Hostname: page-component-7bb8b95d7b-5mhkq Total loading time: 0 Render date: 2024-09-26T13:47:26.592Z Has data issue: false hasContentIssue false

The Okubo–Weiss-type topological criteria in two-dimensional magnetohydrodynamic flows

Published online by Cambridge University Press:  16 April 2024

B.K. Shivamoggi*
Affiliation:
Department of Mathematics, University of Central Florida, Orlando, FL 32816, USA
G.J.F. van Heijst
Affiliation:
J. M. Burgers Centre and Fluid Dynamics Laboratory, Eindhoven University of Technology, NL-5600MB Eindhoven, The Netherlands
L.P.J. Kamp
Affiliation:
J. M. Burgers Centre and Fluid Dynamics Laboratory, Eindhoven University of Technology, NL-5600MB Eindhoven, The Netherlands
*
Email address for correspondence: bhimsen.shivamoggi@ucf.edu

Abstract

The Okubo–Weiss (Okubo, Deep-Sea Res., vol. 17, issue 3, 1970, pp. 445–454; Weiss, Physica D, vol. 48, issue 2, 1991, pp. 273–294) criterion has been widely used as a diagnostic tool to divide a two-dimensional (2-D) hydrodynamical flow field into hyperbolic and elliptic regions. This paper considers extension of these ideas to 2-D magnetohydrodynamic (MHD) flows, and presents an Okubo–Weiss-type criterion to parameterize the magnetic field topology in 2-D MHD flows. This ensues via its topological connections with the intrinsic metric properties of the underlying magnetic flux manifold, and is illustrated by recasting the Okubo–Weiss-type criterion via the 2-D MHD stationary generalized Alfvénic state condition to approximate the slow-flow-variation ansatz imposed in its derivation. The Okubo–Weiss-type parameter then turns out to be related to the sign definiteness of the Gaussian curvature of the magnetic flux manifold. A similar formulation becomes possible for 2-D electron MHD flows, by using the generalized magnetic flux framework to incorporate the electron-inertia effects. Numerical simulations of quasi-stationary vortices in 2-D MHD flows in the decaying turbulence regime are then given to demonstrate that the Okubo–Weiss-type criterion is able to separate the MHD flow field into elliptic and hyperbolic field configurations very well.

Type
Research Article
Copyright
Copyright © The Author(s), 2024. Published by Cambridge University Press

1 Introduction

A major issue in two-dimensional (2-D) hydrodynamic turbulent flows concerns dividing a vorticity field into hyperbolic and elliptic regions because these regions exhibit different particle transport properties. The hyperbolic regions correspond to the relative dominance of flow deformation and hence exhibit exponential divergence of nearby particles (Elhmaïdi, Provenzale & Babiano Reference Elhmaïdi, Provenzale and Babiano1993; Babiano et al. Reference Babiano, Boffetta, Provenzale and Vulpiani1994). The elliptic regions correspond, on the other hand, to the relative dominance of flow rotation and hence exhibit near-zero divergence of nearby particles. Okubo (Reference Okubo1970) and Weiss (Reference Weiss1991) gave a kinematic criterion to provide a diagnostic tool to address this issue, and it has been validated by numerical simulations (Mcwilliams Reference Mcwilliams1984; Brachet et al. Reference Brachet, Meneguzzi, Politano and Sulem1988; Ohkitani Reference Ohkitani1991; Babiano & Provenzale Reference Babiano and Provenzale2007) and laboratory experiments (Ouellette & Gollub Reference Ouellette and Gollub2007) of 2-D hydrodynamic flows. However, this validation has been primarily based on empirical grounds by the results of its application. Recently, Shivamoggi, Van Heijst & Kamp (Reference Shivamoggi, Van Heijst and Kamp2022) explored some topological interpretations of the Okubo–Weiss criterion in terms of the intrinsic metric properties of the underlying vorticity manifold. The Okubo–Weiss parameter was shown, to within a positive multiplicative factor, to be the negative of the Gaussian curvature of the vorticity manifold.

The purpose of this paper is to formulate an Okubo–Weiss-type criterion for 2-D magnetohydrodynamic (MHD) flows. The Okubo–Weiss-type parameter is shown to provide a useful diagnostic tool to parametrize the magnetic field topology in 2-D MHD flows and to provide some topological interpretations in terms of the intrinsic metric properties of the underlying magnetic flux manifold. This is accomplished by recasting the Okubo–Weiss-type criterion via the 2-D MHD stationary generalized Alfvénic state condition to approximate the slow flow-variation ansatz imposed in its derivation. The Okubo–Weiss parameter then turns out to be related to sign definiteness of the Gaussian curvature of the magnetic flux manifold. Reformulation of the Okubo–Weiss-type criterion in plane polar coordinates is shown to facilitate its validation for several real plasma examples like the MHD Rankine-type flux tube and the $Z$-pinch. A similar formulation is shown to become possible for non-standard MHD cases, like 2-D electron MHD (EMHD) flows, by using the generalized magnetic flux framework to incorporate the electron-inertia effects. Numerical simulations of quasi-stationary vortices in 2-D MHD flows in the decaying turbulence regime are then given to demonstrate the Okubo–Weiss-type criterion is able to separate the MHD flow field into elliptic and hyperbolic field configurations very well.

2 The Okubo–Weiss-type criterion for MHD flows

The equation governing the transport of the magnetic field ${\boldsymbol {B}} = \langle B_1, B_2 \rangle$ in 2-D (or 3-D) incompressible MHD flow is (Goedbloed & Poedts Reference Goedbloed and Poedts2004), in usual notation,

(2.1a)\begin{align} \frac{\partial {\boldsymbol{B}}}{\partial t} = \boldsymbol{\nabla} \times ( {\boldsymbol{v}} \times {\boldsymbol{B}} ) \end{align}

or

(2.1b)\begin{align} \frac{{\rm D} {\boldsymbol{B}}}{{\rm D} t} = \frac{\partial {\boldsymbol{B}}}{\partial t} + \left( {\boldsymbol{v}}\boldsymbol{\cdot}\boldsymbol{\nabla}\right) {\boldsymbol{B}} = ({\boldsymbol{B}} \boldsymbol{\cdot} \boldsymbol{\nabla} ){\boldsymbol{v}} \equiv \boldsymbol{\mathcal{A}} \boldsymbol{\cdot}{\boldsymbol{B}}, \end{align}

where $\boldsymbol {\mathcal {A}}$ is the flow-velocity gradient matrix and ${\boldsymbol {v}}$ is defined by ${\boldsymbol {v}} = \langle u, v \rangle$.

