Hostname: page-component-848d4c4894-pjpqr Total loading time: 0 Render date: 2024-06-18T23:08:22.565Z Has data issue: false hasContentIssue false

Microscopic scales of linear tearing modes: a tutorial on boundary layer theory for magnetic reconnection

Published online by Cambridge University Press:  23 November 2022

H. Betar*
Affiliation:
Laboratoire M2P2, UMR CNRS 7340 – Université Aix-Marseille, F-13451 Marseille, France Institut Jean Lamour, UMR 7198 CNRS – Université de Lorraine, F-54000 Nancy, France
D. Del Sarto*
Affiliation:
Institut Jean Lamour, UMR 7198 CNRS – Université de Lorraine, F-54000 Nancy, France
M. Ottaviani
Affiliation:
CEA, IRFM, F-13108, Saint-Paul-lez-Durance, France
A. Ghizzo
Affiliation:
Institut Jean Lamour, UMR 7198 CNRS – Université de Lorraine, F-54000 Nancy, France
*
Email addresses for correspondence: homam.betar@univ-amu.fr, daniele.del-sarto@univ-lorraine.fr
Email addresses for correspondence: homam.betar@univ-amu.fr, daniele.del-sarto@univ-lorraine.fr
Rights & Permissions [Opens in a new window]

Abstract

We revise in detail and in a pedagogical way the analysis of the boundary layer theory of warm tearing modes in slab, reduced magnetohydrodynamics (MHD), when magnetic reconnection is driven by electron inertia and/or resistivity, and ion-sound Larmor radius effects are included. By comparison with the numerical solution of the corresponding eigenvalue problem, we interpret these results by means of a heuristic approach, which in the warm-electron regime, we show to be in general not feasible without knowledge of the scaling of the gradient of the magnetic flux function, differently from what happens in the cold-electron regimes. We put in evidence for a non-trivial relation between the first derivative of the magnetic flux function and of the velocity parallel to the neutral line, evaluated in its proximity, by thus providing insight to the multiple boundary layer analysis that Pegoraro & Schep (Plasma Phys. Control. Fusion, vol. 28, 1986, p. 647) first showed to be required in warm-tearing regimes. In this way, we also suggest and justify a general operational definition of the reconnecting layer width and we discuss the linear appearance of microscopic scales related to the gradients of the eigenfunctions of the tearing modes.

Type
Tutorial
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2022. Published by Cambridge University Press

1 Introduction

Magnetic reconnection is one of the chief processes of conversion between electromagnetic energy and particle kinetic energy in magnetised plasmas (see, e.g. Biskamp Reference Biskamp2000; Yamada, Kulsrud & Ji Reference Yamada, Kulsrud and Ji2010). At the basis of extreme energy releasing phenomena that naturally occur in space plasmas, such as solar flares, coronal mass ejections or magnetic substorms, magnetic reconnection is of fundamental importance also in magnetically confined thermonuclear fusion plasma experiments: in fusion devices like tokamaks, it can cause disruptions (see, e.g. Wesson Reference Wesson1990), that is, the sudden loss of the magnetic confinement. This is usually due to magnetic island formation via tearing-like instabilities, and can affect the transport of matter and energy. Among the different scenarios, which have been devised since the first formulations of the ‘magnetic reconnection’ concept in the works by Giovanelli (Reference Giovanelli1946), Hoyle (Reference Hoyle1949) and Dungey (Reference Dungey1950), tearing-type modes (Furth, Killeen & Rosenbluth Reference Furth, Killeen and Rosenbluth1963) are the prototypical example of spontaneous reconnecting instabilities, that is, of magnetic perturbations that grow exponentially in time by inducing magnetic reconnection. Their ‘explosive’ behaviour in time make them the chief candidates for the explanation of several fast energy releasing events, which, both in Nature and in laboratory plasmas, are attributed to magnetic reconnection processes. Their theoretical modelling is however not trivial, since the mathematics required to solve the corresponding eigenvalue problem is complicated by some technical features that make its integration typically more complex than that of most linear instabilities. Notably, due to the multi-scale nature of the linear problem in which linear differential operators intervene, a boundary layer integration procedure is required.

Since Furth's remark (Furth Reference Furth1963, Reference Furth1964) of the relevance of electron inertia in allowing magnetic reconnection and since the early model by Furth (Reference Furth1964) and Coppi (Reference Coppi1964c,Reference Coppia) for inertia-driven reconnecting instabilities, kinetic scale effects have attracted increasing attention in the attempt to model the rapid magnetic reconnection phenomena observed in low collisionality plasmas. After the seminal paper by Furth et al. (Reference Furth, Killeen and Rosenbluth1963), in which the theory of resistive tearing modes was first formulated, several attempts have been made to include collisionless and/or kinetic physics in the linear tearing mode theory. Even if some works exist in which a full kinetic or girokinetic analytical treatment has been considered (Coppi, Laval & Pellat Reference Coppi, Laval and Pellat1966; Laval, Pellat & Vuillemin Reference Laval, Pellat and Vuillemin1966; Hazeltine & Ross Reference Hazeltine and Ross1975; Drake & Lee Reference Drake and Lee1977; Cowley, Kulsrud & Hahm Reference Cowley, Kulsrud and Hahm1986; Daughton Reference Daughton1999; Daughton & Karimabadi Reference Daughton and Karimabadi2005; Zocco & Schekochihin Reference Zocco and Schekochihin2011; Connor, Hastie & Zocco Reference Connor, Hastie and Zocco2012b), the complexity of the boundary layer analysis and of the heuristic assumptions on the ordering of the different microscopic scales makes the identification of the different reconnection regimes a quite difficult task, especially when two or more non-ideal magnetohydrodynamic (MHD) parameters enter in the reconnection rate. This is true even in a relatively simple fluid description extended to include non-ideal MHD effects.

For this reason, heuristic approaches have been developed to tackle the boundary layer problem with a simplified dimensional-like analysis, based on estimations about the characteristic gradients and about the balance of the terms of the linear equations. These techniques have been successfully used to recover the asymptotic scalings of the growth rate and of the reconnection layer width in the purely resistive and purely inertial regimes of reduced MHD (RMHD) instabilities. They are also often presented in textbooks on magnetic reconnection as a ‘short-cut’ procedure to obtain these results without carrying out the full boundary layer integration of the eigenmodes. Providing an estimate of the characteristic temporal and spatial scales of spontaneous reconnecting instabilities, and comparing them with the values inferred from experimental measures of reconnection events occurring in the laboratory or in Nature, are indeed among the elements of principal interest, in this context. Heuristic methods based on dimensional analysis have also proven to work for both small and large values of the $\varDelta '$ instability parameter (Ottaviani & Porcelli Reference Ottaviani and Porcelli1995) and for the fastest growing mode in a large aspect ratio current sheet (Bhattacharjee et al. Reference Bhattacharjee, Huang, Yang and Rogers2009; Comisso et al. Reference Comisso, Grasso, Waelbroeck and Borgogno2013; Del Sarto et al. Reference Del Sarto, Pucci, Tenerani and Velli2016). Moreover, they are generally used to obtain insight on the physics of the problem and on the interpretation of some non-trivial results of the boundary layer analysis (see, e.g. Drake & Lee Reference Drake and Lee1977; Cowley et al. Reference Cowley, Kulsrud and Hahm1986). However, as we are going to discuss in this work, heuristic methods ‘fail’ to obtain the scalings first computed by Pegoraro & Schep (Reference Pegoraro and Schep1986) and Porcelli (Reference Porcelli1991) with a boundary layer approach, when electron temperature effects are included, unless some further careful assumptions are made, which are related to the boundary layer decomposition of the spatial domain: whether it is possible to develop heuristc arguments which allow one to obtain these results without relying on further information from boundary layer analysis (or from numerical integration of the linear problem) is, at the moment, an open question.

One of the main purposes of this work is to revise in a tutorial way the linear analysis of tearing-type modes in these regimes, by discussing their analytical solutions in the coordinate space and by showing in this way the advantages and limitations of heuristic-type derivations. To the best of our knowledge, indeed, details of the analytical theory for this double-boundary layer approach have been seldom discussed in the existing literature. When this has occurred, the analysis was based on the solution of the eigenvalue problem in the Fourier space (Pegoraro & Schep Reference Pegoraro and Schep1986), or by using a different linear model based on a reduced fluid-kinetic approach (Zocco & Schekochihin Reference Zocco and Schekochihin2011) and by taking a somewhat different analytical approach based on perturbation methods to solve the boundary layer equations that here we will tackle by direct integration instead.Footnote 1 It should be nevertheless recalled that other analytical works on this subject exist. In these, the boundary layer analysis of tearing modes is however complicated by further ingredients related, for example, to the geometry of the magnetic equilibrium profile (see, e.g. Militello et al. Reference Militello, Huysmans, Ottaviani and Porcelli2004; Connor et al. Reference Connor, Hastie, Marchetto and Roach2012a; Zocco, Helander & Weitzner Reference Zocco, Helander and Weitzner2020) or to the inclusion of full kinetic effects (see the references previously cited in this regard). Moreover, also in the few works that addressed these subjects with different approaches with respect to the one we develop here, most details of the complex analysis involved in the calculations were often not reported and are difficult to track.

To simplify the analysis and to show in a pedagogical way the essential features of the double boundary layer separation, we focus here on the limit of cold ions, which allows us to take a fluid closure for the latter, differently from the semi-kinetic models considered in practically all previous works that already treated this problem (see, e.g. Cowley et al. Reference Cowley, Kulsrud and Hahm1986; Pegoraro & Schep Reference Pegoraro and Schep1986; Pegoraro, Porcelli & Schep Reference Pegoraro, Porcelli and Schep1989; Porcelli Reference Porcelli1991; Zocco & Schekochihin Reference Zocco and Schekochihin2011). In particular, we rely on the equations of the two-fluid reduced-MHD model extended to include ion-sound Larmor radius effects, for which different derivations exist (Zank & Matthaeus Reference Zank and Matthaeus1992; Schep, Pegoraro & Kuvshinov Reference Schep, Pegoraro and Kuvshinov1994; Bergmans Reference Bergmans2001; Del Sarto, Califano & Pegoraro Reference Del Sarto, Califano and Pegoraro2006; Bian & Tsiklauri Reference Bian and Tsiklauri2009), which are based on two different types of ordering between fluctuations of the ion density and of the guide field magnetic component (cf. Appendix A). Several more recent and refined gyrofluid models exist, which are based on the evolution of more than two scalar fields. However, the two-field model we focus on, which has been used in several nonlinear studies of different magnetic reconnection scenarios (see Kleva, Drake & Waelbroeck Reference Kleva, Drake and Waelbroeck1995; Cafaro et al. Reference Cafaro, Grasso, Pegoraro, Porcelli and Saluzzi1998; Grasso et al. Reference Grasso, Pegoraro, Porcelli and Califano1999; Bergmans & Schep Reference Bergmans and Schep2001; Del Sarto, Califano & Pegoraro Reference Del Sarto, Califano and Pegoraro2003; Wang et al. Reference Wang, Wei, Wang, Zheng and Liu2011 to cite the earliest) or even of turbulence (Milosevich, Morrison & Tassi Reference Milosevich, Morrison and Tassi2018), contains the whole essential physics of the problem. More specifically, it yields the same linear system to which all other, more refined, cold-ion models converge in the isothermal electron limit.

