two expressions for B(A/t)x\vec{B}\cdot(\partial\vec{A}/\partial t)_{\vec{x}}, which can be equated to obtain

Ht\displaystyle\frac{\partial H}{\partial t} =\displaystyle= BhB0\displaystyle\frac{\vec{B}\cdot\vec{\nabla}h}{B_{0}} (24)
=\displaystyle= hz+(z^×H).\displaystyle\frac{\partial h}{\partial z}+(\hat{z}\times\vec{\nabla}_{\bot}H)\cdot\vec{\nabla}_{\bot}. (25)

Since the magnetic field line velocity u=z^×h\vec{u}_{\bot}=\hat{z}\times\vec{\nabla}h,

Ht+uH=hz.\frac{\partial H}{\partial t}+\vec{u}_{\bot}\cdot\vec{\nabla}H=\frac{\partial h}{\partial z}. (26)

This equation can be written in Lagrangian coordinates in which ordinary Cartesian coordinates x=xx^+yy^+zz^\vec{x}=x\hat{x}+y\hat{y}+z\hat{z} are given as x(xL,t)\vec{x}(\vec{x}_{L},t) with

(xt)L\displaystyle\left(\frac{\partial\vec{x}}{\partial t}\right)_{L} =\displaystyle= u, so\displaystyle\vec{u}_{\bot},\mbox{ so } (27)
(Ht)L\displaystyle\left(\frac{\partial H}{\partial t}\right)_{L} =\displaystyle= (Ht)x+Hx(xt)L\displaystyle\left(\frac{\partial H}{\partial t}\right)_{\vec{x}}+\frac{\partial H}{\partial\vec{x}}\cdot\left(\frac{\partial\vec{x}}{\partial t}\right)_{L} (28)
=\displaystyle= Ht+uH. Consequently \displaystyle\frac{\partial H}{\partial t}+\vec{u}_{\bot}\cdot\vec{\nabla}H.\mbox{ Consequently }\hskip 14.45377pt (29)
(Ht)L\displaystyle\left(\frac{\partial H}{\partial t}\right)_{L} =\displaystyle= hz\displaystyle\frac{\partial h}{\partial z} (30)

using Equation (26).

Although the form is more complicated involving the metric tensor, Laplacians can be calculated in Lagrangian coordinates, and the relations 2H=K\nabla_{\bot}^{2}H=K Equation (22), and 2h=Ω\nabla_{\bot}^{2}h=\Omega, Equation (23), remain valid. Applying 2\nabla_{\bot}^{2} to both sides of Equation (30),

(Kt)L=(Ω)L,\left(\frac{\partial K}{\partial t}\right)_{L}=\left(\frac{\partial\Omega}{\partial\ell}\right)_{L}, (31)

where the subscript LL on the partial derivatives implies the use of Lagrangian coordinates, which means x0x_{0} and y0y_{0} are held constant;

(Kt)LKt+uK, and\displaystyle\left(\frac{\partial K}{\partial t}\right)_{L}\equiv\frac{\partial K}{\partial t}+\vec{u}_{\bot}\cdot\vec{\nabla}K,\mbox{ and } (32)
(Ω)LBΩ.\displaystyle\left(\frac{\partial\Omega}{\partial\ell}\right)_{L}\equiv\vec{B}\cdot\vec{\nabla}\Omega. (33)

The differential distance along a magnetic field line, dd\ell, is equivalent to dzdz with x0x_{0} and y0y_{0} held constant.

The implications of Equation (31) will be found to be extremely profound.

III.2 Constraint of force balance

The Lagrangian time and \ell derivatives of the parallel current distribution Kμ0j||/BK\equiv\mu_{0}j_{||}/B and the vorticity Ω\Omega obey not only Equation (31) but also Equation (37), which is implied by force balance.

For simplicity, the plasma is assumed to have a negligible pressure and a constant density ρ\rho, so force balance is ρ(u/t+uu)=fL\rho(\partial\vec{u}_{\bot}/\partial t+\vec{u}_{\bot}\cdot\vec{\nabla}\vec{u}_{\bot})=\vec{f}_{L}, where fLj×B\vec{f}_{L}\equiv\vec{j}\times\vec{B} is the Lorentz force. The condition j=0\vec{\nabla}\cdot\vec{j}=0 can be written as

BK\displaystyle\vec{B}\cdot\vec{\nabla}K =\displaystyle= B×μ0fLB2, and\displaystyle\vec{B}\cdot\vec{\nabla}\times\frac{\mu_{0}\vec{f}_{L}}{B^{2}},\mbox{ and } (34)
×(uu)\displaystyle\vec{\nabla}\times(\vec{u}_{\bot}\cdot\vec{\nabla}\vec{u}_{\bot}) =\displaystyle= ×(Ω×u)\displaystyle\vec{\nabla}\times(\vec{\Omega}\times\vec{u}_{\bot}) (35)
=\displaystyle= uΩΩu,\displaystyle\vec{u}_{\bot}\cdot\vec{\nabla}\vec{\Omega}-\vec{\Omega}\cdot\vec{\nabla}\vec{u}_{\bot}, (36)

where Ω=Ωz^\vec{\Omega}=\Omega\hat{z}. The z^\hat{z} component of the curl of the force balance equation gives

(Ωt)L=VA2(K)L, where VA2B02μ0ρ\displaystyle\left(\frac{\partial\Omega}{\partial t}\right)_{L}=V_{A}^{2}\left(\frac{\partial K}{\partial\ell}\right)_{L},\mbox{ where }V_{A}^{2}\equiv\frac{B_{0}^{2}}{\mu_{0}\rho} (37)

is the Alfvén speed.

III.3 Implications of the KK and Ω\Omega equations

Equations (31) and (37) together with the mixed partials theorem applied to either Ω\Omega or KK imply both Ω\Omega and KK obey the equation for shear Alfvén waves, (2K/t2)L=VA2(2K/2)L(\partial^{2}K/\partial t^{2})_{L}=V_{A}^{2}(\partial^{2}K/\partial\ell^{2})_{L}. Any variation in KK along the magnetic field lines relaxes by Alfvén waves. Reconnection or ideal kink-instabilities will generally drive Alfvén waves. The inclusion of resistivity or viscosity causes these waves to diffuse across the magnetic field lines and produces wave decay Boozer:j-|| . In a completely ideal theory, the energy that goes into Alfvén waves will bounce back and forth forever, but they can be damped without directly affecting reconnection by adding viscosity or a drag-force to the force equation.