It may be of interest to note that (2.1a) for the 2-D transport of a magnetic field in MHD is the same as that for the 2-D transport of the fluid divorticity (which is the curl of the fluid vorticity). This may be understood by noting the well-known analogy between vorticity in 3-D hydrodynamics and a magnetic field in 3-D MHD, and the analogy between vorticity in 3-D hydrodynamics and divorticity in 2-D hydrodynamics (Kuznetsov et al. Reference Kuznetsov, Naulin, Nielsen and Rasmussen2007).

If we assume, following Okubo (Reference Okubo1970) and Weiss (Reference Weiss1991), the straining flow-velocity gradient tensor $\boldsymbol {\nabla }{\boldsymbol {v}}$ to temporally evolve slowly, so the magnetic field can be taken to evolve adiabatically with respect to the straining flow-velocity gradient field,Footnote 1 (2.1b) may be locally approximated by an eigenvalue problem, with eigenvalues given by

(2.2)\begin{align} \lambda^2 = u_y v_x + {v_y}^2 \equiv Q. \end{align}

An interesting interpretation of the Okubo–Weiss-type parameter $Q$ for the MHD case follows upon noting that the slow-flow-variation ansatz used above may be approximated by the MHD stationary Beltrami flow state (Shivamoggi Reference Shivamoggi2011). From (2.1a), the latter corresponds to the generalized Alfvénic state

(2.3)\begin{equation} {\boldsymbol{B}} = a {\boldsymbol{v}}, \end{equation}

with $a$ being an arbitrary constant.Footnote 2 Equation (2.3) implies that the vorticity $\boldsymbol {\omega } = \boldsymbol {\nabla } \times {\boldsymbol {v}}$ is proportional to the electric current density ${\boldsymbol {J}} = \boldsymbol {\nabla } \times {\boldsymbol {B}}$ for a generalized Alfvénic state in 2-D MHD. Using (2.3), (2.2) becomes

(2.4)\begin{equation} \lambda^2\equiv Q = \frac{1}{a^2} \left(\displaystyle B_{1y} {B_2}_x + {B^2_{2y}} \right)=\frac{1}{4a^2}\left( b_1^2+ b_2^2- J^2/c^2\right), \end{equation}

where

(2.5a–c)\begin{equation} b_1\equiv-2B_{2y}, \quad b_2\equiv{B_2}_x+B_{1y},\quad\frac{J}{c}\equiv B_{2x}-B_{1y}.\end{equation}

Equation (2.4) shows that the Okubo–Weiss-type parameter $Q$ is a measure of the relative importance of magnetic shear ($Q>0$, hyperbolic) and electric current ($Q < 0$, elliptic).Footnote 3 This result, in conjunction with (2.2) and (2.3), is in accord with the eigenvalues of the magnetic field gradient tensor $\boldsymbol {\nabla }{\boldsymbol {B}}$ becoming purely imaginary near the $O$-points or real near the $X$-points of the magnetic field lines (Greene Reference Greene1993).

The hyperbolic magnetic field topology is of great interest in plasmas in space and fusion devices because it becomes a seat for magnetic reconnection processes (Giovanelli Reference Giovanelli1949; Shivamoggi Reference Shivamoggi1986Reference Shivamoggi1999; Rollins & Shivamoggi Reference Rollins and Shivamoggi2007).

In terms of the magnetic vector potential ${\boldsymbol {A}}$ given by

(2.6a,b)\begin{equation} {\boldsymbol{B}} \equiv \boldsymbol{\nabla} \times {\boldsymbol{A}}, \quad {\boldsymbol{A}} = A \,{\hat{\boldsymbol{i}}}_z, \end{equation}

the Okubo–Weiss-type parameter (2.4) becomes

(2.7)\begin{equation} Q = \frac{1}{a^2} \left[ \left( \frac{\partial^2 A}{\partial x \partial y}\right)^2 - \frac{\partial^2 A}{\partial x^2} \frac{\partial^2 A}{\partial y^2} \right]\!. \end{equation}

Equation (2.7) shows that the Okubo–Weiss-type parameter $Q$ for the MHD case is the negative of the Gaussian curvature of the magnetic flux surface, to within a positive multiplicative factor. Since the 2-D MHD magnetic field topology is characterized by the local Gaussian curvature of the underlying magnetic flux manifold, (2.2) has the potential to serve as a useful diagnostic tool to parameterize the magnetic field topology in 2-D MHD flows in terms of magnetic-shear-dominated and electric-current-dominated regions. This is numerically investigated in § 5.

This development may be formally extended to the 2-D EMHD case via the generalized magnetic flux framework to incorporate the effects of electron inertia. The Okubo–Weiss-type parameter for the 2-D EMHD case turns out to be related to the intrinsic metric properties of the out-of-plane magnetic field manifold (see Appendix A).

Example 2.1 As an example of the above formulation, consider a MHD flow with the magnetic flux function given by

(2.8a)\begin{equation} A( x,y)=\frac{k}{2}\left( \alpha x^2-y^2\right)\!. \end{equation}

So, the magnetic field is given by the hyperbolic/elliptic configuration near an $X(O)$-type $(\alpha \gtrless 0)$ magnetic neutral point

(2.8b)\begin{equation} B_1=-ky,\quad B_2=-k\alpha x.\end{equation}

Thus, we obtain

(2.9a–c)\begin{equation} b_1=0,\quad b_2=-k( 1+\alpha),\quad \frac{J}{c}=k( 1-\alpha),\end{equation}

which shows that the current density $J\not =0$, unless $\alpha = 1$. On the other hand, the magnetic field topology is determined only by whether $\alpha \lessgtr 0$, as shown below.