To accomplish our pedagogical purpose, we then give a step-by-step presentation of the boundary layer integration procedure in the coordinate space, with the aim of keeping it ‘self-contained’, that is, by trying to provide all the analytical tools useful for the purpose (e.g. element of complex analysis), when they result necessary for the algebra. By then comparing the analytical results with those obtained with a numerical eigen-solver (Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) based on a multi-precision toolbox (Holoborodko Reference Holoborodko2012), we analyse the spatial behaviour of the corresponding eigenfunctions and we discuss the limitations and delicate points of heuristic estimations that are sometimes used as a ‘quicker’ alternative to full boundary layer calculations. Although the main focus of this article is on the warm-electron regime, we also consider, for comparison, the cold-electron limit, in which a single boundary layer analysis suffices. In particular, we consider in a unified way both the resistive and collisionless regimes, and we separately address the cold- and warm-electron regimes. Each regime is associated with some characteristic non-ideal parameter: the Lundquist number $S^{-1}$, related to resistivity; the electron skin depth $d_e$, related to electron inertia in the collisionless limit; and the ion-sound Larmor radius $\rho _s$, related to the electron temperature.

We thus discuss the results of the boundary layer analysis, which yields the asymptotic scaling (in terms of the non-ideal parameters ruling the reconnection process) of the growth rates and of some characteristic spatial scales associated with the eigenfunctions, as well as the approximated profile of the latter in some regions of the domain of integration. We then discuss the asymptotic estimates of the first derivatives of the current and velocity field on the neutral line. We also provide (to the best of our knowledge, for the first time) a formal quantitative definition of the reconnection layer width, which we argue to be valid in all reconnection regimes and which is given in terms of numerically measurable quantities that are related to the spatial profile of the eigenfunctions.

Then, in interpreting these quantities in terms of a heuristic approach, whose limitations in the warm-electron case we point out, we provide evidence that the introduction of the scale $\rho _s$ makes a further characteristic ‘mixed’ scale appear in the collisionless large wavelength limit, which is smaller than $d_e$ when $\rho _s>d_e$, since it scales like ${\sim }\rho _s^{-1/3}d_e^{4/3}$. Such a characteristic length, which in previous works (Porcelli Reference Porcelli1991) had been already noted, is identifiable as the reconnecting layer width, in agreement both with the operational definition here proposed and with previous works that already recognised it as such (Zocco & Schekochihin Reference Zocco and Schekochihin2011).

By discussing the boundary layer results in light of a heuristic dimensional-type approach, we also show a new, non-trivial relation between the first derivative of the magnetic flux function and the first derivative of the parallel velocity, which holds close to the neutral line. This velocity gradient displays characteristic scalings that depend on the wavelength regime. Information about these scalings is a priori not evident, and the analysis we provide on this subject at the end of this work suggests indeed that such information can not be obtained via simple dimensional analysis, if not in the cold-electron regimes. Nevertheless, introducing this scale length allows, in any reconnection regime, the writing of the scaling laws in a way that results to be perfectly symmetric between the small and large wavelength limit, provided the characteristic instability parameter of the small wavelength limit (i.e. the $\varDelta '$ parameter of Furth et al. Reference Furth, Killeen and Rosenbluth1963) is replaced in the large wavelength limit by the gradient of the velocity (which we have here called $\varDelta _{v_y}'$, by analogy). This fact, and the correspondence between some characteristic scale length associated with this velocity gradient and scale lengths that in other works based on a kinetic approach have been interpreted in terms of inherently kinetic features (Drake & Lee Reference Drake and Lee1977; Cowley et al. Reference Cowley, Kulsrud and Hahm1986; Ottaviani & Porcelli Reference Ottaviani and Porcelli1993; Zocco & Schekochihin Reference Zocco and Schekochihin2011), suggests that the role of this quantity in the tearing mode linear dynamics deserves future investigation.

The structure of this article is as follows.

  1. (i) In § 2, we discuss the model equations and some general aspects of magnetic reconnection which include: a brief historical review of the notion of magnetic reconnection associated with that of magnetic topology and some general features of tearing modes and of the magnetic structures they induce (§ 2.1); the notion of reconnection rate and its relation with the growth rate of the eigenvalue problem (§ 2.2).

  2. (ii) In § 3, we introduce the notion of boundary layer, we discuss its relevance to tearing mode analysis and we outline the key points of the corresponding integration strategy. Then we introduce the notion of instability parameter ($\varDelta '$) and the way it intervenes in the matching of the solutions between the ideal and non-ideal regions of the domain (§ 3.1). We discuss the wavelength regimes of the eigenmode solution, in terms of the amplitude of $\varDelta '$, by pointing out similarities and differences between tearing modes in slab and in cylindrical geometry (§ 3.2). In § 3.3, we briefly review previous works in which a boundary layer analysis similar to the one we develop here has been discussed.

  3. (iii) In § 4, we address the boundary layer integration by starting from the ideal region, where the hypotheses of ideal MHD hold, and we detail the integration procedure which allows the analytical evaluation of $\varDelta '(k)$ in terms of the instability wavenumber $k$ for a specific equilibrium profile, which exemplifies a larger class of equilibria (§ 4.1).

  4. (iv) In § 5, we introduce the notions of ‘generalised’ resistivity and electron inertia, which allow a unified treatment of the resistive and collisionless case altogether, and we discuss the combination of the two non-ideal parameters. We then discuss the strategy for the identification of the integration layers in the non-ideal region via the general approximations valid in the non-ideal region (§ 5.1) and the criteria of normalisation (§ 5.2). We then introduce the auxiliary function, useful for the integration of the boundary layer problem in some regimes, by discussing its small and large wavelength limits (§ 5.3) and then its normalisation (§ 5.4). Finally, we outline the criteria that may make one prefer to perform the integration of the non-ideal equation by using the auxiliary equations rather than the equations for the two scalar fields (§ 5.5).

  5. (v) In § 6, we solve the boundary layer problem in the cold-electron regimes. First we find the solution in the large-$\varDelta '$ limit (§ 6.1) and then in the small-$\varDelta '$ limit (§ 6.2).

  6. (vi) In § 7, we address the warm-electron regimes: after discussing the general form of the equations in the non-ideal region (§ 7.1), we identify the two boundary layers of interest in this case (§ 7.2) and then we outline the integration strategy that will be pursued, by relying on the integral representation of hypergeometric functions (§ 7.3). First we consider the solution in the large-$\varDelta '$ (§ 7.4) and then in the small-$\varDelta '$ limit (§ 7.5).

  7. (vii) In § 8, we address the problem of identifying the characteristic width of the reconnecting layer by relying on further hypotheses of physical character and by starting from the characteristic scales obtained from the boundary layer integration. After reviewing different notions of the reconnecting layer that have been adopted in the literature over the years (§ 8.1), we propose an operational definition of its width, related to the distance of the local maxima (or minima) of the current density from the neutral line, which can be useful for both experimental and numerical application, and which is shown via numerical integration to provide the same scalings of the inner layer width (§ 8.2). The asymptotic scalings of further characteristic spatial scales associated with the derivatives of the magnetic field and of the velocity profile on the neutral line are then estimated and compared to the scalings of the reconnecting layer width in different regimes (§ 8.3).

  8. (viii) In § 9, we address the problem of the heuristic derivation of the scaling laws of tearing modes via dimensional analysis. After discussing the relevance and usefulness of the approach and after having outlined its general hypotheses (§ 9.1), we apply it to the ‘textbook’ example of the cold-electron regimes, where its efficiency is well established (§ 9.3). Then we show and discuss its ‘failure’ in the warm-electron regimes, where it yields wrong estimates with respect to those obtained from the boundary layer analysis (§ 9.3).

  9. (ix) In § 10, starting from the physical insight brought by the analysis of the logical points which may lead to the wrong dimensional estimates in the warm-electron regime, we introduce a new characteristic scale length associated with the gradient of the velocity component parallel to the neutral line and evaluated close to it: by analogy with the instability parameter, we call it $\varDelta _{v_y}'$ (§ 10.1). We then postulate a further estimate for the gradient of the magnetic flux function, which, depending on the reconnection regime, can be related to $\varDelta _{v_y}'$ (§ 10.2). After having shown the convergence of a numerical procedure which allows the quantification of $\varDelta _{v_y}'$ (§ 10.3), we show its relevance for the heuristic estimates: they can now lead to the correct results – although in the warm-electron regime, the procedure is not ‘closed’, since it is strongly suggested that the scale $\varDelta _{v_y}'$ can not be obtained by simple dimensional analysis – and they can be cast in a ‘symmetric form’, in which the scalings in the large-$\varDelta '$ limits mirror those of the small-$\varDelta '$ limits provided the substitution $\varDelta '\leftrightarrow \varDelta _{v_y}'$ (§ 10.4). The significance of the inverse scale length $\varDelta _{v_y}'$ is further discussed by comparison with results obtained in previous works on boundary layer calculations, in which a characteristic length displaying similar scalings in some regimes had been already noted (§ 10.5).

  10. (x) Conclusions follow in § 11.

  11. (xi) Further technical details are reported in the Appendices. These include: calculations related to the derivation of the model and a discussion of the relevant hypotheses (Appendix A); a brief discussion of alternative definitions found in literature for the notion of ‘reconnection rate’, which are not directly relevant to the problem we consider here (Appendix B); a discussion on the general strategy of integration of the boundary layer problem of tearing modes in the Fourier space, and a comparison of the equations we have integrated with those appearing in previous works available in the literature (Appendix C); a didactical example of normalisation of a differential equation (Appendix D); the proof of convergence and linear independence of the solution found in § 7.4 (Appendix E); and the detailed discussion of the logical steps, which allow one to generalise the heuristic estimate discussed in § 10.4 in terms of $\varDelta _{v_y}'$ (Appendix F).

2 Model equations and linear problem: general features

We consider the reduced MHD limit, from now on noted as RMHD (Strauss Reference Strauss1976, Reference Strauss1977; Zank & Matthaeus Reference Zank and Matthaeus1992), and we consider the standard tearing equations in slab geometry ($\partial /\partial z=0$) for a strongly magnetised plasma whose guide field component is along the $z$ direction. Fluid incompressibility, valid for the bulk plasma at the leading ${\boldsymbol E}\times {\boldsymbol B}$-drift ordering and possibly with inclusion of drift-diamagnetic corrections (see Appendix A), allows us to reduce the number of independent vector components and therefore to consider a limited set of scalar quantities.

Choosing perturbations of the form ${\sim }f(x){\rm e}^{{\rm i}ky+\gamma t}{+{\rm c.c.}}$ we linearise the equations around equilibria with an odd $B_y^0(x)$ component, which vanishes at $x=0$, and with a null in-plane fluid velocity at equilibrium, ${\boldsymbol U}^0(x,y)=0$. Here we have already used knowledge of the fact that the eigenmodes we are considering have low frequencies, so that $\omega /\gamma \sim 0$. This can be taken here as a verifiable heuristic assumption.

It can be shown that only two scalar fields are necessary. The relevant equations can be then represented by the equation for one of the in-plane components of ${\boldsymbol B}$ and by the equation for one of the in-plane components of the fluid velocity. These can be cast in the form of two coupled equations for the stream functions $\psi$ and $\varphi$. Details about all these features, and bibliographical references as well, are provided in Appendix A.

In particular, $\psi (x,y,t)$ and $b(x,y,t)$ are the normalised scalar functions defining the magnetic field components, $\boldsymbol {B}=\boldsymbol {\nabla }\psi \times \boldsymbol {e}_z+b\boldsymbol {e}_z$, whereas $\varphi (x,y,t)$ is the stream function for the in-plane $\boldsymbol {E}\times \boldsymbol {B}$-drift velocity, $\boldsymbol {U}_\perp =-\boldsymbol {\nabla }\varphi \times \boldsymbol {e}_z$, expressing the normalised gradient of the electrostatic potential in RMHD. Here, the large guide field ordering allows us to completely neglect the dynamics related to the scalar function $b$.