Equations (31) and (37) imply that during any period in which the evolution is slow compared to the Alfvén transit time L/VAL/V_{A} that

(K)L=0, and\displaystyle\left(\frac{\partial K}{\partial\ell}\right)_{L}=0,\mbox{ and } (38)
(2Ω2)L=0, so\displaystyle\left(\frac{\partial^{2}\Omega}{\partial\ell^{2}}\right)_{L}=0,\mbox{ so } (39)
Ω=Ωt(x0,y0,t)L, and\displaystyle\Omega=\Omega_{t}(x_{0},y_{0},t)\frac{\ell}{L},\mbox{ and } (40)
(Kt)L=Ωt(x0,y0,t)L, where\displaystyle\left(\frac{\partial K}{\partial t}\right)_{L}=\frac{\Omega_{t}(x_{0},y_{0},t)}{L},\mbox{ where } (41)
Ωtz^×vt.\displaystyle\Omega_{t}\equiv\hat{z}\cdot\vec{\nabla}\times\vec{v}_{t}. (42)

The flow of the top perfectly-conducting surface is specified, and Ωt(x0,y0,t)\Omega_{t}(x_{0},y_{0},t) is obtained from that specified flow alone. For an example of a calculation of the evolution of KLKL, see Figure 2c.

Equations (22) and (41) provide a Poisson equation, 2H=K\nabla_{\bot}^{2}H=K, for HH and an expression for KK, which can be solved for each value of zz and tt. The boundary condition is that the component of H\vec{\nabla}_{\bot}H that is tangential to the wall must vanish, otherwise the magnetic field would penetrate the perfectly conducting wall.

For any physically reasonable flow, Ωt(x0,y0,t)\Omega_{t}(x_{0},y_{0},t) is bounded, |Ωt|Ωmax\big{|}\Omega_{t}\big{|}\leq\Omega_{max}, which can be easily calculated analytically for any analytic ht(x,y,t)h_{t}(x,y,t). An extremely important result is that the maximum current density along the magnetic field j||j_{||} satisfies

KmaxΩmaxLt.\displaystyle K_{max}\leq\frac{\Omega_{max}}{L}t. (43)

The fraction of the values of KK that have a particular value is illustrated for magnetic field lines started on a small circle of radius a/100a/100 in Figure 3b and for lines started uniformly over the full region, r<ar<a in Figure 4b. For the small circle, the current KK is more likely to be negative than positive, which is also clearly illustrated in Figure 2c, but nonetheless KK’s of both signs are present. Over the full region, frequency distribution of currents, Figure 4b, is essentially symmetric between the negative and the positive values, and the most probable KK is essentially zero. The absence of smoothness in the current distribution KK, even in a small region, Figure 2c, due to a smooth flow may be surprising to some. An implication is that currents in the corona must be extremely complicated with a short correlation distance across the magnetic field lines.

When the specified flow in the top surface is chaotic, the spatial derivatives of KK will tend to become exponentially large in some directions and exponentially small in others, but the current density itself is strictly bounded by a linear increase in time. In other words, the current density within the plasma lies in ribbons with a decreasing thickness in one direction across B\vec{B}, an increasing width in the other direction across B\vec{B}, and a constant amplitude along B\vec{B}. This thinning with increasing width is illustrated by the top row of Figure 5.

The anisotropy of the spatial derivatives of KK follows from the exponentially large anisotropy of the spatial derivatives of Ωt(x0,y0,t)\Omega_{t}(x_{0},y_{0},t) in x0,y0x_{0},y_{0} space. For simple stream functions ht(x,y,t)h_{t}(x,y,t), such as those defined by Equation (6), Ωt(x,y,t)=2ht\Omega_{t}(x,y,t)=\nabla^{2}h_{t} has a simple and smooth variation in x,yx,y coordinates. But, the streamlines, x(x0,y0,t)x(x_{0},y_{0},t) and y(x0,y0,t)y(x_{0},y_{0},t), of a two-dimensional, divergence-free chaotic flow separate exponentially in time in one direction, which implies they must exponentially converge in the other. For a divergence-free flow, the two singular values of the Jacobian matrix x/x0\partial\vec{x}/\partial\vec{x}_{0} must be inverses of each other. The spatial derivatives of Ωt(x0,y0,t)\Omega_{t}(x_{0},y_{0},t) in the converging direction become exponentially large and those in the diverging direction become exponentially small.

Employing Equation (41), Ωt(x0,y0,t)\Omega_{t}(x_{0},y_{0},t) determines the distribution of the parallel current density K(x0,y0,z,t)K(x_{0},y_{0},z,t). Consequently, the streamlines of vt\vec{v}_{t} determine the properties of K(x0,y0,z,t)K(x_{0},y_{0},z,t) throughout the plasma. In the limit VA>>L/τV_{A}>>L/\tau, the Lorentz force and, therefore, jj_{\bot} are negligible. Figure 2c illustrates how a hundred magnetic field lines that initially had nearby x0,y0x_{0},y_{0} locations develop a large variation in KLKL. As can be seen in Figure 5, there is only a weak correlation between regions where |K|\big{|}K\big{|} is large and where the Frobenius norm is large. This is consistent with the results of Reid et al Reid:2020 that the quasi-squashing factor, which is determined by the Frobenius norm, Equation (16), has a little correlation with a large ηj||𝑑=B0(η/μ0)KL\int\eta j_{||}d\ell=B_{0}(\eta/\mu_{0})KL.

Magnetic field lines that have distant intersection points with the bottom surface of the cylinder, x0,y0x_{0},y_{0} and x0,y0x^{\prime}_{0},y^{\prime}_{0} can interchange their intersections on the top surface if anywhere along their trajectories they are sufficiently close, <Δd<\Delta_{d}, to be indistinguishable. This means they come closer than c/ωpec/\omega_{pe} or the distance through which they resistively diffuse, (η/μ0)t\sqrt{(\eta/\mu_{0})t}.

III.4 Required current density

The current density required for a large exponentiation is relatively small Boozer:B-line.sep . The minimum number of exponentiations is given by the properties of hth_{t}, but more are possible. The separation Δ=Δxx^+Δyy^\vec{\Delta}=\Delta_{x}\hat{x}+\Delta_{y}\hat{y} between two neighboring magnetic field lines obeys dΔ/dz=Δb^d\vec{\Delta}/dz=\vec{\Delta}\cdot\vec{\nabla}\hat{b}, where b^=z^+z^×H\hat{b}=\hat{z}+\hat{z}\times\vec{\nabla}H. That is,

dΔxdz\displaystyle\frac{d\Delta_{x}}{dz} =\displaystyle= 2HxyΔx2Hy2Δy\displaystyle-\frac{\partial^{2}H}{\partial x\partial y}\Delta_{x}-\frac{\partial^{2}H}{\partial y^{2}}\Delta_{y} (44)
dΔydz\displaystyle\frac{d\Delta_{y}}{dz} =\displaystyle= 2Hx2Δx+2HxyΔy.\displaystyle\frac{\partial^{2}H}{\partial x^{2}}\Delta_{x}+\frac{\partial^{2}H}{\partial x\partial y}\Delta_{y}. (45)

An exact answer for the separation requires a solution of the equation 2H=K\nabla_{\bot}^{2}H=K. But, the typical magnitude of the second derivatives of HH, which appear in Equations (44) and (45) is KK, which suggests that the number of e-folds is typically of order KLKL. Figure 2 illustrates this scaling, and Figure 4c shows the accuracy with which the scaling holds. Although the scaling holds for the ensemble averages, it does not hold magnetic field line by field line, which is another way of saying that the correlation between the magnitude of the Frobenius norm and the current density is weak, Figure 5. A current along a magnetic field line affects not only the Hamiltonian and its derivatives on that line but elsewhere as well.