Using (2.9ac), (2.4) gives

(2.10)\begin{equation} Q=\left( \frac{k^2}{a^2}\right) \alpha \gtrless 0, \quad\alpha \gtrless 0, \end{equation}

as to be expected.

3 Reformulation in polar coordinates

In plane polar coordinates $(r,\theta )$, (2.1b) written in the component form is

(3.1a)\begin{gather} \frac{\partial B_r}{\partial t}+v_r\frac{\partial B_r}{\partial r}+v_\theta\frac{\partial B_r}{r\partial\theta}-\frac{v_\theta B_\theta}{r}=B_r\frac{\partial v_r}{\partial r}+B_\theta\frac{\partial v_r}{r\partial \theta}-\frac{B_\theta v_\theta}{r} , \end{gather}
(3.1b)\begin{gather}\frac{\partial B_\theta}{\partial t}+v_r\frac{\partial B_\theta}{\partial r}+v_\theta\frac{\partial B_\theta}{r\partial\theta}+\frac{v_\theta B_r}{r}=B_r\frac{\partial v_\theta}{\partial r}+B_\theta\frac{\partial v_\theta}{r\partial \theta}+\frac{B_\theta v_r}{r}. \end{gather}

The velocity gradient matrix $\boldsymbol {\mathscr {A}}$ in (2.1b) then becomes

(3.2)\begin{equation} \boldsymbol{\mathscr{A}}\equiv\left[ \begin{array}{@{}cc@{}} \dfrac{\partial v_r}{\partial r} & \dfrac{\partial v_r}{r\partial\theta}-\dfrac{v_\theta}{r}\\[10pt] \dfrac{\partial v_\theta}{\partial r} & \dfrac{\partial v_\theta}{r\partial\theta}+\dfrac{v_r}{r} \end{array} \right]\!. \end{equation}

On using the mass-conservation equation

(3.3)\begin{equation} \frac{1}{r}\frac{\partial}{\partial r}\left( rv_r\right)+\frac{\partial v_\theta}{r\partial\theta}=0, \end{equation}

and assuming again that the magnetic field evolves adiabatically with respect to the straining flow-velocity gradient field, (3.1a,b) may again be locally approximated by an eigenvalue problem, with eigenvalues given by

(3.4)\begin{equation} \lambda^2=\left(\frac{\partial v_r}{\partial r}\right)^2+ \displaystyle \frac{1}{r}\frac{\partial v_r}{\partial\theta}\frac{\partial v_\theta}{\partial r}-\frac{1}{r}v_\theta\frac{\partial v_\theta}{\partial r}\equiv Q. \end{equation}

Equation (3.4) is the same as that for the divorticity field in 2-D hydrodynamics (Shivamoggi et al. Reference Shivamoggi, Van Heijst and Kamp2022).

On using the stationary generalized Alfvénic state condition (2.3), (3.4) becomes

(3.5a)\begin{equation} \lambda^2=\frac{1}{a^2}\left[\left(\frac{\partial B_r}{\partial r}\right)^2+\frac{1}{r}\frac{\partial B_r}{\partial\theta}\frac{\partial B_\theta}{\partial r}-\frac{1}{r}B_\theta\frac{\partial B_\theta}{\partial r}\right]\equiv Q.\end{equation}

On using the Gauss law

(3.6)\begin{equation} \frac{1}{r}\frac{\partial}{\partial r}\left( r B_r\right) +\frac{\partial B_\theta}{r\partial\theta}=0, \end{equation}

equation (3.5a) may be alternatively expressed as

(3.5b)\begin{equation} \lambda^2=\frac{1}{a^2}\left[\left(\frac{\partial B_\theta}{\partial r}\right)\left(\frac{\partial B_r}{r\partial\theta}-\frac{B_\theta}{r}\right)+\left(\frac{B_r}{r}+\frac{\partial B_\theta}{r\partial\theta}\right)^2\right]\equiv Q.\end{equation}

In terms of the magnetic vector potential ${\boldsymbol {A}} = A \hat {{\boldsymbol {i}}}_z$, given by (2.6ac), (3.5b) becomes

(3.7)\begin{equation} \lambda^2=\frac{1}{a^2r^4}\left[-r^2\frac{\partial^2 A}{\partial r^2}\left(\frac{\partial^2A}{\partial\theta^2}+r\frac{\partial A}{\partial r}\right)+\left(\frac{\partial A}{\partial\theta}-r\frac{\partial^2 A}{\partial r\partial\theta}\right)^2\right] \equiv Q. \end{equation}

Equation (3.7) shows that $\lambda ^2$ is the negative of the Gaussian curvature of the magnetic flux manifold, to within a positive multiplicative factor.

Example 3.1 As an example, consider an axisymmetric MHD flow with magnetic flux function given by

(3.8a)\begin{equation} A=A( r). \end{equation}

Equation (3.8a) leads to the magnetic field

(3.8b)\begin{equation} B_r=0,\quad B_\theta=-\frac{{\rm d}A}{{\rm d}r}. \end{equation}

Using (3.8), (3.7) gives for the eigenvalues

(3.9)\begin{equation} \lambda^2=-\frac{1}{r}\frac{{\rm d}A}{{\rm d}r}\frac{{\rm d}^2A}{{\rm d}r^2}. \end{equation}

Consider the $Z$-pinch problem, where the magnetic field is generated by a uniform, unidirectional current ${\boldsymbol {J}}=J{\hat {\boldsymbol {i}}}_z$,Footnote 4

(3.10)\begin{equation} A=-\frac{1}{4c}Jr^2 . \end{equation}

Using (3.10), (3.9) becomes

(3.11)\begin{equation} \lambda^2=-\frac{J^2}{4c^2}<0, \end{equation}

as to be expected.