Labelling with indices ‘$0$’ and ‘$1$’ the equilibrium quantities and first-order perturbations, respectively, the linearisation is then performed around equilibria of the form $\psi (x,y,0)\equiv \psi _0(x)$, even with respect to $x=0$, with $\phi (x,y,0)=0$. Normalised as indicated below and using, for simplicity of notation, the same symbols for the fields defined in the coordinate space and for their Fourier transform with respect to the $y$ variable, the equations we will work with, read:

(2.1)\begin{gather} \gamma [\psi_1-{d}_e^2(\psi_1''-k^2\psi_1)] -{\rm i}k\varphi_1 (\psi_0'-{d}_e^2\psi_0''')={-}{\rm i}k{\rho}_s^2(\varphi_1''-k^2\varphi_1)\psi_0'+S^{{-}1}(\psi_1''-k^2\psi_1), \end{gather}
(2.2)\begin{gather} \gamma [\varphi_1''-k^2\varphi_1]={\rm i}k\psi_0'(\psi_1''-k^2\psi_1)-ik\psi_1 \psi_0'''. \end{gather}

We can also cast (2.1)(2.2) in a matrix form (useful for the eigenvalue solution) as

(2.3)\begin{equation} \gamma \left(\begin{array}{@{}cc@{}} \mathcal{F} & 0 \\ 0 & \mathcal{L} \\ \end{array}\right) \left( \begin{array}{@{}c@{}} \psi_1 \\ \varphi_1 \\ \end{array}\right)= \left( \begin{array}{@{}cc@{}} S^{{-}1}\mathcal{L} & \mathcal{B} \\ \mathcal{A} & 0 \\ \end{array}\right) \left( \begin{array}{@{}c@{}} \psi_1 \\ \varphi_1 \\ \end{array}\right), \end{equation}

where we have introduced the differential operators

(2.4)\begin{gather} \mathcal{L}\equiv \frac{\partial^2}{\partial x^2}-k^2,\quad \mathcal{F}\equiv 1-d_e^2\mathcal{L}, \end{gather}
(2.5)\begin{gather} \mathcal{A}\equiv {\rm i}k(\psi_0'\mathcal{L}-\psi_0'''), \quad \mathcal{B}\equiv {\rm i}k\left[(\psi_0'-d_e^2\psi_0''')-\rho_s^2\psi_0'\mathcal{L}\right], \end{gather}

with $\mathcal {C}$ being the $\rho _s=0$ limit of $\mathcal {B}$.

The non-ideal parameters of the model are normalised to the reference scale length $L_0$. They correspond to the electron skin-depth, $d_e$, to the ion-sound Larmor radius, $\rho _s$, and to the Lundquist number, $S$. The reference length $L_0$ is chosen to be the equilibrium shear length $a$ of $\psi _0(x)$ at $t=0$, whereas times are normalised to the reference Alfvén time $\tau _A$ computed with respect to the in-plane equilibrium magnetic field component, evaluated sufficiently far from the neutral line.

Several derivations exist of (2.1)(2.2). The parameter $\rho _s$, which violates the Lagrangian invariance of the parallel electron canonical momentum $F\equiv \psi -d_e^2\nabla ^2\psi$, is related to electron parallel compressibility effects (Grasso et al. Reference Grasso, Pegoraro, Porcelli and Califano1999). It is also considered to be a finite-Larmor-radius (FLR) effect, since in a strong guide field limit, it can be shown to be due to a component of the non-isotropic electron pressure tensor that enters in the equation of $\psi$ via the diamagnetic drift component of the fluid electron velocity (see Schep et al. (Reference Schep, Pegoraro and Kuvshinov1994) and Appendix A).

2.1 Magnetic reconnection in tearing modes and formation of magnetic islands

A magnetic reconnection process is characterised by the formation of an $X$-point, where initially distinct magnetic lines have connected thanks to non-ideal effects that become important as the gradients of the magnetic field components, i.e. the components of the current density, are large enough. As a result of the magnetic reconnection event, the plasma system typically relaxes to a final state of lower ‘potential magnetic energy’, the diminished energy being converted into plasma kinetic and internal energy as well as into electron acceleration. During this process, the ‘magnetic topology’ in the plasma changes since the connection of fluid elements is globally modified – see figure 1.

Figure 1. Cartoon schematising three steps of a simplified 2-D magnetic reconnection process that emphasises the notion of change of magnetic topology. (a) Two magnetic lines connecting different pairs of fluid elements, A $\leftrightarrow$ B and C $\leftrightarrow$ D, are twisted by the ideal fluid motion so as to locally generate strong spatial gradients, i.e. current densities. (b) Magnitude of the spatial gradients locally compensates the smallness of the non-ideal parameters in the generalised Ohm's law, by thus allowing the relaxation of the topological constraints that in the ideal limit, forbid the intersection of initially distinct magnetic lines; an $X$-point is formed. (c) Magnetic reconnection process undergone at the $X$-point, by which magnetic energy is dissipated and/or converted into other forms of plasma energy, which makes other magnetic configurations accessible to the plasma; the magnetic topology has globally changed as the magnetic lines now connect the pairs of fluid elements A $\leftrightarrow$ D and B $\leftrightarrow$ C.

Non-ideal effects allow a local violation of the topological conservations implied by the ideal Ohm's law (Cowling Reference Cowling1933; Alfvén Reference Alfvén1942; Batchelor Reference Batchelor1950; Elsasser Reference Elsasser1950a,Reference Elsasserb; Truesdell Reference Truesdell1950; Lundquist Reference Lundquist1951; Newcomb Reference Newcomb1958), which would otherwise prevent initially distinct magnetic lines embedded in the plasma to intersect. Mathematically speaking, this kind of conservation is a consequence of the fact that Faraday–Ohm's law in an ideal MHD plasma,

(2.6)\begin{equation} \frac{\partial {\boldsymbol B}}{\partial t}={\boldsymbol \nabla}\times({\boldsymbol U}\times{\boldsymbol B}), \end{equation}

is equivalent, thanks to a well-known vector identity (${\boldsymbol \nabla }\times ({\boldsymbol U}\times {\boldsymbol B})= {\boldsymbol U}({\boldsymbol \nabla }\boldsymbol {\cdot }{\boldsymbol B})- {\boldsymbol B}({\boldsymbol \nabla }\boldsymbol {\cdot }{\boldsymbol U})+ ({\boldsymbol B}\boldsymbol {\cdot }{\boldsymbol \nabla }){\boldsymbol U}-({\boldsymbol U}\boldsymbol {\cdot }{\boldsymbol \nabla }){\boldsymbol B}$) combined with the continuity equation (Truesdell Reference Truesdell1950), to a vector expression that corresponds to the null Lie-derivative of the vector field ${\boldsymbol B}/n$ dragged by the velocity field ${\boldsymbol U}$: this directly implies the Lagrangian invariance of magnetic lines from which a set of topological conservations follow, the most famous of which go under the names of ‘(Cowling–)Alfvén theorem’ (Cowling Reference Cowling1933; Alfvén Reference Alfvén1942), ‘Woltjer invariants’ Woltjer (Reference Woltjer1958) and ‘connection theorem’ (Newcomb Reference Newcomb1958) – see (Tur & Yanovsky Reference Tur and Yanovsky1993; Kuvshinov & Schep Reference Kuvshinov and Schep1997; Del Sarto et al. Reference Del Sarto, Califano and Pegoraro2006) and references therein for a more detailed discussion; see, e.g. Dubrovin, Nivikov & Fomenko (Reference Dubrovin, Nivikov and Fomenko1991, § III.23) for a definition of Lie derivative in tensor notation in the coordinate representation and Schouten (Reference Schouten1989, § II.8) for the Lie-derivative of tensor densities of arbitrary ‘weight’ (cf. also Lovelock & Rund (Reference Lovelock and Rund1989, p. 105) for the notion of a ‘relative tensor’). This is mathematically analogous to the set of topological conservations related to the fluid vorticity, which were well known to follow from the inviscid vorticity equation in a barotropic fluid (see Truesdell Reference Truesdell1954). In this context, the Alfvén theorem of magnetic flux conservation in an ideal MHD plasma is the mirror correspondent of the Helmholtz–Kelvin theorem of vorticity conservation in ideal hydrodynamics (Batchelor Reference Batchelor1950; Elsasser Reference Elsasser1950b; Truesdell Reference Truesdell1950; Axford Reference Axford1984; Greene Reference Greene1993). We incidentally note that some formal similarities can be also recognised in the eigenmode analysis of resistive tearing modes and of ideal instabilities in presence of viscosity in a Kolmogorov hydrodynamic flow (Fedele, Negulescu & Ottaviani Reference Fedele, Negulescu and Ottaviani2021). In terms of the RMHD equations above, the Lagrangian invariance of magnetic lines is expressed by the ideal limit ($d_e=\rho _s=S^{-1}=0$) of (2.1), although a finite $\rho _s$ alone allows preservation of the ideal MHD topological conservation, provided a redefinition of the stream function of the velocity field ${\boldsymbol U}$ according to $\varphi \to \varphi -\rho _s^2\nabla ^2\varphi$ (Pegoraro et al. Reference Pegoraro, Borgogno, Califano, Del Sarto, Echkina, Grasso, Liseikina and Porcelli2004). In particular, in the collisionless regime, even the nonlinear evolution of tearing-type instabilities can be shown to be ruled by the conservation of Lagrangian invariants whose existence is related to the condition $\rho _s\neq 0$ (Cafaro et al. Reference Cafaro, Grasso, Pegoraro, Porcelli and Saluzzi1998; Grasso et al. Reference Grasso, Califano, Pegoraro and Porcelli2001).

The early notion of magnetic reconnection has been formalised by Dungey (Reference Dungey1950, Reference Dungey1953) after the intuitions by Giovanelli and Hoyle. The former one first noted the occurrence of solar flares in correspondence with regions of local inversion of the magnetic field perpendicular to the Sun surface, and thus made the hypothesis that this could have been a signature of a mechanism of conversion from the magnetic energy to the plasma kinetic energy allowed by resistivity (Giovanelli Reference Giovanelli1946); the second one conjectured that the same may have occurred also in the terrestrial magnetotail (Hoyle Reference Hoyle1949). Resistivity is one of the non-ideal effects capable of violating alone the Lagrangian invariance of ${\boldsymbol B}$, so as the electron inertia, or the electron–electron viscosity, or a non-zero out-of-plane component of the rotational of the divergence of the pressure tensors, also are. All these non-ideal effects, together with other ones such as the Hall-term, can be synthetically expressed with the vector ${\boldsymbol \varPhi }_\varepsilon$ in generalised Ohm's law, $\varepsilon$ generally indicating the infinitesimally small parameter weighting the non-ideal contribution (we use the same normalisation of (2.1)(2.2)):

(2.7)\begin{equation} {\boldsymbol E}+{\boldsymbol U}\times {\boldsymbol B}= {\boldsymbol \varPhi}_\varepsilon. \end{equation}

Magnetic reconnection at least requires ${\boldsymbol \nabla }\times {\boldsymbol \varPhi }_\varepsilon \neq 0$. Although ideal Ohm's law at the MHD scale takes the form of (2.7) with ${\boldsymbol \varPhi }_\varepsilon =0$, note that the dominant contribution to generalised Ohm's law, obtained by summing (A7)(A8) multiplied by the respective charge over mass of the species, comes from electron momentum equation. That is, electrons fix the strongest ‘frozen-in’ condition between the plasma flow and the magnetic field, which, by neglecting all electron inertia, electron–electron viscosity and electron–ion viscosity (i.e. resistivity), and electron-pressure anisotropy, is expressed by the ‘ideal Hall-MHD Ohm's law’: ${\boldsymbol E}+{{\boldsymbol u}^e}\times {\boldsymbol B}= 0 \; \Leftrightarrow \; {\boldsymbol E}+{\boldsymbol u}^i\times {\boldsymbol B}= d_i ({\boldsymbol J}\times {\boldsymbol B})/{n}$. While at MHD scales, where ${\boldsymbol U}\simeq {\boldsymbol u}_i$, the frozen-in condition equally applies for both ions and electrons. At smaller spatial scales, the magnetic field can decouple from the ion fluid, but, as long as ${\boldsymbol E}+{{\boldsymbol u}^e}\times {\boldsymbol B}\simeq 0$ (or, better, as long as ${\boldsymbol E}+{{\boldsymbol u}^e}\times {\boldsymbol B}\simeq \nabla f$) holds, magnetic lines are dragged by the electron fluid flow. Note in this regard that (2.1) for the magnetic stream function $\psi$, corresponding to the $z$-component of Ohm's law, also expresses the variation of the $z$-component of the electron canonical momentum (cf. Appendix A).