Refer to caption

Figure 5: Ten thousand starting points were uniformly spread over r<ar<a. The top row is the locations of the five hundred points that had the largest current density |K|\big{|}K\big{|} after one, three, and five transits. Red implies KK is negative and blue positive. The bottom row is the locations of the five hundred points that had the largest Frobenius norm, which measures the rapidity of streamline separation. The correlation between a large KK and a large Frobenius norm is weak. The regions with a high current density KK tend to become long and thin, but small regions that have a correlated current density do not entirely disappear. Figure 2c also illustrates this.

IV Kink stability

The current flowing along the magnetic field lines causes the lines to twist through an angle Θ=KL/2\Theta=KL/2 from one end of the cylinder to the other. When the twist has a smooth variation with radius, Hood and Priest Hood:kink1979 found the magnetic field becomes unstable to an ideal kink when Θ\Theta is greater than a critical value, which in their calculations lay in the range 2π2\pi to 6π6\pi. Studies of the onset of reconnection in the model of Figure 1 are much simpler when ideal kink instabilities are not an issue.

The largest KLKL values in Figure 2c correspond to Θ8π\Theta\approx 8\pi, but the current KK has an extremely complicated spatial distribution, not only in magnitude but also in sign; spatial averages are far smaller than the maximum value. As will be discussed, the anisotropy of the derivatives of KK across the magnetic field lines and the smallness of the spatial averages of KK makes the system highly stable to kinks.

It is not required that KK, or equivalently the twist Θ\Theta, have small spatial averages when the flow is chaotic. The spatially averaged twist Θ\Theta can be made arbitrarily large by choosing c0c_{0} to be large and ω0\omega_{0} to be either zero or small in Equation (6) for h~\tilde{h}. The choice c0=0c_{0}=0 was made to show that a large average field line twist is not needed to obtain chaos.

Even when c0=0c_{0}=0, the spatial average of KK over small regions can be non-zero. This is illustrated by Figure 2c and by the first row of Figure 5.

When the stream function is chosen so the flow is chaotic but with a large spatially-averaged KK, the resulting magnetic field will generally evolve not only into a kinked but also into an eruptive state. As shown in Section VI, the evolution properties of magnetic helicity imply the spatial and temporal average of hth_{t} must be zero for a non-eruptive steady-state solution for the magnetic field when Rm>>1R_{m}>>1—no matter how spatially concentrated the current may become. Consequently, non-eruptive chaotic models tend to have spatially complicated distributions of Θ\Theta in which Θ\Theta has both signs and a near-zero spatial average as in the vt\vec{v}_{t} example used to construct Figure 2.

As discussed in Section V.B.1 of Boozer:RMP , the stability of force-free equilibria can be determined using the perturbed equilibrium equation ×δB=(μ0δj||/B)B\vec{\nabla}\times\delta\vec{B}=(\mu_{0}\delta j_{||}/B)\vec{B}, where δB=×(δA||z^)\delta\vec{B}=\vec{\nabla}\times(\delta A_{||}\hat{z}). The perturbed parallel current is determined by the constancy of Kμ0j||/BK\equiv\mu_{0}j_{||}/B along magnetic field lines, which in linear order in the perturbation implies BδK+δBK=0\vec{B}\cdot\vec{\nabla}\delta K+\delta\vec{B}\cdot\vec{\nabla}K=0. Stability is determined by whether it takes positive or negative energy to drive a perturbation that obeys the equations

2δA||\displaystyle\nabla_{\bot}^{2}\delta A_{||} =\displaystyle= δKB0;\displaystyle-\delta KB_{0}; (46)
B0(δK)x0y0\displaystyle B_{0}\left(\frac{\partial\delta K}{\partial\ell}\right)_{x_{0}y_{0}} =\displaystyle= z^(A||×K).\displaystyle\hat{z}\cdot\left(\vec{\nabla}_{\bot}A_{||}\times\vec{\nabla}_{\bot}K\right).\hskip 14.45377pt (47)

The system is at marginal stability when δK\delta K is just strong enough to produce a solution δA||\delta A_{||} that fits within the perfectly conducting cylindrical walls. The implication is that when Equation (46) is multiplied by δA||\delta A_{||}, then at marginal stability

{(δA||)2δKδA||B0}d3x=0.\displaystyle\int\left\{\left(\vec{\nabla}_{\bot}\delta A_{||}\right)^{2}-\delta K\delta A_{||}B_{0}\right\}d^{3}x=0. (48)

When δK\delta K has only a rapid spatial variation, as it does when KK does, then δA||\delta A_{||} must also have a rapid variation to avoid a self-cancelation of the destabilizing term δKδA||\delta K\delta A_{||} in Equation (48). Equation (47) for δK\delta K involves two spatial derivatives across B\vec{B} and one might think they could be large and balance the stabilizing effect of the two spatial derivatives in (δA||)2\left(\vec{\nabla}_{\bot}\delta A_{||}\right)^{2}, but that is not the case. The two spatial derivatives across B\vec{B} in the equation for δK\delta K are orthogonal and the spatial derivatives of KK in the two directions across the magnetic field lines tend to be of exponentially different magnitudes, so the large term in K\vec{\nabla}_{\bot}K forces a large spatial derivative in δA||\delta A_{||}, which quadratically enhances (δA||)2\left(\vec{\nabla}_{\bot}\delta A_{||}\right)^{2} but only linearly enhances δK\delta K.

V Magnetic flux

The change in the magnetic flux associated with a particular magnetic field line ψ(x0,y0,t)\psi(x_{0},y_{0},t) is the integral from one perfectly conducting surface to the other, ψ/t=E𝑑\partial\psi/\partial t=-\int\vec{E}\cdot d\vec{\ell}. When the cylindrical conductor is stationary and the plasma is resistive, the flux decays as ψ/t=ηj𝑑\partial\psi/\partial t=-\int\eta\vec{j}\cdot d\vec{\ell}.