Consider next a magnetic field generated by a current-carrying cylinder $(0< r< r_0)$

(3.12)\begin{equation} J=J_0h(r_0-r), \end{equation}

where $h\left ( x\right )$ is the unit step function, and $J_0$ is the uniform current density in this cylinder. The magnetic flux function associated with (3.12) is given by

(3.13)\begin{equation} A=\begin{cases} \displaystyle-\frac{1}{2}\frac{J_0r^2}{c}, & r< r_0\\[10pt] \displaystyle-\frac{J_0r_0^2}{c}\ln(r/r_0), & r>r_0, \end{cases} \end{equation}

and the magnetic field is given by

(3.14)\begin{equation} B_\theta= \begin{cases} \displaystyle\frac{J_0r}{c}, & r< r_0\\[10pt] \displaystyle\frac{J_0r_0^2}{rc}, & r>r_0, \end{cases} \end{equation}

which represents an MHD Rankine-type flux tube.

Using (3.14), (3.5a) gives

(3.15)\begin{equation} \lambda^2= \begin{cases} \displaystyle-\frac{J_0^2}{a^2c^2}, & r< r_0\\[10pt] \displaystyle\frac{J_0^2r_0^4}{c^2a^2r^4}, & r>r_0, \end{cases} \end{equation}

signifying an elliptic region inside the current-carrying cylinder and a hyperbolic region outside it, as to be expected.

4 Numerical set-up

The Okubo–Weiss-type criteria for 2-D MHD flows are now illustrated by performing numerical simulations of the relevant equations in dimensionless form. The fluid flow occurs in a planar 2-D domain, with the velocity field given by ${\boldsymbol {v}}=u(x,y,t){\hat {\boldsymbol {i}}}_x+v(x,y,t){\hat {\boldsymbol {i}}}_y$, where ${\hat {\boldsymbol {i}}}_x$ and ${\hat {\boldsymbol {i}}}_y$ are the unit vectors in the $x$- and $y$-directions, respectively. Additionally, there is a (dimensionless) magnetic field, which is normalized using some characteristic magnetic field strength $B_0$, ${\boldsymbol {B}}=B_1(x,y,t){\hat {\boldsymbol {i}}}_x+B_2(x,y,t){\hat {\boldsymbol {i}}}_y$.

For MHD flows, the governing (dimensionless) equations are the equation of motion

(4.1)\begin{equation} \frac{\partial{\boldsymbol{v}}}{\partial t}+\left({\boldsymbol{v}}\boldsymbol{\cdot}\boldsymbol{\nabla}\right){\boldsymbol{v}}=- \frac{1}{\rho} \boldsymbol{\nabla} p+ \frac{1}{R_e}\nabla^2{\boldsymbol{v}}+J{\hat{\boldsymbol{i}}}_z\times{\boldsymbol{B}}, \end{equation}

where ${\hat {\boldsymbol {i}}}_z$ is the unit vector in the $z$-direction and the magnetic field transport equation

(4.2)\begin{equation} \frac{\partial{\boldsymbol{B}}}{\partial t}=\boldsymbol{\nabla}\times\left({\boldsymbol{v}}\times {\boldsymbol{B}}\right)+\frac{1}{R_m}\nabla^2{\boldsymbol{B}}. \end{equation}

In (4.1), velocity has been non-dimensionalized using a reference speed $U$, and length has been normalized using a reference length $L$; $R_e$ is the Reynolds number given by $R_e\equiv UL/\nu$, where $\nu$ is the kinematic viscosity of the fluid; $R_m$ is the magnetic Reynolds number given by $R_m\equiv \mu _0UL/\eta$, $\mu _0$ being the free space permeability; and $\eta$ is the resistivity. The electric current density $J$ follows from Ampère's law $J=(\partial B_2/\partial x-\partial B_1/\partial y)$.

Equations (4.1) and (4.2) have been numerically solved with a finite-element code Comsol,Footnote 5 using double-periodic boundary conditions on a computational domain defined by $-5\leq x\leq 5$ and $-5\leq y\leq 5$. The velocity field has been initialized with a statistically steady turbulent flow is field given by $u(x,y,t=0)=u_{0}(x,y)$ and $v(x,y,t=0)=v_{0}(x,y)$, with root-mean-square value equal to 1. In consistency with the generalized Alfvénic state condition (2.3), the magnetic field has also been initialized with this turbulent field, i.e. $B_1(x,y,t=0)=u_0(x,y)$ and $B_2(x,y,t=0)=v_0(x,y)$. This corresponds to putting $a=1$ (on taking $\rho =1$, plasma being incompressible) in (2.3), which as mentioned in footnote 2, reflects the nonlinear Alfvénic state associated with perturbations in a uniformly magnetized plasma (Hasegawa Reference Hasegawa1985). All numerical simulations were performed with $R_e=1000=R_m$. Integrations have been carried out using a second-order backward-difference scheme and an implicit time stepping.

This statistically steady turbulent state is set up by starting with a quiescent fluid which is mechanically forced using an external body force acting on the fluid. The details of this forcing become irrelevant once a turbulent flow develops, and then the forcing is turned off. The kinetic energy of the flow increases with time, eventually reaching a quasi-steady value, which indicates the attainment of a statistically steady state. The root-mean-square value of the flow velocity then becomes quasi-steady, even though the flow is turbulent and time dependent. As time progresses, this turbulent flow state, which is not forced anymore, starts to decay concomitantly creating large-scale coherent structures. Figure 1 gives a transient snapshot of this time-dependent decaying turbulent flow showing these large-scale structures, which may be identified with an inverse cascade.

Figure 1. Snapshot of the vorticity field (a) and the associated Okubo–Weiss function (b) taken during the decaying phase of MHD turbulence. The colour bar denotes the value of the Okubo–Weiss function, which has been rescaled to range from $-1$ to $+1$.

Our numerical simulations are carried out using a finite-element method for which the computational domain is discretized using an ultra-fine triangular mesh – a typical mesh element has a size of 0.01 (in normalized units), resulting in $10^6$ mesh elements. The intended accuracy of the method is, among other things, determined by the number of mesh elements in the computational domain rather than the number of grid points (which are dealt with better by a finite-difference method rather than a finite-element method).