The first analytical model of magnetic reconnection was the well-known Sweet–Parker model (Parker Reference Parker1957; Sweet Reference Sweet1958), which assumes a steady inflow condition in an $X$-point: in this case, the resistive reconnection steadily occurs in one point (the $X$-point) of a static current sheet of finite elongation $L$ and of large aspect ratio $L/a$, which asymptotically scales as $S$, when the Lundquist number is defined with respect to the current sheet thickness $a$, as in (2.1) – see figure 2(a) (when $S$ is instead defined – let us call it $S_L$ – with respect to the current sheet length, $L$, the scaling of the Sweet-Parker aspect ratio reads $L/a\sim S_L^{1/2}$). This steady reconnection scenario has been then extended to the inertia-driven regime by Wesson (Reference Wesson1990), for applications to the sawtooth crash in tokamaks, and by Bulanov, Pegoraro & Sakharov (Reference Bulanov, Pegoraro and Sakharov1992) for applications to reconnection in the electron-MHD regime. Variations to this model, in which a different choice of the boundary conditions around the reconnecting region has been made and different rates of magnetic reconnection have been obtained, have been done in subsequent works, starting from that of Petschek (Reference Petschek1964) – see also (Vasyliunas Reference Vasyliunas1975) and references therein.

Figure 2. Comparison of four reconnection scenarios. (a) Sketch of the steady Sweet–Parker reconnection scenario. No particular hypotheses are done on the boundary conditions of the reconnecting current sheet, whereas steadiness and a specific scaling of the aspect ratio with non-ideal parameters like resistivity (i.e. $L/a\sim S^{-1}$, Parker Reference Parker1957; Sweet Reference Sweet1958) or electron inertia (Wesson Reference Wesson1990) are required. The configuration does not represent an instability: no distinction is made between equilibrium and perturbed quantities. (b) Typical ‘magnetic island’ configuration in the tearing-type reconnection. Periodic boundary conditions along the neutral line are required and hyperbolic patterns of the velocity field near the $X$- and $O$-points develop together with the island formation. The wavenumber of oscillations of the mode fixes the number of magnetic islands which are generated. In the linear stage, the island width is $w\simeq \delta \ll a$. (c,d) Merger configuration (c) of two undisplaced line currents (Syrovatskii Reference Syrovatskii1966a; Biskamp & Welter Reference Biskamp and Welter1980; Kleva et al. Reference Kleva, Drake and Waelbroeck1995), which resembles (d) a non-periodic version of the ‘coalescence instability’ scenario. The latter concerns a periodic structure constituted by magnetic islands in a chain (Finn & Kaw Reference Finn and Kaw1977; Pritchett, Lee & Drake Reference Pritchett, Lee and Drake1979; Bhattacharjee, Brunel & Tajima Reference Bhattacharjee, Brunel and Tajima1983). In the displaced current merger case, the attraction of two parallel current filaments squeezes the magnetic field that piles up and typically dissipates at the $X$-point: in this configuration, one may rather speak of ‘magnetic annihilation’ (cf. e.g. Priest & Sonnerup Reference Priest and Sonnerup1975; Watson & Craig Reference Watson and Craig1998; Gu et al. Reference Gu, Pegoraro, Sasarov, Golovin, Yogo, Korn and Bulanov2019). In the coalescence instability scenario, a somewhat similar process occurs (an out-of-plane current density being associated with the $O$-point of each island) once a non-rigid displacement along the neutral line perturbs the periodic configuration by making the islands get closer, pairwise. Similarly, a periodic 2-D ‘paving’ of magnetic cell-like structures of different shape (not shown here) can be destabilised the same way by planar, non-rigid displacements (Longcope & Strauss Reference Longcope and Strauss1993).

Instead, the notion of a tearing mode, the first example of a spontaneous reconnecting instability, was formalised a few years later in the pioneering work by Furth et al. (Reference Furth, Killeen and Rosenbluth1963). In this case, the current sheet is assumed to have periodic boundary conditions along the ‘resonant line’ where the wave vector of the linear perturbation is locally orthogonal to the equilibrium magnetic field. That is, if ${\boldsymbol B}_0(x_s)\boldsymbol {\cdot }{\boldsymbol k}=0$, then the equation $x=x_s$ defines a resonant line in two dimensions and a resonant surface in three dimensions. After a translation of the reference amplitude of the sheared magnetic field component, it can always be assumed that the sheared equilibrium magnetic field is zero at $x=x_s$, and changes sign in its neighbourhood: therefore, in slab geometry, the resonant line is usually termed the ‘neutral line’. It can be seen from the energy principle that the ${\boldsymbol B}_0(x_s)\boldsymbol {\cdot }{\boldsymbol k}=0$ condition minimises the stabilising role played by Alfvénic perturbations: this is why rational surfaces $(m,n)$ in large aspect ratio tokamaks are candidate resonant surfaces for tearing-type instabilities. We recall that a rational surface in a tokamak is a magnetic surface characterised by the condition $q(r)=-m/n$. In the large aspect ratio limit, i.e. the ‘cylindrical tokamak’ approximation, the safety factor reads $q(r)=B_\varphi r/(B_\theta R)$, where $B_\varphi$ and $B_\theta$ are respectively the toroidal and poloidal component of the magnetic field, and $r$ and $R$ are the minor and major radius of the tokamak, respectively. It is easy to verify that the condition ${\boldsymbol k}(r)\boldsymbol {\cdot }{\boldsymbol B}(r)=0$ defines a rational surface for ${\boldsymbol k}=m/(2{\rm \pi} r){\boldsymbol e}_\theta +n/(2{\rm \pi} R){\boldsymbol e}_\varphi$. The harmonic perturbation along the neutral line direction (the wave vector $k$ of (2.1)(2.2)) induces a sinusoidal modulation of the perturbed magnetic field lines that gives rise to the characteristic ‘magnetic island’ pattern: in this case, a pair of $X$-points (at the local minima of $\psi _1$) delimits each magnetic island, and an elliptic point for the magnetic field, called the $O$-point (at the local maximum of $\psi _1$), is at the centre of the island – see figure 2(b) (see also White Reference White1983, Reference White1986) for a tutorial discussion about further features of tearing mode analysis that go beyond the purpose of the present work). In figure 2c,d), other reconnection scenarios related to the merger of current filaments (Syrovatskii Reference Syrovatskii1966a; Biskamp & Welter Reference Biskamp and Welter1980) and to the ‘coalescence’ of a chain of magnetic islands (Finn & Kaw Reference Finn and Kaw1977; Longcope & Strauss Reference Longcope and Strauss1993) are shown for comparison.

As tearing-type modes are linear instabilities, they grow exponentially in time, which marks an important, fundamental difference with respect to the Sweet–Parker scenario: the structure of the reconnecting current sheet is not static but is altered by the current density associated with the tearing mode perturbation, which makes the magnetic island grow in thickness until nonlinear saturation of the instability (see Ottaviani et al. Reference Ottaviani, Arcis, Escande, Grasso, Maget, Militello, Porcelli and Zwingmann2004 and references therein for a survey of different saturation scenarios). This dynamics is associated with hyperbolic patterns of the velocity field around both the $X$- and $O$-points, whereas only the hyperbolic flow at the $X$-point is required in the Sweet–Parker scenario. The hyperbolic flow at the $X$-point is responsible for the well-known quadrupole pattern of the stream function of the velocity field, and the elliptic and hyperbolic structure of $\psi$, $\varphi$ and of their derived fields near the critical points ($X$ and $O$) fix the symmetries of the eigenmode solutions. We incidentally note that this also provides a quite simple explanation (cf. Del Sarto et al. Reference Del Sarto, Pucci, Tenerani and Velli2016, App. A) of the quadrupole pattern associated with the out-of-plane magnetic field component at the $X$-point, in the regime where the Hall term is dominant in the generalised Ohm's law (quadrupole structure which is often recognised as an experimental proxy of Hall-mediated reconnection – cf. e.g. Deng & Matsumoto Reference Deng and Matsumoto2001): in this regime, the Lagrangian invariance of the magnetic field is expressed with respect to the electron flow (see Fruchtman Reference Fruchtman1991; Pegoraro et al. Reference Pegoraro, Borgogno, Califano, Del Sarto, Echkina, Grasso, Liseikina and Porcelli2004; Del Sarto et al. Reference Del Sarto, Califano and Pegoraro2006), which is in turn associated with the in-plane current density, and therefore to the spatial modulation of the out-of-plane magnetic field. If the equilibrium magnetic profile is even with respect to the neutral line, as it is typically assumed in analytical models for tearing modes (as we will do here), the matrix operators of (2.3) commute with the parity operator inducing the transformation $x\leftrightarrow -x$, which makes tearing modes have a fixed parity in $x$. Taking into account the parity in $y$ associated with the harmonic perturbation, the point symmetries of the $\psi _1$ and $\varphi _1$ fields near the critical points ($X$ and $O$) for tearing modes developing in an even magnetic equilibrium can be summarised as

(2.8)\begin{equation} {\text{X-point}: \left\{\begin{array}{@{}ll} \psi\ \text{even in} \ x, & \text{even in}\ y ,\\ \varphi\ \text{odd in}\ x, & \text{odd in}\ y, \end{array}\right. \quad \text{O-point}: \left\{\begin{array}{@{}ll} \psi\ \text{even in}\ x, & \text{even in}\ y ,\\ \varphi\ \text{odd in}\ x, & \text{odd in}\ y .\end{array}\right.} \end{equation}

These symmetries nonlinearly mix via the Poisson bracket operators, of which the terms of (2.1)(2.2) are a rewriting, after linearisation (cf. (A1)(A2), Appendix A). A local quadratic expansion of the eigenmodes near the critical points, which accounts for these nonlinear couplings, allows interesting insight in the early nonlinear, local dynamics of tearing-type instabilities (Pegoraro et al. Reference Pegoraro, Kuvshinov, Romanelli and Schep1995; Del Sarto et al. Reference Del Sarto, Marchetto, Pegoraro and Califano2011).