As was shown in the derivation of Equation (19) for A/dt\partial\vec{A}/dt, the effective inductive electric field along the magnetic field is B(A/t)=B(B0h)-\vec{B}\cdot(\partial A/\partial t)=\vec{B}\cdot(B_{0}\vec{\nabla}h), which gives a change in the flux, ψ/t=E𝑑\partial\psi/\partial t=-\int\vec{E}\cdot d\vec{\ell}, or

ψt=B0ht(x0,y0,t)\frac{\partial\psi}{\partial t}=-B_{0}h_{t}(x_{0},y_{0},t) (49)

since the at the top of the cylinder h=hth=h_{t}.

The appearance of B0htB_{0}h_{t} in the electric-field integral can be understood using the expression E=vt×B0\vec{E}=-\vec{v}_{t}\times\vec{B}_{0} for the electric field in the flowing conductor when observed from a stationary frame of reference. The velocity is vt=z^×ht\vec{v}_{t}=\hat{z}\times\vec{\nabla}h_{t} and B0=B0z^\vec{B}_{0}=B_{0}\hat{z}, so E=B0ht\vec{E}=-B_{0}\vec{\nabla}_{\bot}h_{t}. The electromotive force from the intersection point of the field line to a stationary point, where ht=0h_{t}=0, is B0htB_{0}h_{t}.

The rate at which plasma resistivity destroys magnetic flux is η=ηj||𝑑\mathcal{E}_{\eta}=\int\eta j_{||}d\ell. Since K=μ0j||/BK=\mu_{0}j_{||}/B is constant along a magnetic field line, η=B0(η/μ0)K(x0,y0,t)L\mathcal{E}_{\eta}=B_{0}(\eta/\mu_{0})K(x_{0},y_{0},t)L. Equation (41) implies η/t=B0(η/μ0)Ωt(x0,y0,t)\partial\mathcal{E}_{\eta}/\partial t=-B_{0}(\eta/\mu_{0})\Omega_{t}(x_{0},y_{0},t). Since Ωt=2ht\Omega_{t}=\nabla^{2}h_{t}, the ratio of flux creation to flux destruction is

|2ht/t(η/μ0)2ht|Rm1012\Big{|}\frac{2\partial h_{t}/\partial t}{(\eta/\mu_{0})\nabla^{2}h_{t}}\Big{|}\approx R_{m}\sim 10^{12} (50)

in the solar corona.

VI Magnetic helicity

As will be shown, an argument based on magnetic helicity implies that a long-term relevant solution to the problem outlined in Figure 1 requires the long-term spatial and temporal average of hth_{t} to be zero. When the average of hth_{t} is zero over a chaotic region, the interchange of penetration points implies the poloidal magnetic flux associated with a field line x0,y0x_{0},y_{0} fluctuates but has no systematic increase.

Chaotic streamlines can cause two field lines that penetrate the bottom of the cylinder at two distinct points x0,y0x_{0},y_{0} and x0,y0x^{\prime}_{0},y^{\prime}_{0} to interchange their penetration points through the top plane due to exponentially small non-ideal effects.

Equation (60) for the evolution of the magnetic helicity limits the degree to which a magnetic field driven as in Figure 1 can be simplified by magnetic field lines exchanging connections even if the current density were to obtain arbitrarily high local values by being concentrated in thin sheets. As has been shown, the maximum current density increases only linearly in time, Equation (43), and does not reach the enhancement by a factor of order 1/Rm1/R_{m}, which would be required for the loop voltage to balance the poloidal flux creation, before reconnection has already occurred. But, even if it did the rate of helicity dissipation would not be significantly enhanced, Equation (60). Magnetic turbulence can reduce the magnetic energy, but not the helicity Taylor:1974 ; Berger:1984 as RmR_{m}\rightarrow\infty.

Equation (58) for the rate of helicity increase implies that unless the stream function integrated over each chaotic region, ht𝑑at\int h_{t}da_{t}, has a zero time average, the magnetic helicity can increase without limit. In the model of this paper, the perfectly conducting cylindrical boundary conditions will keep the system confined no matter how strong or contorted the magnetic field may become. But, in a natural system, such as the solar corona, a drive hth_{t} that does not have a zero long-term average will presumably cause the eruption of a magnetic flux tube.

The derivation of the helicity evolution equation starts with the definition of the magnetic helicity enclosed by the cylinder,

𝒦\displaystyle\mathcal{K} \displaystyle\equiv BAd3x.\displaystyle\int\vec{B}\cdot\vec{A}d^{3}x. (51)

Equations (17) and (18) together with x=2\vec{\nabla}\cdot\vec{x}_{\bot}=2 imply BA=B02(2H+(Hx)\vec{B}\cdot\vec{A}=B_{0}^{2}(-2H+\vec{\nabla}_{\bot}\cdot(H\vec{x}_{\bot}). The helicity is then

𝒦=2B02Hd3x.\mathcal{K}=-2B_{0}^{2}\int Hd^{3}x. (52)

The time derivative of the helicity is calculated using

BAt\displaystyle\frac{\partial\vec{B}\cdot\vec{A}}{\partial t} =\displaystyle= BAt+A×At\displaystyle\vec{B}\cdot\frac{\partial\vec{A}}{\partial t}+\vec{A}\cdot\vec{\nabla}\times\frac{\partial\vec{A}}{\partial t} (53)
=\displaystyle= 2BAt(A×At) and\displaystyle 2\vec{B}\cdot\frac{\partial\vec{A}}{\partial t}-\vec{\nabla}\cdot\left(\vec{A}\times\frac{\partial\vec{A}}{\partial t}\right)\mbox{ and }
A×At\displaystyle\vec{A}\times\frac{\partial\vec{A}}{\partial t} =\displaystyle= B02x2Ht, so\displaystyle-B_{0}^{2}\frac{\vec{x}_{\bot}}{2}\frac{\partial H}{\partial t},\mbox{ so } (54)
BAt\displaystyle\frac{\partial\vec{B}\cdot\vec{A}}{\partial t} =\displaystyle= 2BAt+(B02x2Ht)\displaystyle 2\vec{B}\cdot\frac{\partial\vec{A}}{\partial t}+\vec{\nabla}\cdot\left(B_{0}^{2}\frac{\vec{x}_{\bot}}{2}\frac{\partial H}{\partial t}\right) (55)

The side of the cylinder is a rigid perfect conductor, so H/t=0\partial H/\partial t=0 within its sides. Consequently,

d𝒦dt\displaystyle\frac{d\mathcal{K}}{dt} =\displaystyle= 2BAtd3x\displaystyle 2\int\vec{B}\cdot\frac{\partial\vec{A}}{\partial t}d^{3}x (56)
=\displaystyle= 2B0(hB)d3x\displaystyle-2B_{0}\int\vec{\nabla}\cdot(h\vec{B})d^{3}x (57)
=\displaystyle= 2B02ht𝑑at,\displaystyle-2B_{0}^{2}\int h_{t}da_{t}, (58)

using Equation (19) with dat=rdrdθda_{t}=rdrd\theta the area integral over the top of the cylinder. Since the magnetic field lines are tied to the flow of the perfectly conducting top, h=hth=h_{t} within the top surface.