5 Numerical results

For a generalized Alfénic state in 2-D MHD, as discussed in § 2, the spatial pattern of the electrical current density ${\boldsymbol {J}}$ is similar to that of the vorticity field $\boldsymbol {\omega }$. Therefore, the regions of concentrated electrical current density can be visualized via the Okubo–Weiss function $Q = u_y v_x+v_y^2$. Figure 1 shows a snapshot of the numerically calculated vorticity field (a) and the associated Okubo–Weiss function (b) taken during the decaying phase of a MHD turbulent field that was initialized with the generalized Alfvénic state (2.3) and evolved according to (4.1) and (4.2). Note that $Q$ has been rescaled each time so as to have it range from $-1$ to 1.

For the MHD case, the pertinent quantity for consideration, as per (2.4), would be the electric current density ${\boldsymbol {J}}$. However, thanks to the initialization with the generalized Alfvénic state (2.3), the snapshot of the vorticity would be similar to that of the electric current density (as confirmed also by the direct numerical simulation (DNS) of decaying 2-D MHD by Kinney, McWilliams & Tajima Reference Kinney, McWilliams and Tajima1995). Figure 1 shows that coherent vortices are indeed located in elliptic regions ($Q<0$) while divorticity sheets are located in hyperbolic regions ($Q>0$). The Okubo–Weiss-type parameter given in § 2 is therefore found to track the calculated location of the magnetic-shear-dominated and electric-current-dominated regions for the MHD case very well.

It may be mentioned that similar numerical results were obtained by Banerjee & Pandit (Reference Banerjee and Pandit2014Reference Banerjee and Pandit2019), verifying (2.4) given in our earlier preliminary arXiv preprint (Shivamoggi, van Heijst & Kamp Reference Shivamoggi, van Heijst and Kamp2016).

6 Discussion

The Okubo–Weiss (Okubo Reference Okubo1970; Weiss Reference Weiss1991) criterion provides a very useful diagnostic tool to divide a 2-D hydrodynamical flow field into hyperbolic and elliptic regions. In this paper, we have considered an extension of this device to 2-D MHD flows, and have formulated an Okubo–Weiss-type criterion for 2-D MHD flows. The Okubo–Weiss-type parameter is shown to provide a useful diagnostic tool to parametrize the magnetic field topology in 2-D MHD flows and to provide interesting topological connections with the intrinsic metric properties of the underlying magnetic flux manifold. This is accomplished by recasting the Okubo–Weiss-type criterion by using the 2-D MHD stationary generalized Alfvénic state condition (Hasegawa Reference Hasegawa1985) to approximate the slow-flow-variation ansatz imposed in its derivation. The Okubo–Weiss-type parameter then turns out to be related to the sign definiteness of the Gaussian curvature of the magnetic flux manifold. More specifically, the Okubo–Weiss-type parameter is shown to be the negative of the Gaussian curvature of the magnetic flux surface, to within a positive multiplicative factor. Reformulation of the Okubo–Weiss-type criterion in plane polar coordinates is shown to facilitate its validation for several important real plasma examples like the MHD Rankine-type flux tube and the $Z$-pinch. A similar formulation is shown to become possible for non-standard MHD cases, like 2-D EMHD flows, provided one uses the generalized magnetic flux framework to incorporate the electron-inertia effects (see Appendix A). The Okubo–Weiss-type parameter for the 2-D EMHD case turns out to be related to the intrinsic metric properties of the out-of-plane magnetic field manifold. Numerical simulations of quasi-stationary vortices in 2-D MHD flows in the decaying turbulence regime are then given to demonstrate that the Okubo–Weiss-type criterion given here is shown to be able to separate the flow field into elliptic and hyperbolic field configurations very well.

Acknowledgements

Editor Steve Tobias thanks the referees for their advice in evaluating this article.

Funding

B.K.S. is thankful to Professor J. Rasmussen for helpful discussions and would like to thank The Netherlands Organization for Scientific Research (NWO) for the financial support. The authors are thankful to the referees for their helpful remarks.

Declaration of interests

The authors report no conflict of interest.

Appendix A

The EMHD model is pertinent for collisionless plasmas in space (e.g. solar flares and magnetospheric substorms) and the laboratory (e.g. tokamak discharges). In EMHD, with $\rho _e\ll \ell \ll \rho _i, \rho _s$$s = i, e,$ being the gyro-radius, the dynamics is dominated by magnetized electrons with the demagnetized ions serving to provide the neutralizing static background (Gordeev, Kingsep & Rudakov Reference Gordeev, Kingsep and Rudakov1994). The assumptions underlying the EMHD model are $\ell \ll d_i$, where $d_s \equiv c/\omega _{p_s}$ is the skin depth, and that the frequencies involved are greater than $\omega _{c_i}$ and $\omega _{p_i}$, $\omega _c$ being the cyclotron frequency, $\omega _{c_s} \equiv eB/m_s c$, $\omega _{p_s}$ being the plasma frequency, $\omega _{p_s} \equiv \sqrt {n_se^2 / m_s}$ (Shivamoggi Reference Shivamoggi2015bReference Shivamoggi2016; Shivamoggi & Michalak Reference Shivamoggi and Michalak2019).

The equation governing the transport of the generalized magnetic field or a renormalized magnetic field incorporating the effect of electron inertia (or generalized Ohm's law) may be rewritten as follows:

(A1)\begin{equation} \frac{\partial}{\partial t} \left[ \boldsymbol{\nabla} \times \left( m_e {\boldsymbol{v}}_e - \frac{e {\boldsymbol{A}}}{c} \right) \right] = \boldsymbol{\nabla} \times \left[ {\boldsymbol{v}}_e \times \left\{\boldsymbol{\nabla} \times \left( m_e {\boldsymbol{v}}_e - \frac{e {\boldsymbol{A}}}{c} \right) \right\} \right], \end{equation}

or

(A2)\begin{equation} \frac{{\rm D}{\boldsymbol{B}}_e}{{\rm D} t} \equiv \left[ \frac{\partial}{\partial t} + \left( {\boldsymbol{v}}_e \boldsymbol{\cdot} \boldsymbol{\nabla} \right) \right] {\boldsymbol{B}}_e = \left( {\boldsymbol{B}}_e \boldsymbol{\cdot} \boldsymbol{\nabla} \right) {\boldsymbol{v}}_e, \end{equation}

or

(A3)\begin{equation} \frac{{\rm D}{\boldsymbol{B}}_e}{{\rm D}t} = \boldsymbol{\mathscr{A}}_e \boldsymbol{\cdot} {\boldsymbol{B}}_e, \end{equation}

where

(A4)\begin{equation} {\boldsymbol{v}}_e=\langle v_{e_1},v_{e_2}\rangle \equiv \boldsymbol{\nabla}\times\left(\psi \hat{\boldsymbol{i}}_z\right), \end{equation}

and ${\boldsymbol {v}}_e$ is proportional to the in-plane current density, so $\psi$ also represents the out-of-plane magnetic field. Further