We conclude by recalling that several effects, in addition to the inclusion of further non-ideal terms in Ohm's law, have been considered, over the years, in the linear theory of tearing modes in reduced MHD. These include the effect of flows parallel to the neutral line, which can have a stabilising effect on the tearing mode instability (Bulanov, Sakai & Syrovatskii Reference Bulanov, Sakai and Syrovatskii1979; Syrovatskii Reference Syrovatskii1981; Faganello et al. Reference Faganello, Pegoraro, Califano and Marradi2010), and which can also make the latter compete with the Kelvin–Helmholtz instability (Hofmann Reference Hofmann1974; Chen & Morrison Reference Chen and Morrison1990; Ofman et al. Reference Ofman, Chen, Morrison and Steinolfson1991; Chen, Otto & Lee Reference Chen, Otto and Lee1997) – cf. also (Einaudi & Rubini Reference Einaudi and Rubini1986; Wang, Lee & Wei Reference Wang, Lee and Wei1988; Bettarini et al. Reference Bettarini, Landi, Rappazzo, Velli and Opher2006; Li & Ma Reference Li and Ma2012) for a numerical linear study; the stabilising effect of an in-plane magnetic field component orthogonal to the neutral line (see Nishikawa Reference Nishikawa1982; Somov & Verneta Reference Somov and Verneta1988, Reference Somov and Verneta1989); the role of equilibrium density gradients, which via the diamagnetic drift frequency gives a time-resonant character to the so-called drift-tearing modes (Coppi Reference Coppi1965; Drake & Lee Reference Drake and Lee1977; Coppi et al. Reference Coppi, Mark, Sugiyama and Bertin1979). These diamagnetic effects are particularly important in tokamaks, where they are related to the rotation of tearing structures and contribute to the nonlinear or resonant coupling of tearing modes with different mode numbers (see, e.g. Hicks, Carreras & Holmes Reference Hicks, Carreras and Holmes1984; Cowley & Hastie Reference Cowley and Hastie1988; Fitzpatrick et al. Reference Fitzpatrick, Hastie, Martin and Roach1993), but also influence their linear evolution and stability threshold (see, e.g. Mahajan et al. Reference Mahajan, Hazeltine, Strauss and Ross1978, Reference Mahajan, Hazeltine, Strauss and Ross1979; Migliuolo, Pegoraro & Porcelli Reference Migliuolo, Pegoraro and Porcelli1991; Grasso, Ottaviani & Porcelli Reference Grasso, Ottaviani and Porcelli2001; Yu Reference Yu2010). In tokamaks, also temperature gradients may induce, in the framework of a gyrokinetic description, the onset of the so-called ‘micro-tearing’ modes (see, e.g. Hazeltine & Ross Reference Hazeltine and Ross1975; Drake & Lee Reference Drake and Lee1977; Gladd et al. Reference Gladd, Drake, Chang and Liu1980; Connor, Cowley & Hastie Reference Connor, Cowley and Hastie1990), which cause the formation of microscopic magnetic islands that can affect turbulent transport (see Doerk et al. Reference Doerk, Jenko, Pueschel and Hatch2011) and foster stochastic fluctuations of the magnetic field, which can in turn affect tearing modes (Carreras, Rosenbluth & Hicks Reference Carreras, Rosenbluth and Hicks1981) and the magnetic confinement (see, e.g. Firpo Reference Firpo2015). Also, we recall that the linear tearing mode theory has been extended and adapted so as to study modes simultaneously and interdependently occurring on two (or more) sufficiently close resonant surfaces, i.e. the double- (or multiple-) tearing modes (Furth, Rutherford & Selberg Reference Furth, Rutherford and Selberg1973; Pritchett, Lee & Drake Reference Pritchett, Lee and Drake1980; Wang et al. Reference Wang, Wei, Wang, Zheng and Liu2011); or to include the so-called ‘neo-classical effects’ related to pressure gradient and toroidal curvature in tokamak plasmas. Although the latter effects are intrinsically nonlinear, they can be self-consistently included in the study of the linear growth of the so-called ‘neo-classical tearing’ modes (NTMs, in specialised ‘jargon’): these are modes developing out of a seed magnetic island (e.g. produced by previous tearing-type instabilities which have saturated and have become stable) and which are driven by the so-called ‘boot-strap current’ (see, e.g. Hahm Reference Hahm1988; Lütjens & Luciani Reference Lütjens and Luciani2002; Wilson Reference Wilson2012). In addition to the large number of works developed over the years on these subjects, attention has also been recently drawn to the role that a background plasma turbulence can have in driving the growth of NTMs (see, e.g. Muraglia et al. Reference Muraglia, Agullo, Benkadda, Garbet, Beyer and Sen2009, Reference Muraglia, Agullo, Benkadda, Yagi, Garbet and Sen2011; Agullo et al. Reference Agullo, Muraglia, Benkadda, Poyé, Dubuit, Garbet and Sen2017a,Reference Agullo, Muraglia, Benkadda, Poyé, Dubuit, Garbet and Senb; Choi Reference Choi2021). These latter subjects are somewhat related (see, e.g. Brennan et al. Reference Brennan, Strait, Turnbull, Chu, La Haye, Luce, Taylor, Kruger and Pletzer2002) to the further topic of forced magnetic reconnection, which has also been considered in the framework of tearing mode theory, by looking at the way an external forcing affects the stability and growth of the linear modes (see, e.g. Hahm & Kulsrud Reference Hahm and Kulsrud1985; Wang & Bhattacharjee Reference Wang and Bhattacharjee1997; Fitzpatrick Reference Fitzpatrick2008). None of these subjects is however of further concern to us; here, in the following, we will focus on the linear problem related to (2.1)(2.2) only, with respect to which of each of these ingredients would provide further elements of analytical complication that would lead us beyond the purpose of this work.

2.2 Eigenvalue of the linear problem and reconnection rate in tearing-type modes

In the literature, the notion of ‘reconnection rate’, let us name it as $\mathcal {R}_{rc}$, formally refers to the rate at which the magnetic flux changes over time close to the $X$-point during a magnetic reconnection event associated with it. In general, we can write

(2.9)\begin{equation} \mathcal{R}_{rc}\equiv \frac{1}{\varPhi_\varOmega({\boldsymbol B})}\frac{{\rm d} \varPhi_\varOmega({\boldsymbol B})}{{\rm d} t}, \quad \varPhi_\varOmega({\boldsymbol B})\equiv \int_\varOmega {\boldsymbol B}\boldsymbol{\cdot} {\rm d}{\boldsymbol S}, \end{equation}

where $\varOmega$ is the resonant surface, which, in a planar configuration, corresponds to the surface generated by translation of the neutral line along the invariant coordinate ($z$ in the slab coordinate system we have chosen). If we assume the planar current sheet (or the neutral line) to be static and along the $y$ direction, then $\varOmega =\ell _y\ell _z$, where the extension of the length $\ell _y$ depends on the specific reconnection process considered.

The way this rate is evaluated clearly depends on the reconnection scenario which is considered. A few words on this topic are therefore useful, since the contrasting terminology acquired on this subject by specialised literature dealing with different reconnection scenarios (e.g. steady, Sweet–Parker-like versus tearing-type reconnection) can be sometimes confusing (cf. also Biskamp (Reference Biskamp2000, p. 54) for a brief discussion as well as Appendix B). This is especially important, since the reconnection rate is related to the rate at which magnetic energy is converted into other forms of plasma energy and therefore to the power released in the form of thermal energy and particle acceleration during a reconnection event. This subject is at the basis of many open questions turning around magnetic reconnection in both laboratory and astrophysical plasmas. These subjects, still debated, concern for example: the model capable of accounting for the short time scale of the sawtooth crash and of other reconnection-related disruptions in tokamaks (see, e.g. Wesson Reference Wesson1986, Reference Wesson2004; Porcelli, Boucher & Rosenbluth Reference Porcelli, Boucher and Rosenbluth1996; Boozer Reference Boozer2012; Jardin, Krebs & Ferraro Reference Jardin, Krebs and Ferraro2020); a theoretical model of the fast release of energy during solar flares and coronal mass ejections (see, e.g. Shibata Reference Shibata1998; Cassak & Shay Reference Cassak and Shay2012; Aschwanden, Xu & Jing Reference Aschwanden, Xu and Jing2014; Aschwanden et al. Reference Aschwanden, Holman, O'Flannagain, Caspi, McTiernan and Kontar2016; Janvier Reference Janvier2017; Aschwanden et al. Reference Aschwanden, Caspi, Cohen, Holman, Jing, Kretzschmar, Kontar, McTiernan, Mewaldt and O'Flannagain2019; Aschwanden Reference Aschwanden2020); the mechanism of energy release at the basis of $X$-rays and $\gamma$-ray emissions (gamma-ray bursts) in pulsars, magnetars and nebulae (see, e.g. Lyutikov Reference Lyutikov2003; Tavani et al. Reference Tavani, Bulgarelli, Vittorini, Pellizzoni, Striani, Caraveo, Weisskopf, Tennant, Pucella and Trois2011; Uzdensky Reference Uzdensky2011); the processes by which the kinetic and magnetic energy of the photospheric plasma are likely to heat up the expanding solar wind by a factor $\sim 10^6$ over the distance of just a solar radius, supposedly via reconnection in the turbulent coronal plasma (see, e.g. Leamon et al. Reference Leamon, Matthaeus, Smith, Zank, Mullan and Oughton2000; Matthaeus & Velli Reference Matthaeus and Velli2011); the energy injected in the magnetosphere via solar wind–magnetosphere interactions and released during geomagnetic storms (see, e.g. Lakhina & Tsurutani Reference Lakhina and Tsurutani2016), and the impact that the ensuing space–weather perturbations may have on terrestrial biosphere and human activity (see National Research Council (2008) and further more specific studies like, e.g. Gopalswamy Reference Gopalswamy2016; Nelson Reference Nelson2016; Eastwood et al. Reference Eastwood, Hapgood, Biffis, Benedetti, Bisi, Green, Bentley and Burnett2018; Knipp et al. Reference Knipp, Fraser, Shea and Smart2018). Regardless of whether the tearing mode theory be actually relevant and capable to explain these phenomena or not, it should be noted that the comparison of theoretical estimates of the reconnection rate with the characteristic time scales directly measured or inferred from experiments or simulations is probably the main criterion presently used to assess the pertinence of tearing-type instabilities or of alternative magnetic reconnection models for these energy releasing processes.