The effect of resistivity on the helicity is obtained by letting A/t=u×BB0hηj\partial\vec{A}/\partial t=\vec{u}_{\bot}\times\vec{B}-B_{0}\vec{\nabla}h-\eta\vec{j}, then Equation (58) implies

d𝒦dt\displaystyle\frac{d\mathcal{K}}{dt} =\displaystyle= B02(ht+ηj||𝑑B0)𝑑at\displaystyle-B_{0}^{2}\int\left(h_{t}+\frac{\int\eta j_{||}d\ell}{B_{0}}\right)da_{t} (59)
=\displaystyle= B02(ht+ημ0K𝑑)𝑑at.\displaystyle-B_{0}^{2}\int\left(h_{t}+\int\frac{\eta}{\mu_{0}}Kd\ell\right)da_{t}. (60)

The effect of resistivity on the helicity evolution is given by the volume-averaged KK, and is therefore unaffected by KK being concentrated. When the evolution is slow compared to the Alfvén transit time, KK is independent of \ell, and the ratio of helicity input to its resistive destruction is

|htημ0K𝑑|\displaystyle\left|\frac{h_{t}}{\int\frac{\eta}{\mu_{0}}Kd\ell}\right| \displaystyle\sim avtημ0Lvtta2L\displaystyle\frac{av_{t}}{\frac{\eta}{\mu_{0}}L\frac{v_{t}t}{a^{2}L}} (61)
\displaystyle\sim (μ0a2η)(vta)a/vtt\displaystyle\left(\frac{\mu_{0}a^{2}}{\eta}\right)\left(\frac{v_{t}}{a}\right)\frac{a/v_{t}}{t}
\displaystyle\sim Rma/vtt,\displaystyle R_{m}\frac{a/v_{t}}{t}, (62)

while Rm1012R_{m}\sim 10^{12} in the corona. To extreme accuracy, resistivity has no effect on the rate of helicity increase in the corona.

VII Power input

The condition j=0\vec{\nabla}\cdot\vec{j}=0 implies that the Lorentz force fLj×B\vec{f}_{L}\equiv\vec{j}\times\vec{B} obeys Equation (34). Using

BK\displaystyle\vec{B}\cdot\vec{\nabla}K =\displaystyle= B0K=Kδ(L)\displaystyle B_{0}\frac{\partial K}{\partial\ell}=-K\delta(\ell-L) (63)
=\displaystyle= B0μ0B02z^×fL,\displaystyle B_{0}\frac{\mu_{0}}{B_{0}^{2}}\hat{z}\cdot\vec{\nabla}\times\vec{f}_{L}, (64)
z^×fL\displaystyle\hat{z}\cdot\vec{\nabla}\times\vec{f}_{L} =\displaystyle= B02μ0Kδ(L).\displaystyle-\frac{B_{0}^{2}}{\mu_{0}}K\delta(\ell-L). (65)

The power required to maintain the flow in the top of the cylinder is

𝒫\displaystyle\mathcal{P} =\displaystyle= vtfLd3x\displaystyle-\int\vec{v}_{t}\cdot\vec{f}_{L}d^{3}x
=\displaystyle= z^(ht×fL)d3x\displaystyle-\int\hat{z}\cdot(\vec{\nabla}h_{t}\times\vec{f}_{L})d^{3}x
=\displaystyle= z^(×(htfL)ht×fL)d3x\displaystyle-\int\hat{z}\cdot\Big{(}\vec{\nabla}\times(h_{t}\vec{f}_{L})-h_{t}\vec{\nabla}\times\vec{f}_{L}\Big{)}d^{3}x
=\displaystyle= z^((htz^×fL)\displaystyle-\int\hat{z}\cdot\Big{(}-\vec{\nabla}(h_{t}\hat{z}\times\vec{f}_{L})
+B02μ0htKδ(L))d3x\displaystyle\hskip 14.45377pt+\frac{B_{0}^{2}}{\mu_{0}}h_{t}K\delta(\ell-L)\Big{)}d^{3}x
=\displaystyle= B02μ0ht(x0,y0,t)K(x0,y0,t)𝑑x0𝑑y0.\displaystyle-\frac{B_{0}^{2}}{\mu_{0}}\int h_{t}(x_{0},y_{0},t)K(x_{0},y_{0},t)dx_{0}dy_{0}.\hskip 7.22743pt (67)

The integrand becomes extremely spatially complicated as time evolves but is not very large.

VIII Discussion

The prevalence of magnetic reconnection in situations in which effects that cause a departure from an ideal evolution are arbitrarily small suggests the cause of reconnection must be within the ideal evolution equation itself. Indeed it is Boozer:ideal-ev .

In an ideal evolution, the magnetic field lines move with the velocity u\vec{u}_{\bot} of Equation (1). As Schindler, Hesse, and Birn Schindler:1988 stated in 1988, resistivity η\eta can only compete with the ideal evolution in a region Δd\Delta_{d} that is sufficiently narrow across the magnetic field that the local magnetic Reynolds number, R=(μ0u/η)Δd1R_{\ell}=(\mu_{0}u_{\bot}/\eta)\Delta_{d}\sim 1. The actual scale of the reconnecting region across the magnetic field, aa, gives the usual magnetic Reynolds number Rm=(μ0u/η)aR_{m}=(\mu_{0}u_{\bot}/\eta)a, which is many orders of magnitude larger than unity—in both natural and laboratory plasmas. In the solar corona Rm1012R_{m}\sim 10^{12}. There are two possibilities for addressing the a/ΔdRma/\Delta_{d}\sim R_{m} problem identified by Schindler et al.

The first possibility, which is the dominant reconnection paradigm and the only one considered by Schindler et al, is that the ideal magnetic evolution creates and maintains layers of intense current density jBrec/μ0Δdj\sim B_{rec}/\mu_{0}\Delta_{d}, where BrecB_{rec} is the reconnecting magnetic field. One problem with this possibility is that in an ideal evolution the current density tends to increase only linearly in time. A linear increase in the current density by a factor RmR_{m} takes far too long to explain many natural phenomena. The growth in current density is found to be linear in time, not only in this paper, Equation (43), but also in ideal flows that are known to create a singular current density as time goes to infinity—flows that have a resonant interaction with a rational magnetic surface in a torus Hahm-Kulsrud ; Boozer-Pomphrey . Although more complicated than a linear increase, a slow increase in current density along the magnetic field lines is also found for lines that pass near a magnetic null Elder-Boozer .