(A5)\begin{equation} \left.\begin{gathered} {\boldsymbol{B}}_e=\langle{\boldsymbol{B}}_{e_1},{\boldsymbol{B}}_{e_2}\rangle \equiv \boldsymbol{\nabla} \times {\boldsymbol{A}}_e \equiv \boldsymbol{\nabla} \times \left({\boldsymbol{A}} - \frac{m_e c}{e} {\boldsymbol{v}}_e \right) = {\boldsymbol{B}} - d_e^2\,\boldsymbol{\nabla}^2 {\boldsymbol{B}},\\ {\boldsymbol{A}}_e =A_e\hat{\boldsymbol{i}}_z,\quad A_e\equiv A-d_e^2\,\nabla^2 A . \end{gathered}\right\} \end{equation}

The electron number density $n_e$ is constant, in accordance with the continuity equation of the electron flow

(A6)\begin{equation} \frac{\partial n_e}{\partial t}=-\boldsymbol{\nabla}\boldsymbol{\cdot}\left( n_e{\boldsymbol{v}}_e\right)= \boldsymbol{\nabla}\boldsymbol{\cdot}\left(\frac{\boldsymbol{J}}{e}\right)= \frac{c}{e}\boldsymbol{\nabla}\boldsymbol{\cdot}\left(\boldsymbol{\nabla}\times{\boldsymbol{B}}\right) =0. \end{equation}

This presupposes that the displacement current $\partial {\boldsymbol {E}}/\partial t$ is negligible, which is valid if $\omega \ll \omega _{p_e}^2/\omega _{c_e}$. Here, $\boldsymbol {\mathscr {A}_e}$ is the electron flow-velocity gradient matrix

(A7)\begin{equation} \boldsymbol{\mathscr{A}_e} \equiv \left[\begin{array}{@{}cc@{}} \displaystyle\dfrac{\partial v_{e_1}}{\partial x} & \displaystyle\dfrac{\partial v_{e_1}}{\partial y}\\[10pt] \displaystyle\dfrac{\partial v_{e_2}}{\partial x} & \displaystyle\dfrac{\partial v_{e_2}}{\partial y} \end{array}\right]\!. \end{equation}

If we now assume, following Okubo (Reference Okubo1970) and Weiss (Reference Weiss1991), that the straining electron-flow velocity gradient tensor $\boldsymbol {\nabla } {\boldsymbol {v}}_e$ temporally evolves slowly so the generalized magnetic field ${\boldsymbol {B}}_e$ evolves adiabatically with respect to the straining electron-flow-velocity gradient field, (A3) may be locally approximated by an eigenvalue problem, with eigenvalues given by

(A8)\begin{equation} \lambda^2 = \left( v_{e_{1_y}} v_{e_{2_x}} + v_{e_{2_y}}^2 \right) \equiv Q. \end{equation}

Using (A4), the Okubo–Weiss-type parameter $Q$ may be recast as follows:

(A9)\begin{equation} Q=\left(\frac{\partial^2\psi}{\partial x\partial y}\right)^2-\frac{\partial^2\psi}{\partial x^2}\frac{\partial^2\psi}{\partial y^2}, \end{equation}

which implies that the Okubo–Weiss-type parameter $Q$ is related to the intrinsic metric properties of the out-of-plane magnetic field manifold – it is the negative of the Gaussian curvature of the latter manifold, to within a positive multiplicative factor.

Let us approximate the slow electron-flow-variation ansatz used above by the Beltrami flow condition for the EMHD flows (Shivamoggi Reference Shivamoggi2011)

(A10)\begin{equation} a{\boldsymbol{v}}_e = {\boldsymbol{B}}_e, \end{equation}

which follows from (A1), and represents the generalized Alfvénic state in 2-D EMHD. (There is no force-free state in 2-D EMHD, except in the massless-electron limit (Shivamoggi Reference Shivamoggi2015a). Taking $a$ to be a constant (A8) becomes

(A11)\begin{equation} Q=\displaystyle\frac{1}{a^2}\left( B_{e_1y} B_{e_2x}+B^2_{e_2y}\right) =\displaystyle\frac{1}{a^2}\left( b^2_{e_1}+b^2_{e_2}-\displaystyle\frac{1}{c^2}J^2_e\right), \end{equation}

where

(A12a–c)\begin{equation} b_{e_1}\equiv-2B_{e_2y},\quad b_{e_2}\equiv B_{e_2x}+B_{e_1y}, \quad\frac{1}{c}J_e\equiv B_{e_2x}-B_{e_1y}=\frac{1}{c}\left( J-d^2_e\nabla^2J\right)\!. \end{equation}

So, the Okubo–Weiss-type parameter $Q$ is a measure of the relative importance of the generalized in-plane magnetic shear ($Q>0,$ hyperbolic) and the generalized out-of-plane electric current ($Q<0$, elliptic).