The operational definition by which the reconnected flux can be computed according to (2.9) has been refined over time in different contexts, which also keep account of observational difficulties related to issues encountered in both direct experimental and numerical measurements (see Parker Reference Parker1957; Petschek Reference Petschek1964; Syrovatskii Reference Syrovatskii1966a,Reference Syrovatskiib; Bratenahl & Yeates Reference Bratenahl and Yeates1970; Baum, Bratenahl & White Reference Baum, Bratenahl and White1971; Parker Reference Parker1973; Schnack Reference Schnack1978; Park, Monticello & White Reference Park, Monticello and White1984; Chen et al. Reference Chen, Otto and Lee1997; Hesse, Forbes & Birn Reference Hesse, Forbes and Birn2005; Comisso & Bhattacharjee Reference Comisso and Bhattacharjee2016; Grasso et al. Reference Grasso, Borgogno, Tassi and Perona2020). For the purpose of the present discussion, we here rely on the formal definition: combination of (2.9) with (2.7) via Faraday's law and Kelvin–Stokes theorem of circulation allows us to write

(2.10)\begin{equation} \mathcal{R}_{rc} = \frac{1}{\varPhi_S({\boldsymbol B})} \int_\varOmega \left(\frac{\partial {\boldsymbol B}}{\partial t}-{\boldsymbol \nabla}\times({\boldsymbol U}\times{\boldsymbol B})\right)\boldsymbol{\cdot} {\rm d}{\boldsymbol S} ={-}\frac{c}{\varPhi_S({\boldsymbol B})} \int_\varOmega ({\boldsymbol \nabla}\times\mathcal{\boldsymbol F}_\varepsilon )\boldsymbol{\cdot} {\rm d}{\boldsymbol S}, \end{equation}

where the curl term in the first integral generally accounts for the evolution of the surface over which the magnetic flux is calculated, as it is dragged by the bulk plasma velocity ${\boldsymbol U}$. If the surface over which the flux is evaluated is static, like it can be chosen in the Sweet–Parker scenario or for non-propagating tearing-like modes, that contribution is null. This is the case in which we are interested, where we consider a non-evolving integration surface along the neutral line, whose extension in the $y$ direction depends on the reconnection scenario considered, as it will be specified below. If, under this assumption, we specify the surviving terms of the first integral of (2.10) for the slab geometry, RMHD variables ${\boldsymbol B}=\boldsymbol {\nabla }\psi \times \boldsymbol {e}_z+b\boldsymbol {e}_z$ of (2.1)(2.2), we can write ${\rm d}{S}={{\rm d} y}\,{\rm d}z\,{\boldsymbol e}_x$ and therefore

(2.11)\begin{equation} \varPhi_\varOmega({\boldsymbol B})=\int_\varOmega {\boldsymbol B}\boldsymbol{\cdot} {\rm d}{\boldsymbol S} = \int_\varOmega {B}_{x,1} \,{{\rm d} y} \,{\rm d}z =\int_{l_z}\,{\rm d}z \int_{l_y} \frac{\partial \psi_1}{\partial y} \,{{\rm d} y}. \end{equation}

Thus, posing the $X$-point to lay on the line $x=0$, we obtain

(2.12)\begin{equation} \mathcal{R}_{rc} = \frac{\int_{l_y}{{\rm d} y}\;\partial^2 \psi_1(0,y,t)/\partial t \partial y}{\int_{l_y}{{\rm d}y}\;\partial \psi_1(0,y,t)/\partial y}. \end{equation}

In the case of tearing modes, it is easy to see from the symmetries of the eigenmode solutions in (2.8) that it is appropriate to consider the integration interval in $y$ to not have the $X$-point in the middle (otherwise the spatial integral would vanish). For a single tearing mode, the interval can be therefore taken to extend from the $X$-point to the $O$-point (although each $O$-point is delimited by two $X$-points, the periodicity of the configuration makes the numbering of $X$- and $O$-points be in a 1:1 correspondence). By construction, the eigenmode of the magnetic stream function has a separable dependence on time and on space of the form $\psi _1(x,y,t)=f(x,y) {\rm e}^{\gamma t}$. In reality, although we are here anticipating the formal results, as we will discuss in the next sections, it is easy to determine from the magnetic island shape that the solution is of the form $\psi _1(x,y,t)=g(x)\cos (ky){\rm e}^{\gamma t}$, if $y=0$ is taken at the $O$-point – cf. figure 2(b). Hence, one immediately obtains (the label $TM$ standing for ‘tearing mode’)

(2.13)\begin{equation} \mathcal{R}_{TM}(k) = \gamma(k), \end{equation}

that is, for a tearing-type instability of wavenumber $k$, the notions of reconnection rate and of growth rate are equivalent. The quadratic dependence of the magnetic energy on $\psi$ (cf. (A5)) trivially implies that for a single tearing mode, the rate at which magnetic energy $\mathcal {E}_{B}=\int |{\boldsymbol \nabla }\psi _1|^2\,{{\rm d} x}\,{{\rm d} y}$ is converted into other forms of energy during the linear stage of the instability is $2 \mathcal {R}_{TM} = 2\gamma$.

Of course, if several tearing modes are simultaneously unstable, and/or if one or more reconnection instabilities are in their nonlinear stage, the identification $\mathcal {R}_{rc}\leftrightarrow \gamma (k)$ is not correct any longer, even if a dominant mode can be assumed to dominate the reconnection process. In these cases, an evaluation of the global reconnection rate in a given volume, more accurate at least from a numerical point of view, can be made by taking

(2.14)\begin{equation} \mathcal{R}_{rc} \simeq \frac{1}{2 \mathcal{E}_c} \left|\frac{\partial\mathcal{E}_c}{\partial t}\right|, \end{equation}

where $\mathcal {E}_c$ is a component of the total energy (cf. (A5)) involved in the reconnecting process and the factor $1/2$ accounts for the quadratic relation between the fields and the energy components. Usually, $\mathcal {E}_c$ is the magnetic energy component $\mathcal {E}_{B}$, which in most reconnection regimes (see, e.g. White Reference White1983, Reference White1986) can be shown to provide the ‘reservoir’ of energy of the instability, which is ‘dissipated’ (or, more generally, converted into other forms) during the process. It should be however noted that in some regimes of tearing-like instabilities, numerically studied by using ‘large’ values of $d_e/L_0$ (i.e. $d_e/L_0\lesssim 1$ but not asymptotically smaller than unity), the role of $\mathcal {E}_{B}$ has been observed to be played, instead, by the electron kinetic energy $\mathcal {E}_{J}=\int \,{\rm d}_e^2|{\nabla }^2\psi _1|^2\,{{\rm d} x}\,{{\rm d}\kern 0.05em y}$ associated with the equilibrium current sheet (Del Sarto, Califano & Pegoraro Reference Del Sarto, Califano and Pegoraro2005; Del Sarto et al. Reference Del Sarto, Califano and Pegoraro2006). Also, the increase of some kinetic energy component could be used for the estimate of (2.14), if it assumed that the electromagnetic energy released during the reconnection event is dominantly transferred to particle heating and/or acceleration, or to radiation. This is the way by which, for example, the reconnection rate of processes at the basis of solar flares and coronal mass ejections, or of the sawtooth cycle in tokamaks, is indirectly inferred from observational measures.

Some alternative definitions of ‘reconnection rate’, provided in terms of local estimates of quantities, are also frequently used in the literature. These further definitions, briefly outlined and discussed in Appendix B for comparison, are not of concern for the purpose of this work, since they mostly deal with stationary reconnection or provide approximations to the more accurate (2.13), valid in the case of interest to us.

We conclude this section with a note on the notion of ‘fast’, often used in the literature for models devised in the attempt of modelling the rapid reconnection processes observed in Nature and in experiments, which display growth rates faster – in an asymptotic sense – than those predicted by the early reconnection models (i.e. resistive Sweet–Parker and resistive tearing modes). On the one hand, the notion of ‘fast’ has changed over the years, both depending on the specific reconnection context (space or laboratory) and on the physical process allowing a relative increase of the reconnection rate (e.g. electron inertia, the inclusion of the Hall-effect in Ohm's law, the accounting for other nonlinear effects, 3-D effects and secondary instabilities – see appendices of Del Sarto & Ottaviani Reference Del Sarto and Ottaviani2017) for a short historical review on these subjects). On the other hand, it is unlikely that a primary spontaneous tearing mode developing on a static current sheet overtake values of order unity, when measured with respect to the normalisation time $\tau _A$ defined as below ((2.4)(2.5)) – see Pucci & Velli (Reference Pucci and Velli2014), Del Sarto et al. (Reference Del Sarto, Ottaviani, Pucci, Tenerani and Velli2018), and § IX of Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) for more detailed discussions. It is however worth noting that for linear tearing modes, the normalised growth rate evaluated from (2.1)(2.2) typically becomes of the order of unity or, better, of some decimal fraction of it, when the shear length becomes comparable to the microscopic non-ideal scales of interest: that is, $\gamma \tau _A\sim O(10^{-1})-O(1)$, when some among $d_e$, $\rho _s$ or $(S^{-1}/\gamma )^{1/2}$ become of the order of $O(a/10)-O(a)$. However, this also generally means that we are out of the limits of applicability of the boundary layer theory (cf. § 3, next, and note that we leave aside, here, the open problem of discussing the validity of an extended fluid modelling at these kinetic scales): in this case, the analytical estimates of $\gamma (k)$, which we are going to develop below, are not applicable and a numerical computation is instead required. See figure 3 and table 2 of Del Sarto et al. (Reference Del Sarto, Marchetto, Pegoraro and Califano2011) for an example in the collisionless limit.

Figure 3. Growth rates of collisionless tearing modes numerically computed from (2.1)(2.2) with the solver of Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020), for values of the non-ideal parameters at the limits of (or beyond the) applicability of boundary layer theory, and for a wavelength $ka\sim O(1)$ (likely value for modes on large aspect ratio current sheets having a microscopic thickness – a condition typically met, e.g. in MHD turbulence (see, e.g. Franci et al. Reference Franci, Landi, Matteini, Verdini and Hellinger2016; Del Sarto & Pegoraro Reference Del Sarto and Pegoraro2017; Franci et al. Reference Franci, Papini, Del Sarto, Hellinger, Burgess, Matteini, Landi and Montagud-Camps2022). Note that once $\gamma$ approaches unity, the effect of the non-ideal parameters on the magnetic equilibrium $\psi _0$ is not negligible anymore, and (2.1) and (2.2) in the resistive limit (not shown here) should be modified accordingly (cf. Appendix A). (a) Case of cold electrons. (b) Case of warm electrons, for different values of $d_e/a$ and for values of $\rho _s/a$ approaching one.

3 Boundary layer approach for tearing modes: large- and small-$\varDelta '$ regimes

In this section, we start by describing the general strategy of integration with the boundary layer approach to solve the eigenvalue problem of (2.1)(2.2). Then, we discuss the ordering and the instability parameter, which defines the unstable spectrum of wavenumbers, and its consequences on the classification of different reconnection regimes in both slab and toroidal geometries. Finally, we close the section by giving a brief review on previous works using the boundary layer integration in the collisionless regime.

There are no exact analytical solutions available for the general eigenvalue problem of (2.1)(2.2). Approximated analytical solutions can be obtained by using a boundary layer approach as first shown for the purely resistive tearing by Furth et al. (Reference Furth, Killeen and Rosenbluth1963) in the constant-$\psi$ regime and by Coppi et al. (Reference Coppi, Galvao, Pellat, Rosenbluth and Rutherford1976), Ara et al. (Reference Ara, Basu, Coppi, Laval, Rosenbluth and Waddell1978) and Basu & Coppi (Reference Basu and Coppi1981) in the internal-kink regime. When kinetic-like effects are included, it is instead easier to solve the boundary layer problem if calculations are performed after doing a Fourier transform with respect to the variable $x$: this facilitates the analytical calculations since it lowers the order of the differential equations in the ‘inner region’, as it has been first shown by Pegoraro & Schep (Reference Pegoraro and Schep1986) and Pegoraro et al. (Reference Pegoraro, Porcelli and Schep1989).

The boundary layer approach consists in solving the linear problem in distinct overlapping regions. The number of regions depends on the non-ideal parameters that exist in the problem. Moreover, for the boundary layer approach to be applied, it is necessary that the normalised non-ideal parameters (e.g. $\varepsilon =S^{-1}$ or $\varepsilon =(d_e/a)^2$) be much smaller than unity, i.e. $\varepsilon \ll 1$. This makes it possible to perform an asymptotic analysis by expanding quantities in powers of $\varepsilon$, and also grants the scale separation between the boundary layers, since the characteristic width of the innermost layers, where ideal-MHD breaks down, is a posteriori found to scale with a multiplication of positive powers of the parameters ‘$\varepsilon$’ at play. Quantitatively speaking, comparison between theoretical estimates and numerical integration of the boundary layer problem indicates that the condition $\varepsilon \ll 1$ typically means that the microscopic scales (cf. end of § 2.2) related to the non-ideal terms capable of breaking the ideal MHD conservation of the magnetic topology be not larger than a fraction $\sim 10^{-1}$ of the equilibrium shear length $a$.