What has been presented in this paper is an example of the second possible explanation for fast magnetic reconnection, but one that has aroused little interest. This explanation is based on the characteristic increase in the ratio of the maximum to the minimum separation, Δmax/Δmin\Delta_{max}/\Delta_{min}, between two magnetic lines that is produced by an ideal flow u\vec{u}_{\bot}. Magnetic field lines are defined at a fixed time, as is Δmax/Δmin\Delta_{max}/\Delta_{min}, but when the two lines are adjacent at some location along their trajectories, Δmin0\Delta_{min}\rightarrow 0, the Δmax/Δmin\Delta_{max}/\Delta_{min} ratio characteristically increases exponentially when the field-line trajectories are calculated at different points in time. Characteristically, the rate of exponentiation is comparable to the evolution time τeva/u\tau_{ev}\equiv a/u_{\bot}. Reconnection must have occurred by the time at which Δmax/ΔminRm\Delta_{max}/\Delta_{min}\sim R_{m} with Δmin=Δd\Delta_{min}=\Delta_{d}, the spatial scale over which field line distinguishability is lost, and Δmaxa\Delta_{max}\sim a, the system scale.

All that is required to produce reconnection by the Parker and Krook definition Parker-Krook:1956 , the “severing and reconnection of lines of force,” is that magnetic field lines become indistinguishable on some spatial scale Δd\Delta_{d} and that the exponentiation of field-line separation magnify the indistinguishability scale to the scale over which reconnection occurs. Indeed, one can model magnetic reconnection using an ideal code Pariat-Antiochos because finite numerical resolution will provide an effective distinguishability scale Δd\Delta_{d}. In many examples of reconnection, the distinguishability scale Δd\Delta_{d} is far smaller than can be directly assessed numerically. It is important to carry out reconnection studies minimizing effects that break the ideal magnetic evolution.

Unlike a localized current density, Δmax/Δmin\Delta_{max}/\Delta_{min} becomes large over extended regions. Nonetheless, the exponentiation, which is quantified by the Frobenius norm, Section II.3, differs from one pair of magnetic field lines to another—even over small regions, Figure 2b. Magnetic field lines that have the largest exponentiations reconnect first, which can break force-balance and cause Alfvénic relaxations when the reconnected region becomes sufficiently large, which is the scale Δr\Delta_{r} at which reconnection becomes important to the system dynamics. The scale Δr\Delta_{r} is important but not determined in this paper. That would require more complete simulations related to those of Reid et al Reid:2018 , and Δr\Delta_{r} is far less well defined than the distinguishability scale Δd\Delta_{d}. Nonetheless, Δr\Delta_{r} is an important concept because the typical size of reconnected regions scale as Δd\Delta_{d} times a factor that depends exponentially on time until the scale of reconnection reaches Δr\Delta_{r}, then reconnection spreads at an Alfvénic rate, which presumably accounts for Parker’s observation Parker:1973 that the speed of reconnection is 0.1VA\approx 0.1V_{A}. Even though an Alfvénic relaxation is consistent with an ideal evolution, it generally causes an increase in Δmax/Δmin\Delta_{max}/\Delta_{min} on the Alfvénic time scale for pairs of magnetic field lines in additional regions of space.

Plasma turbulence can be considered to be third possibility for fast reconnection, but turbulence-enhanced reconnection is in effect given by one of the other two possibilities. Plasma turbulence can mean either micro-turblence or macro-turbulence: (1) Micro-turbulence affects the particle distributions of the Fokker-Planck equation and can cause an enhanced turbulent resistivity ηturb\eta_{turb} for the flow of current along B\vec{B}, which can be much larger than the standard Spitzer value for the parallel resistivity, η\eta. This effect fits into the paradigm of Schindler, Hesse, and Birn Schindler:1988 ; just ηturbj||\eta_{turb}j_{||} must compete with the evolution rather than ηj||\eta j_{||}. (2) Macro-turbulence means the mass flow velocity of the plasma v\vec{v} develops small spatial scales with no direct effect on the particle distribution function. Macro-turbulence is analogous to the turbulent flow of water through a pipe.

In 1953 Dungey Dungey:1953 stated that a change in the linkages of magnetic field lines due to resistivity “is usually very slow in astrophysical systems, but may be increased by turbulent motion in the gas.” Macro-turbulent enhanced reconnection has the same relation to enhanced reconnection due to the exponentiation of field line separations as the pre-Aref theories of fluid mixing to that of Aref Aref;1984 . When macro-turbulence is three dimensional, it does cause the magnetic field to become chaotic. Nevertheless when the characteristic spatial scale of the turbulent eddies turb\ell_{turb} is small compared to the scale aa over which strong advection takes place, the advection from the eddies behaves as a diffusive process Boozer:rec-phys . The diffusive advection of the turbulent eddies is much slower than the advective transport due to flows that have a scale comparable to aa unless the maximum flow speed is far larger than the flow speed averaged over the scale turb\ell_{turb}. A deterministic turbulent ideal flow u\vec{u}_{\bot} does not directly produce magnetic reconnection any more than a laminar flow does; both cause an exponentially enhanced sensitivity to non-ideal effects.

The literature on macro-turbulence enhanced reconnection is extensive. However, the focus of these studies was not on on deterministic magnetic fields and flows. The effect on reconnection of indeterminacy in the magnetic field, which means intrinsic magnetic stochasticity, was described in Lazarian:1999 ; Eyink:2011 ; Eyink:2015 ; Kowal:2020 , and reviewed in Lazarian:2020rev . The effect of turbulence and intermittency of plasma flows on two-dimensional magnetic fields was described in Matthaeus1986 ; Matthaeus:2015 ; Matthaeus:2020 .

The simple model developed here illustrates how the laminar flow of an ideal magnetic evolution can force magnetic reconnection on a time scale that is the ideal evolution time, τeva/u\tau_{ev}\equiv a/u_{\bot} times the logarithm of the ratio of the intrinsic magnitude of the ideal to the connection breaking terms.

A spatially bounded flow of a footpoint of a magnetic field line that has a smooth and slow variation in space and time can be chaotic and impart properties to the magnetic field lines in an ideal evolution that many may find to be surprising.

Two magnetic field lines that are separated by the distance Δmin\Delta_{min} at the stationary foot point must have a maximum separation Δmax\Delta_{max} at least as great as the separation of the lines at the moving footpoint. The separation of neighboring streamlines determines the minimum Δmax/Δmin\Delta_{max}/\Delta_{min} ratio that the two magnetic field lines must have obtained. Figure 4c shows that no matter how small Δmin\Delta_{min} may be that Δmax\Delta_{max} will reach a value comparable to aa, the distance scale of the system perpendicular to the magnetic field, within a time that depends only logarithmically on Δmin\Delta_{min}. Identifying Δmin\Delta_{min} with the spatial scale Δd\Delta_{d} on which magnetic field lines become indistinguishable, the time required for large-scale reconnection to occur is the evolution time defined by the footpoint flow speed times a logarithmic factor depending on a/Δda/\Delta_{d}. This argument assumes the reconnection scale Δr\Delta_{r} is comparable to aa, although it could be smaller, which would imply large scale reconnection would occur even sooner.