In terms of the generalized magnetic vector potential $A_e$ given by (A5)

(A13)\begin{equation} Q = \frac{1}{a^2} \left[ \left( \frac{\partial^2 A_e}{\partial x \partial y} \right)^2 - \frac{\partial^2 A_e}{\partial x^2} \frac{\partial^2 A_e}{\partial y^2} \right]\!. \end{equation}

Equation (A13) implies that the Okubo–Weiss-type parameter $Q$ for the 2-D EMHD case is the negative of the Gaussian curvature of the generalized magnetic flux surface, to within a positive multiplicative factor. As with the case of 2-D MHD flows, (A13) can serve as a useful diagnostic tool to parameterize the generalized in-plane magnetic field topology in 2-D EMHD flows in terms of the relative dominance of generalized in-plane magnetic shear ($Q>0$, hyperbolic) and generalized out-of-plane electric current ($Q<0$, elliptic). This result implies that the eigenvalues of the generalized magnetic field gradient tensor $\boldsymbol {\nabla } {\boldsymbol {B}}_e$ become purely imaginary near the $O$-points or real near the $X$-points of the generalized magnetic field lines, generalizing Greene's (Greene Reference Greene1993) result to EMHD.

Equation (A13) may be rewritten, on noting (A5), as

(A14)\begin{equation} Q = \frac{1}{a^2} \left|\begin{array}{@{}cc@{}} \displaystyle\dfrac{\partial^2 A_e}{\partial x \partial y} & \displaystyle\dfrac{\partial^2 A_e}{\partial x^2}\\[10pt] \displaystyle\dfrac{\partial^2 A_e}{\partial y^2} & \displaystyle\dfrac{\partial^2 A_e}{\partial x \partial y} \end{array}\right| = Q_m + Q_e + Q_c, \end{equation}

where

(A15a,b)\begin{gather} Q_m \equiv \frac{1}{a^2} \left|\begin{array}{@{}cc@{}} \displaystyle\dfrac{\partial^2 A}{\partial x \partial y} & \displaystyle\dfrac{\partial^2 A}{\partial x^2}\\[10pt] \displaystyle\dfrac{\partial^2 A}{\partial y^2} & \displaystyle\dfrac{\partial^2 A}{\partial x \partial y} \end{array}\right|\!,\quad Q_e \equiv \frac{d_e ^4}{ a^2} \left|\begin{array}{@{}cc@{}} \displaystyle\dfrac{\partial^2}{\partial x \partial y} ( \nabla^2 A ) & \displaystyle\dfrac{\partial^2}{\partial x^2} ( \nabla^2 A ) \\[10pt] \displaystyle\dfrac{\partial^2}{\partial y^2} ( \nabla^2 A ) & \displaystyle\dfrac{\partial^2}{\partial x \partial y} ( \nabla^2 A ) \end{array}\right|\!, \end{gather}
(A16)\begin{gather} Q_c \equiv- \frac{d_e ^2}{ a^2} \left\{ \left| \begin{array}{@{}cc@{}} \displaystyle\dfrac{\partial^2 A}{\partial x \partial y} & \displaystyle\dfrac{\partial^2 A}{\partial x^2} \\[10pt] \displaystyle\dfrac{\partial^2}{\partial y^2} ( \nabla^2 A ) & \displaystyle\dfrac{\partial^2}{\partial x \partial y} ( \nabla^2 A ) \end{array}\right| +\left|\begin{array}{@{}cc@{}} \displaystyle\dfrac{\partial^2 A}{\partial x \partial y} & \displaystyle\dfrac{\partial^2 A}{\partial y^2} \\[10pt] \displaystyle\dfrac{\partial^2}{\partial x^2} ( \nabla^2 A ) & \displaystyle\dfrac{\partial^2}{\partial x \partial y} ( \nabla^2 A ) \end{array}\right|\right\}\!. \end{gather}

So, the Okubo–Weiss-type parameter $Q$ for 2-D EMHD may be viewed as the sum of the negative of the Gaussian curvatures of the in-plane magnetic flux surface ($Q_m$) and the out-of-plane electron-current manifold ($Q_e$) along with cross-terms ($Q_c$). Here, $Q_e$ and $Q_c$ both represent electron-inertia contributions. This result is commensurate with the topological result for EMHD (Shivamoggi & Michalak Reference Shivamoggi and Michalak2019) that the self-linkage of electron-flow vorticity field lines in EMHD (which is an invariant) can be expressed as the sum of the self-linkage of magnetic field lines, the self-linkage of electron-flow vorticity field lines and the mutual linkage among these two sets of field lines.

If the electron-inertia effects are ignored (i.e. in the limit $d_e \Rightarrow 0$) $Q_e$ and $Q_c$ drop out and, (A13) reduces to the result (2.7) for classical MHD, given in § 2.

Footnotes

1 A similar ansatz was also invoked in Toda's criterion (Toda Reference Toda1974), which is aimed at predicting the onset of chaotic motion by investigating the linear stability of trajectories in the neighbourhood of the reference trajectory. The former are assumed to evolve adiabatically with respect to the latter.

2 For the exact nonlinear Alfvénic state associated with perturbations in a uniformly magnetized plasma considered by Hasegawa (Reference Hasegawa1985), $a = \pm \sqrt {\rho }$, $\rho$ being the mass density of the incompressible plasma.

3 This is similar to the way $Q$ measures the relative importance of flow strain and flow vorticity in hydrodynamics (Okubo Reference Okubo1970; Weiss Reference Weiss1991). However, it may be pointed out that, due to the reported apparent non-robust nature of the Alfvénic state in MHD (Núñez Reference Núñez2007) (which may be traceable to the dynamical significance of electric current $J$ unlike vorticity $\omega$, although $\omega$, like $J$, exhibits self-degradation as the two systems evolve), the electromagnetic interpretation of $Q$ in MHD (as in (2.4)) may not be on a totally firm footing.

4 In a $Z$-pinch, the azimuthal magnetic field is generated by an axial electric current in the plasma, which in turn, compresses and confines this magnetic field.