For example, for purely resistive or purely inertia-driven tearing modes in RMHD, the linear problem is solved in two distinct regions and the solutions are asymptotically matched in an intermediate layer, as shown in figure 4(a). In this case, the first region is the ‘outer’ region in which the plasma is assumed to be at ideal MHD force-balance equilibrium at spatial scales of the order of $L_0 = a$. The second region is the ‘inner’ one, in which a solution is sought for the differential equations while retaining the non-ideal parameters in the limit $x\ll 1$ (i.e. $x/L_0\ll 1$ in dimensional coordinates). The characteristic size of the inner region is identified to correspond to the reconnection layer width (cf. § 8), whose characteristic thickness we hereafter call $\delta$, with $\delta \to 0$ as the non-ideal parameters tend to zero. The inner region is therefore defined by the inequality $x/\delta \lesssim 1$.

Figure 4. (a) Sketch of the boundary layer decomposition for the calculation of the tearing eigenmodes with asymptotic matching techniques, when a single boundary layer is sufficient (e.g. case of the resistive tearing first considered by Furth et al. Reference Furth, Killeen and Rosenbluth1963). The width $\delta$ of the inner boundary layer is here the natural candidate for the width of the ‘reconnecting layer’, although the latter is not formally defined, i.e. we can assume $\delta =\delta _1$ (an assumption we will support in § 8 with further arguments). (b) Sketch of the boundary layer decomposition when two boundary layers are required (e.g. case of the warm-collisionless RMHD). Note that the reconnecting layer width, which we call $\delta$, is, again, a priori not defined but, as we will see (§ 8), further arguments can be found to identify it with the innermost layer width, i.e. $\delta =\delta _1$.

Seen from the inner region, the convergence to some point in the outer domain is expressed with the limit $x/\delta \to \infty$ with respect to the ‘stretched variable’ $x/\delta$. In the matching layer, this limit should match the limit $x/L_0\to 0$ (or, simply, $x\to 0$ in normalised units) for quantities evaluated in the outer region.

When more than one non-ideal parameter is involved in the problem, applying the boundary layer approach becomes more difficult. This is due to the presence of at least two boundary layers and therefore of two matching regions (cf. figure 4b), whose characteristic widths, i.e. the normalisation scales for the asymptotic matching, say $\delta _2$ and $\delta _1$, depend on the spatial scales of the problem in a non-trivial way. Moreover, the notion of ‘reconnection layer width’, which we keep on naming $\delta$, in this case is not a priori related to the width of one of the boundary layers, although, as we will see (§ 8), we can make it correspond to the extension of the innermost layer, $\delta _1$.

Operationally speaking, the integration of tearing mode equations by means of the boundary layer approach, which we will detail in the nextsections, can be split in the following sequence of steps.

  1. (i) Solving the equations in the ideal MHD limit provides the solution in the outer layer (the ‘outer solution’) and allows one to evaluate the instability parameter (the so-called $\varDelta '$ parameter, see § 3.1), which identifies the range of wavenumbers that are tearing-unstable and allows one to distinguish among different wavelength regimes of reconnection – the constant-$\psi$ or tearing regime, the large-$\varDelta '$ or internal-kink regime, and the fastest growing mode in a continuum wavelength spectrum of slab tearing modes.

  2. (ii) Considering the equations in the non-ideal region by assuming that the inverse spatial gradients are small enough to count the microscopic scales associated with the non-ideal parameters. This also allows some simplifications to the equations, related to the fact that the $n$th derivative with respect to $x$ dominates over the $n$th power of $k$.

  3. (iii) Establishing, on heuristic basis, an ordering among terms of the non-ideal equations depending on the value of the non-ideal microscopic parameters. This allows one to distinguish different boundary layers (cf. figure 4b) in which the non-ideal equations are differently approximated. This task becomes more complex as more non-ideal parameters are present, and is subordered to the consistency of the solutions a posteriori found after integration in each sub-region (hence, the ‘heuristic’ nature of the ordering of the terms in the equations).

  4. (iv) Integrating the non-ideal equations in each sub-region, by taking the solution in the next-most, outer domain as a boundary condition. For example, with reference to figure 4(b), the outer solution in the ideal region provides the boundary condition in the matching layer II for the solution in the sub-domain $|x|\lesssim \delta _2/2$, which in turn provides the boundary condition in the matching layer I for the solution in the region $|x|\lesssim \delta _1/2$.

  5. (v) Finally note that, in any case, constructing the global solution in the whole domain in an explicit form is a not evident task and often it is not possible, since the integral solutions in each non-ideal region may be not obtained in closed form. Nevertheless, it is possible to obtain a quantitative estimate of the eigenvalues of the linear problem (i.e. the growth rate of the reconnecting mode) from the conditions on the solutions in the non-ideal region. Similarly, other spatial scales of interest (which, by further arguments, can be interpreted as corresponding, e.g. to the reconnecting layer width, $\delta$ – cf. § 8) can be quantitatively evaluated.

3.1 Orderings and instability parameter $\varDelta '$

In the present section, we discuss the different orderings of the operators in the eigenvalue problem, which will allow us to obtain the governing equations in different regions of the domain, and we introduce the notion of an ‘instability parameter’.

In the outer, ideal MHD region, the ordering of different operators in (2.3) is

(3.1)\begin{equation} \gamma\sim 0, \quad \mathcal{L}\sim 1, \quad \mathcal{F}\simeq 1, \quad \mathcal{A}\sim 1, \quad \mathcal{B}=\mathcal{C}={\rm i}k\psi_0'\sim 1. \end{equation}

Solving the equations resulting from this ordering allows us to obtain the solution in the outer region, $\psi _{\text {out}}$.

At the neutral line, the equilibrium magnetic field (shear field) vanishes and changes its sign when reconnection occurs. Therefore, this field should be an odd function in the vicinity of $x=0$. Hence, $\psi _0$, whose $x$-derivative represents the shear field $B_y^0(x)$, is an even function. Moreover, close to the neutral line, $\psi _0(x)$ is continuous at least up to its first derivative. We then expand it as a Taylor series which, for symmetric tearing modes in slab geometry, has non-zero coefficients only for even powers of $x$,

(3.2)\begin{equation} \psi_0(x)|_{x\to 0}\simeq C_0+C_2x^2+C_4x^4+\cdots. \end{equation}

Therefore, we can relate $\delta$ to the intensity of the equilibrium magnetic field and to the gradient of the associated current density inside of the inner layer, when looked at from the matching (or from the outer) region. So we write

(3.3)\begin{equation} \left.\psi_0'\right|_{x\to 0}\sim \left.\psi_0'''\right|_{x\to 0}\sim x|_{\text{inner}}\simeq {\delta}. \end{equation}

Before going further, we note that condition (3.2) is too restrictive, in general, for slab tearing modes on magnetic equilibria that are not symmetric with respect to the shear coordinate. An example is provided by tearing modes in a cylinder, for which $C_1=0$ but $C_{3} \neq 0$ (Bertin Reference Bertin1982; Militello et al. Reference Militello, Huysmans, Ottaviani and Porcelli2004, Reference Militello, Borgogno, Grasso, Marchetto and Ottaviani2011). It must be also noted that the inner layer width, $\delta$, has not yet been ‘defined’ here. As a quantity, it only explicitly appears in boundary layer calculations as a normalisation scale for the differential equations in the non-ideal region, whose boundaries are defined by the condition $|x|/ \delta \sim O(1)$. To date, indeed, no general criterion exists to quantitatively define $\delta$, whose estimation is made, when possible, thanks to further hypothesis and heuristic ansatz (e.g. comparison with the characteristic scales of the problem). In the present section, we therefore assume $\delta$ to be simply defined as the innermost layer width (in § 8, we will propose a quantitative general definition of $\delta$, whose appropriateness in the different regimes we will prove numerically by comparing it with theoretical estimates).

Unless a large aspect ratio current sheet is considered, for which an almost continuous spectrum of wavenumbers can be destabilised that also allows $k$ to be large, the further ordering $\partial _x\gg k$ is assumed for $x\ll 1$. This implies the following orderings inside the non-ideal region:

(3.4)\begin{align} \mathcal{L}\simeq \partial_x^2,\quad \mathcal{F}= 1-d_e^2\mathcal{L}, \quad \mathcal{A}={-} k\delta(1-\mathcal{L})\sim{-}k\delta\mathcal{L}, \quad \mathcal{B}= ik\delta\left[(1-d_e^2)-\rho_s^2\mathcal{L}\right]. \end{align}

The corresponding equations are solved for the ‘inner’ functions $\psi _1=\psi _{\text {in}}$ and $\varphi _{1}=\varphi _{\text {in}}$.

The matching with the outer solution is assumed as a boundary condition to be imposed across an intermediate matching layer, where one must compare the asymptotic series $\sum a_{n}^{(\text {in})}(x/ \delta )^n$ for $(x/ \delta )\gg 1$ and $\sum _n a_{n}^{(\text {out})}x^n$ for $x\ll 1$ (we recall that all lengths appear here as normalised to $L_0=a$), respectively representing the inner and outer solutions – see Bender & Orszag (Reference Bender and Orszag1978, § 9). This translates in the condition

(3.5)\begin{equation} \lim_{x/ \delta\to\pm\infty} a_{m}^{(\text{in})}\,(x/ \delta)^m\simeq \lim_{x\to \pm 0} a_{m}^{(\text{out})}\,x^m\quad \begin{array}{c} \text{(for }m\text{ that corresponds to}\\ \text{the leading term of the series).} \end{array} \end{equation}

Condition (3.5) is usually expressed in a looser notation as

(3.6)\begin{equation} \lim_{x/ \delta\to\pm\infty}\psi_{\text{in}}(x/ \delta)=\lim_{x\to \pm 0}\psi_{\text{out}}(x). \end{equation}

Although (3.6) formally compares the two numerical values of the limits of the eigenfunctions solving the differential equations, in the following, we will mean this expression as a shortcut writing of (3.5), as it is usually done in tearing mode analysis.

The dependence on the outer solution becomes then explicit through the relation

(3.7)\begin{equation} \int_{-\infty}^{+\infty}\psi_{\text{in}}''\,{\rm d}\left({x}/ {\delta}\right)= \psi_{\text{out}}(0)\varDelta', \end{equation}

which introduces the instability parameter (Furth et al. Reference Furth, Killeen and Rosenbluth1963) related to the discontinuity which is met in the first derivative of $\psi _{\text {out}}$ as $x\to \pm 0$,

(3.8)\begin{equation} \varDelta'(k;\psi_0)\equiv \lim_{\epsilon\to 0} \frac{\psi_{\text{out}}'(\epsilon)-\psi_{\text{out}}'(-\epsilon)}{\psi_{\text{out}}(0)}. \end{equation}

In the expression above, $\epsilon > 0$. Definition (3.8) enters in (3.6) as

(3.9)\begin{equation} \int_{-\infty}^{+\infty} \psi''_{\text{in}}\,{\rm d}(x/ \delta) = \psi'_{\text{in}}(+\infty) - \psi'_{\text{in}}(-\infty) = \psi'_{\text{out}}({+}0) - \psi'_{\text{out}}(- 0) = \psi_{\text{out}}(0)\varDelta', \end{equation}

where we have used the key fact that the outer solution $\psi _{\text {out}}$ is continuous as $x/L_0\to 0$ while this is generally not the case for its derivative $\psi _{\text {out}}'$ (as it is found a posteriori when solving the inner equation for tearing-type modes).