Many more properties of an ideally evolving magnetic field can be determined from the footpoint motion when its evolution time is long compared to the time for an Alfvén wave to propagate from one footpoint interception to the other and the magnetic field remains kink stable. The example of the footpoint flow that was studied was chosen to be consistent with the maintenance of kink stability. Two properties that can be determined are the distribution of the current density Kμ0j||/BK\equiv\mu_{0}j_{||}/B, which is constant along the magnetic field lines under these conditions, and the power that is required to drive the footpoint flow. The figures show the quantity KLKL rather than KK because KLKL is the quantity that is determined by the theory; not KK and LL separately. KL/2KL/2 is the dimensionless twist angle that a magnetic field line makes in going between the two footpoints of each field line, which are separated by the distance LL.

Despite the smoothness and simplicity of the footpoint flow that was studied, the current density distribution KK varies wildly among field lines that have their stationary foopoints in a tiny region, Figure 2c. The currents in the corona must be extremely complicated with a short correlation distance across the magnetic field lines. Observations of spatial complexity in the current density are clearly not a proof of turbulence. Nearby magnetic field lines can have currents flowing in opposite directions. In addition the magnitude of KK increases only linearly in time, Figure 2c, and forms ribbons along the magnetic field that tend to widen exponentially and thin as one over the exponential, Figure 5. As discussed in Appendix E of Boozer:part.acc , the currents produced by the more localized photospheric motions could produce runaway electrons and explain the hot corona.

To be consistent with Ampere’s law, KK must have a typical magnitude that is proportional to the logarithm of the exponentiation factor Boozer:B-line.sep . A heuristic argument is given in Section III.4. The exponentiation factor is the Frobenius norm of a Jacobian matrix, Section II.3. This approximate proportionality is illustrated in Figure 4c and requires the current density be enhanced by factor of ln(Rm)\ln(R_{m}) above its characteristic value before reconnection due to exponentially enhanced resistive diffusion can compete with evolution. An enhancement of the current density by ln(Rm)ln(1012)28\ln(R_{m})\sim\ln(10^{12})\approx 28 is modest compared to an enhancement by Rm1012R_{m}\sim 10^{12}. Despite this relationship between KK and the Frobenius norm, the regions in which they are particularly large are not closely correlated. Reid et al Reid:2020 also found a weak correlation between the current density j||j_{||} and ratio at the two footpoint locations of the distance between neighboring field lines, which is the definition of the quasi-squashing factor Priest:1995 ; Titov: 2002 .

The power required to drive the footpoint motion, Section VII, is a spatial average of the footpoint motion times the current distribution KK, so it increases approximately linearly in time and never becomes extremely large.

Section VI demonstrated that unless the stream function ht(x,y,t)h_{t}(x,y,t), which defines the footpoint flow, has a zero spatial and temporal average over a footpoint region, the magnetic helicity increases without limit in the volume occupied by the magnetic field lines that strike that region. When Rm>>1R_{m}>>1, magnetic helicity cannot be destroyed faster that it is created in such regions—even by macro-turbulence—and the result must be the ejection of a magnetic flux tube, called a plasmoid, from the region.

The example of a footpoint flow that was discussed here had a very simple and smooth spatial and temporal dependence; realistic footpoint motions of coronal loops are far more complicated, which makes them more prone to chaos. However, the parts of the footpoint motion that cause the most rapid development of states in which large-scale reconnection is inevitable have a spatial scale comparable to the region aa over which reconnection will take place.

The calculations made in this paper demonstrate that a magnetic field reaches a state in which reconnection is inevitable on a time scale that is only logarithmically longer than the ideal evolution time scale, even when drive for the evolution is simple.

Nonetheless, more complete simulations related to those of Reid et al Reid:2018 , are required for a more complete understanding of how the reconnection proceeds. It is particularly important to use such simulations to understand how the magnetic reconnection region is determined, not because it is bounded by a rigid conductor but by unstressed magnetic field lines. This can be done by choosing a non-zero λ2\lambda^{2} in Equation (5). As λ2\lambda^{2} is made larger within the model of this paper, simulations become more expensive but more representative of the corona. Such simulations are numerically simpler when the circular cylinder of Figure 1 is replaced by a cylinder of square cross section that encloses the volume with |x|<a|x|<a, |y|<a|y|<a, and 0<z<L0<z<L. As mentioned at the end of Section II.2, the only significant modification required is a change in the factors of 1r2/a21-r^{2}/a^{2} in Equation (5) to (1x2/a2)(1y2/a2)(1-x^{2}/a^{2})(1-y^{2}/a^{2}).

Acknowledgements

The authors wish to thank Yi-Min Huang for pointing out an error in the derivation of Equation (31) in an earlier version of the manuscript. This work was supported by the U.S. Department of Energy, Office of Science, Office of Fusion Energy Sciences under Award Numbers DE-FG02-95ER54333, DE-FG02-03ER54696, DE-SC0018424, and DE-SC0019479.