References

Babiano, A., Boffetta, G., Provenzale, A. & Vulpiani, A. 1994 Chaotic advection in point vortex models and two-dimensional turbulence. Phys. Fluids 6 (7), 24652474.CrossRefGoogle Scholar
Babiano, A. & Provenzale, A. 2007 Coherent vortices and tracer cascades in two-dimensional turbulence. J. Fluid Mech. 574, 429448.CrossRefGoogle Scholar
Banerjee, D. & Pandit, R. 2014 Statistics of the inverse-cascade regime in two-dimensional magnetohydrodynamic turbulence. Phys. Rev. E 90 (1), 013018.CrossRefGoogle ScholarPubMed
Banerjee, D. & Pandit, R. 2019 Two-dimensional magnetohydrodynamic turbulence with large and small energy-injection length scales. Phys. Fluids 31 (6), 065111.CrossRefGoogle Scholar
Brachet, M.E., Meneguzzi, M., Politano, H. & Sulem, P.L. 1988 The dynamics of freely decaying two-dimensional turbulence. J. Fluid Mech. 194, 333349.CrossRefGoogle Scholar
Elhmaïdi, D., Provenzale, A. & Babiano, A. 1993 Elementary topology of two-dimensional turbulence from a Lagrangian viewpoint and single-particle dispersion. J. Fluid Mech. 257, 533558.CrossRefGoogle Scholar
Giovanelli, R.G. 1949 XVII. Electron energies resulting from an electric field in a highly ionized gas. Lond. Edinb. Dub. Phil. Mag. J. Sci. 40 (301), 206214.CrossRefGoogle Scholar
Goedbloed, J.P.H. & Poedts, S. 2004 Principles of Magnetohydrodynamics: With Applications to Laboratory and Astrophysical Plasmas. Cambridge University Press.CrossRefGoogle Scholar
Gordeev, A.V., Kingsep, A.S. & Rudakov, L.I. 1994 Electron magnetohydrodynamics. Phys. Rep. 243 (5), 215315.CrossRefGoogle Scholar
Greene, J.M. 1993 Reconnection of vorticity lines and magnetic lines. Phys. Fluids B 5 (7), 23552362.CrossRefGoogle Scholar
Hasegawa, A. 1985 Self-organization processes in continuous media. Adv. Phys. 34 (1), 142.CrossRefGoogle Scholar
Kinney, R., McWilliams, J.C. & Tajima, T. 1995 Coherent structures and turbulent cascades in two-dimensional incompressible magnetohydrodynamic turbulence. Phys. Plasmas 2 (10), 36233639.CrossRefGoogle Scholar
Kuznetsov, E.A., Naulin, V., Nielsen, A.H. & Rasmussen, J.J. 2007 Effects of sharp vorticity gradients in two-dimensional hydrodynamic turbulence. Phys. Fluids 19 (10), 105110.CrossRefGoogle Scholar
Mcwilliams, J.C. 1984 The emergence of isolated coherent vortices in turbulent flow. J. Fluid Mech. 146, 2143.CrossRefGoogle Scholar
Núñez, M. 2007 Long time convergence of magnetohydrodynamic flows to alfvénic states. J. Plasma Phys. 73 (6), 947955.CrossRefGoogle Scholar
Ohkitani, K. 1991 Wave number space dynamics of enstrophy cascade in a forced two-dimensional turbulence. Phys. Fluids A 3 (6), 15981611.CrossRefGoogle Scholar
Okubo, A. 1970 Horizontal dispersion of floatable particles in the vicinity of velocity singularities such as convergences. Deep-Sea Res. 17 (3), 445454.Google Scholar
Ouellette, N.T. & Gollub, J.P. 2007 Curvature fields, topology, and the dynamics of spatiotemporal chaos. Phys. Rev. Lett. 99 (19), 194502.CrossRefGoogle ScholarPubMed
Rollins, D.K. & Shivamoggi, B.K. 2007 Current sheet formation near a hyperbolic magnetic neutral line in a variable-density plasma: an exact solution. Phys. Lett. A 366 (1), 97100.CrossRefGoogle Scholar
Shivamoggi, B.K. 1986 Evolution of current sheets near a hyperbolic magnetic neutral point. Phys. Fluids 29 (3), 769772.CrossRefGoogle Scholar
Shivamoggi, B.K. 1999 Current-sheet formation near a hyperbolic magnetic neutral line in the presence of a plasma flow with a uniform shear-strain rate: an exact solution. Phys. Lett. A 258 (2), 131134.CrossRefGoogle Scholar
Shivamoggi, B.K. 2011 Characteristics of plasma Beltrami states. Eur. Phys. J. D 64 (2), 393404.CrossRefGoogle Scholar
Shivamoggi, B.K. 2015 a Beltrami states in 2D electron magnetohydrodynamics. arXiv:1506.06094.Google Scholar
Shivamoggi, B.K. 2015 b Electron magnetohydrodynamic turbulence: universal features. Eur. Phys. J. D 69 (2), 55.CrossRefGoogle Scholar
Shivamoggi, B.K. 2016 Plasma relaxation and topological aspects in electron magnetohydrodynamics. Phys. Plasmas 23 (7), 072302.CrossRefGoogle Scholar
Shivamoggi, B.K. & Michalak, M. 2019 Topological implications of the total generalized electron-flow magnetic helicity invariant in electron magnetohydrodynamics. Phys. Plasmas 26 (4), 44501.CrossRefGoogle Scholar
Shivamoggi, B.K., van Heijst, G.J.F. & Kamp, L.P.J. 2016 The Okubo–Weiss criteria in two-dimensional hydrodynamic and magnetohydrodynamic flows. arXiv:1110.6190.Google Scholar
Shivamoggi, B.K., Van Heijst, G.J.F. & Kamp, L.P.J. 2022 The Okubo–Weiss criterion in hydrodynamic flows: geometric aspects and further extension. Fluid Dyn. Res. 54 (1), 015505.CrossRefGoogle Scholar
Toda, M. 1974 Instability of trajectories of the lattice with cubic nonlinearity. Phys. Lett. A 48 (5), 335336.CrossRefGoogle Scholar
Weiss, J. 1991 The dynamics of enstrophy transfer in two-dimensional hydrodynamics. Physica D 48 (2), 273294.CrossRefGoogle Scholar
Figure 0

Figure 1. Snapshot of the vorticity field (a) and the associated Okubo–Weiss function (b) taken during the decaying phase of MHD turbulence. The colour bar denotes the value of the Okubo–Weiss function, which has been rescaled to range from $-1$ to $+1$.