Data availability statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  • (1) P. Cargill (2015), Magnetic Reconnection in the Solar Corona: Historical Perspective and Modern Thinking. In: Cowley FRS S., Southwood D., Mitton S. (eds) Magnetospheric Plasma Physics: The Impact of Jim Dungeyś Research, Astrophysics and Space Science Proceedings, vol 41, Springer, Cham.
  • (2) E. N. Parker and M. Krook, Diffusion and severing of magnetic lines of force, Ap. J. 124, 214 (1956).
  • (3) W. A. Newcomb, Motion of magnetic lines of force, Ann. Phys. 3, 347 (1958).
  • (4) A. H. Boozer, Fast magnetic reconnection and the ideal evolution of a magnetic field, Phys. Plasmas 26, 042104 (2019).
  • (5) L. Spitzer, Physics of Fully Ionized Gases, second edition (John Wiley and Sons, New York, 1962). See Equation (2-12).
  • (6) A. H. Boozer, Magnetic Reconnection with null and X-points, Phys. Plasmas 26, 122902 (2019).
  • (7) K. Schindler, M. Hesse, and J. Birn, General magnetic reconnection, parallel electric-fields, and helicity, Journal of Geophysical Research—Space Physics 93, 5547 (1988).
  • (8) D. A. Uzdensky, N. F. Loureiro, and A. A. Schekochihin, Fast Magnetic Reconnection in the Plasmoid-Dominated Regime, Phys. Rev. Lett. 105, 235002 (2010).
  • (9) A. H. Boozer, Magnetic reconnection and thermal equilibration, Phys. Plasmas 28, 032102 (2021).
  • (10) H. Aref, Stirring by chaotic advection, Journal of Fluid Mechanics [143], 1 (1984).
  • (11) H. Aref, J. R. Blake, Marko Budisić, S. S. S. Cardoso, J. H. E. Cartwright, H. J. H. Clercx, K. El Omari, U. Feudel, R. Golestanian, E. Gouillart, G. J. F. van Heijst, T. S. Krasnopolskaya, Y. Le Guer, R. S. MacKay, V. V. Meleshko, G. Metcalfe, I. Mezić, A. P. S. de Moura, O. Piro, M. F. M. Speetjens, R. Sturman, J.-L. Thiffeault, and I. Tuval, Frontiers of chaotic advection, Rev. Mod. Phys. 89, 025007 (2017).
  • (12) L. F. Richardson, Atmospheric diffusion shown on a distance-neighbour graph, Proc. R. Soc. London Ser. A 110, 709 (1926).
  • (13) A. Lazarian and E. T. Vishniac, Reconnection in a weakly stochastic field, Ap. J. 517, 700 (1999).
  • (14) T. S. Hahm and R. M. Kulsrud, Forced magnetic reconnection, Phys. Fluids 28, 2412 (1985).
  • (15) T. Elder and A. H. Boozer, Magnetic nulls in interacting dipolar fields, https://arxiv.org/pdf/2005.08242.pdf, accepted for publication by the Journal of Plasma Physics.
  • (16) E. N. Parker EN (1972) Topological dissipation and the small-scale fields in turbulent gases, Astrophys. J. 174, 499 (1972).
  • (17) David I. Pontin and Gunnar Hornig, The Parker problem: existence of smooth force-free fields and coronal heating, Living Reviews in Solar Physics 17, 5 (2020).
  • (18) A. H. Boozer, Why fast magnetic reconnection is so prevalent, Journal of Plasma Physics 84, 715840102 (2018).
  • (19) J. Reid, A. W. Hood, C. E. Parnell, P. K. Browning, and P. J. Cargill, Coronal energy release by MHD avalanches: continuous driving, Astronomy and Astrophysics 615, A84 (2018).
  • (20) A. H. Boozer, Fast magnetic reconnection and particle acceleration, Phys. Plasmas 26, 082112 (2019).
  • (21) J. Reid, C. E. Parnell, A. W. Hood, and P. K. Browning, Determining whether the squashing factor, Q, would be a good indicator of reconnection in a resistive MHD experiment devoid of null points, Astronomy and Astrophysics 633, A92 (2020).
  • (22) D. Borgogno, D. Grasso, F. Pegoraro, and T. J. Schep, Barriers in the transition to global chaos in collisionless magnetic reconnection. I. Ridges of the finite time Lyapunov exponent field, Phys. Plasmas 18, 102307 (2011).
  • (23) Yi-Min Huang, A. Bhattacharjee, and A. H. Boozer, Rapid change of field line connectivity and reconnection in stochastic magnetic fields, Ap. J. 793, 106 (2014).
  • (24) B. B. Kadomtsev and O. P. Pogutse, Nonlinear helical perturbations of a plasma in the tokamak, Sov. Phys.-JETP, 38, 283 (1974).
  • (25) H. R. Strauss, Nonlinear, 3-dimensional magnetohydrodnamics of noncircular tokamaks, Phys. Fluids 19, 134 (1976).
  • (26) A. H. Boozer and N. Pomphrey, Current density and plasma displacement near perturbed rational surfaces, Phys. Plasmas 17, 110707 (2010).
  • (27) A. H. Boozer, Flattening of the tokamak current profile by a fast magnetic reconnection with implications for the solar corona, Phys. Plasmas 27, 102305 (2020).
  • (28) A. H. Boozer, Separation of magnetic field lines, Phys. Plasmas 19, 112901 (2012).
  • (29) A. W. Hood and E. R. Priest, Kink instability of solar coronal loops as the cause of solar flares, Solar Physics 64, 303 (1979).
  • (30) A. H. Boozer, Physics of magnetically confined plasmas, Rev. Mod. Phys. 76, 1071 (2004).
  • (31) J. B. Taylor, Relaxation of Toroidal Plasma and Generation of Reverse Magnetic Fields, Phys. Rev. Lett. 33, 1139 (1974).
  • (32) M. A. Berger, Rigorous new limits on magnetic helicity dissipation in the solar corona, Geophysical and Astrophysical Fluid Dynamics, 30, 79 (1984).
  • (33) E. Pariat, S. K. Antiochos, and C. R. DeVore, A model for solar polar jets, Ap. J. 691, 61 (2009).
  • (34) E. N. Parker, The reconnection rate of magnetic fields, Ap. J. 180, 247 (1973).
  • (35) J. W. Dungey, The motion of magnetic fields, Monthly Notices of the Royal Astronomical Society 113, 679 (1953).
  • (36) G. L. Eyink, A. Lazarian, E. T. Vishniac, Fast magnetic reconnection and spontaneous stochasticity, Ap. J. 743, 51 (2011).
  • (37) G. L. Eyink, Turbulent general magnetic reconnection, Ap. J 807 137 (2015).
  • (38) G. Kowal, Di A. Falceta-Gon¸alves, A. Lazarian, and E. T. Vishniac, Kelvin-Helmholtz versus Tearing Instability: What Drives Turbulence in Stochastic Reconnection?, Ap. J. 892, 50 (2020).
  • (39) A. Lazarian, G. L. Eyink, A. Jafari, G. Kowal, H. Li, S-Y Xu, and E. T. Vishniac, 3D turbulent reconnection: Theory, tests, and astrophysical implications, Phys. Plasmas 27, 012305 (2020).
  • (40) W. H. Matthaeus and S. L. Lamkin, Turbulent magnetic reconnection, Phys. Fluids 29, 2513 (1986).
  • (41) W. H. Matthaeus, M. Wan, S. Servidio, A. Greco, K. T. Osman, S. Oughton, and P. Dmitruk, Intermittency, nonlinear dynamics and dissipation in the solar wind and astrophysical plasmas, Phil. Trans. R. Soc. A 373 20140154 (2015).
  • (42) S. Adhikari, M. A. Shay, T. N. Parashar, P. S. Pyakurel, W. H. Matthaeus, D. Godzieba, J. E. Stawarz, J. P. Eastwood, and J. T. Dahlin, Reconnection from a turbulence perspective, Phys. Plasmas 27, 042305 (2020).
  • (43) E. R. Priest and P. Démoulin, Three-dimensional magnetic reconnection without null points: 1. Basic theory of magnetic flipping, J. Geophys. Res., 100, 23443 (1995).
  • (44) V. S. Titov, G. Hornig, and P. Démoulin, Theory of magnetic connectivity in the solar corona, J. Geophys. Res., 107, 1164 (2002).