The linear problem in the inner region is closed by combining (3.7) with the relation that can be established between $\psi _{\text {in}}$ and $\varphi _{\text {in}}$. According to the inner region ordering, (2.2) becomes $\gamma \varphi ''_{\text {in}} = {\rm i}kx\psi ''_{\text {in}}$. Therefore,

(3.10)\begin{equation} \int_{-\infty}^{+\infty} \psi''_{\text{in}}\,{\rm d}\left({x}/ {\delta}\right) ={-} ({\rm i} \gamma/ k) \int_{-\infty}^{+\infty} \frac{\varphi''_{\text{in}}}{ x/ \delta} \,{\rm d}\left({x}/ {\delta}\right). \end{equation}

It must be however noted that when more than one non-ideal parameter is involved and/or when the microscopic scale of variation of $\psi$ and $\varphi$ differ (this happens, for example, in 2-D electron-MHD (Bulanov et al. Reference Bulanov, Pegoraro and Sakharov1992), where the fluid stream function is related to the fluctuation $b$ of the $B_z$ magnetic component and an equation for $b$ replaces that for $\varphi$), the matching procedure above requires more care, as we are going to see in § 7. In these cases, more than two boundary layers must be considered and the corresponding solutions must be matched – see Pegoraro & Schep (Reference Pegoraro and Schep1986), Porcelli (Reference Porcelli1991) and Bulanov et al. (Reference Bulanov, Pegoraro and Sakharov1992). These are the cases for which heuristic estimations are difficult, since a different width must be associated with each boundary layer (e.g. $\delta _2$ and $\delta _1$, with $\delta _1<\delta _2$). In this kind of analysis, the width of each layer appears as the characteristic normalisation scale, say $l_{\text {norm}}$, with respect to which to consider the limits $x/l_{\text {norm}}\ll 1$ and $x/l_{\text {norm}}\gg 1$ while performing the asymptotic matching in the intermediate region between the layers. In this sense, the limit $\varepsilon \to 0$ of (3.8) also must be reinterpreted, which we should read as $\varepsilon =|x|/l_{\text {norm}}$ for $l_{\text {norm}}=L_0=a$. At the same time, also the relations between $\varDelta '$ and the spatial gradients of $\varphi _{\text {in}}$ expressed by combinations of (3.7) and (3.10) become non-trivial, as we will discuss in § 10.

3.2 Small- and large-$\varDelta '$ limits in slab geometry and in tokamaks

For each eigenvalue problem in slab geometry, the numerical value of $\varDelta '$ depends both on the choice of the magnetic equilibrium profile and on the value of $k$. A classification of regimes of slab tearing modes can be generally done depending on the numerical comparison of $(\varDelta ')^{-1}$ and of the innermost layer width, $\delta _1$, regardless of the number of boundary layers involved. For simplicity of notation, we generically use for them the symbol $\delta$, although it is only later (§ 8) that we will provide some argument to identify the ‘reconnecting layer width’. In particular, one speaks of a small-$\varDelta '$ limit for $\varDelta '\delta \ll 1$ and of a large-$\varDelta '$ limit for $\varDelta '\delta \gg 1$.

Once the scaling of $\delta$ on the non-ideal parameters involved is known, the large- and small-$\varDelta '$ limits in slab geometry can be made to respectively correspond to the small and large wavelength limits of the tearing dispersion relation, which for each reconnection regime can be defined with respect to the comparison of $kL_0=ka$ with powers of the non-ideal parameters involved (as it has been discussed for example by Bulanov (Reference Bulanov2017) for the purely resistive case).

In particular, since $\varDelta '$ defined through (3.7)(3.8) is, when analytically obtained (cf. example below – § 4.1), a continuous function of the variable $k$, the large wavelength limit can be postulated to correspond to a power-law dependence of the kind

(3.11)\begin{equation} \lim_{ka\to 0} \varDelta'(ka)=(ka)^{{-}p},\quad p>0, \end{equation}

where $p$ depends on the initial equilibrium profile (Del Sarto et al. Reference Del Sarto, Pucci, Tenerani and Velli2016; Pucci et al. Reference Pucci, Velli, Tenerani and Del Sarto2018; Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020). While the limit above corresponds to $\varDelta '\to \infty$, the marginal stability condition $a\varDelta '(ka)\to 1$ is approached as $ka\to 1$.

A fundamental difference between large- and small-$\varDelta '$ tearing-type modes in tokamak devices and their corresponding small- and large-$k$ limits in slab geometry must be however remembered, when the results of the two models are compared: in tokamak devices, the wave vector is fixed by the resonant surface on which reconnection occurs, and the large- or small-$\varDelta '$ condition is therefore determined by the specific shape of the unstable magnetic profile (sometimes in turn determined by some ideal instability, which has previously occurred and which has modified the otherwise stable magnetic configuration). In slab Cartesian geometry, instead, the transition between the small- and large-$\varDelta '$ limit occurs by moving along the $k$ interval for a fixed, tearing-unstable magnetic profile, and depends on the aspect ratio of the associated current sheet (see figure 5).

Figure 5. Scheme of the correspondence between linear reconnecting instabilities in the RMHD slab geometry limit and in the RMHD ‘cylindrical tokamak’ approximation, which is compatible with the strong guide field assumption (Strauss Reference Strauss1976). At the bottom of the figure, we highlight the fact that the slab geometry easily allows modelling both of tearing modes of different wavenumber (centre frame) and of the double tearing mode, i.e. the tearing-like instability simultaneously occurring on two sufficiently close resonant surfaces (Furth et al. Reference Furth, Rutherford and Selberg1973), of which we have shown here only the symmetric case (rightmost frame). Note however that we have considered here the ‘cylindrical tokamak’, or ‘large aspect ratio tokamak’ approximation, in the strict limit of $\partial / \partial \varphi \to 0$, which maps into the $\partial / \partial z \to 0$ assumption that we consider in this article. In reality, the RMHD modelling also allows for inclusion of the neglected derivatives with respect to the axis-symmetric coordinate ordered as $\partial / \partial \varphi \sim \partial / \partial z \sim \varepsilon _B$, in terms of the ratio $\varepsilon _B$ between the in-plane and the guide magnetic field components at equilibrium (cf. (A11) and discussion in the last paragraph of Appendix A).

We remark indeed that the distinction between small- and large-$\varDelta '$ limits has been historically introduced to characterise, in terms of the instability parameter $\varDelta '$ defined by Furth et al. (Reference Furth, Killeen and Rosenbluth1963), two different types of unstable modes observed in tokamaks. In tokamak physics, for positive values of $\varDelta '$, a distinction is made between the tearing mode or constant-$\psi$ mode, first studied in the cylindrical geometry approximation by Furth et al. (Reference Furth, Killeen and Rosenbluth1963) (and which formally corresponds to $\varDelta '\delta \ll 1$) and the internal kink mode or $m=n=1$ non-ideal kink mode, first identified in cylindrical geometry by Coppi et al. (Reference Coppi, Galvao, Pellat, Rosenbluth and Rutherford1976) (and which formally has $\varDelta '=\infty$) (we recall indeed that, even if $\varDelta '<0$, ideal instabilities (i.e. the ‘ideal kink mode’) can also occur in a tokamak, depending on the value of the safety factor $q$ – cf. Wesson (Reference Wesson1990), § 6). In a cylindrical tokamak approximation, these two modes generally occur on different magnetic surfaces and for specific values of the wavenumbers, and also display different relations between the eigenfunctions $\varphi _1$ and $\psi _1$: differently from the tearing mode, for which the fluid displacement $\xi \equiv k^2 \varphi _1/\gamma$ is proportional to $\sim \psi _1/x$, the internal kink mode corresponds to a rigid displacement $\xi \sim const.$ of the plasma inside the inner region; more specifically, for the internal kink mode, $\xi \sim const.$ in the whole subdomain $r\leq r_s$, where $r_s$ is the radius of the resonant surface, whereas $\xi = 0$ for $r>r_s$ (see, e.g. Porcelli Reference Porcelli1987; Del Sarto & Ottaviani Reference Del Sarto and Ottaviani2017 for a more detailed discussion). It has been however shown by Ara et al. (Reference Ara, Basu, Coppi, Laval, Rosenbluth and Waddell1978) that, thanks to the formal analogy between the two corresponding eigenvalue problems in cylindrical and Cartesian geometry, the transition between the two types of modes can be modelled in a slab geometry configuration by varying the value of the wavenumber for a fixed magnetic equilibrium and for fixed values of the non-ideal parameters. This corresponds to a transition between the small- and the large-$\varDelta '$ limits. In this context, both the tearing (constant-$\psi$) mode and the internal-kink mode of tokamak physics can be considered as examples of tearing-type modes when modelled in slab, Cartesian geometry. In this case, a larger ‘free energy’ can be associated with tearing-type modes in the large-$\varDelta '$ limit. This is the approach we take here.

Another fundamental difference between tearing and internal kink modes in tokamaks on the one side, and tearing-type modes in slab geometry on the other side, is the fact that in slab geometry, it makes sense to identify a further wavenumber ‘regime’ that is characterised by the condition $\varDelta '\delta \sim 1$ (see Loureiro, Schekochihin & Cowley Reference Loureiro, Schekochihin and Cowley2007; Bhattacharjee et al. Reference Bhattacharjee, Huang, Yang and Rogers2009; Del Sarto et al. Reference Del Sarto, Pucci, Tenerani and Velli2016; Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) at the varying of the non-ideal parameters involved. This wavelength limit, which is met thanks to the possibility to perform a ‘continuous’ transition from $\varDelta '\delta \ll 1$ to $\varDelta '\delta \gg 1$, corresponds to the fastest growing mode that can be destabilised in a periodic current sheet when a continuum spectrum of wavenumbers is admitted (see Furth et al. Reference Furth, Killeen and Rosenbluth1963, appendix D, and also Biskamp Reference Biskamp1982). Its scaling is exemplified in figure 6 for the case of purely resistive tearing modes (i.e. $d_e=0,\rho _s=\nu =0$ in (2.1)(2.2)) destabilised on an equilibrium profile $\psi _0=\cos (x/a)$, whose $\varDelta '(k)$ formula has been discussed by Ottaviani & Porcelli (Reference Ottaviani and Porcelli1993, Reference Ottaviani and Porcelli1995) and for which the scaling $\gamma _M\sim S^{-1/2}$ can be deduced (dashed line in the figure).

Figure 6. Dispersion relation of the purely resistive tearing as a function of $ka$ (dots correspond to numerically computed values), evaluated for the equilibrium $\psi _0=\cos (x/a)$. The smaller box is a ‘zoom’ of the plot for low values of $\gamma$, which correspond to the range of smaller values of $S^{-1}$ indicated in the caption. The dashed line corresponds to the fastest mode scaling, which for $\psi _0=\cos (x/a)$ is $\gamma _M\sim S^{-1/2}$. This can be obtained by using in (24) of Del Sarto et al. (Reference Del Sarto, Pucci, Tenerani and Velli2016) the value $p=2$, which can be deduced from the $ka\ll 1$ limit of the corresponding $\varDelta '(ka)$ formula (see Ottaviani & Porcelli Reference Ottaviani and Porcelli1993 or (8) of Ottaviani & Porcelli Reference Ottaviani and Porcelli1995). An 8th-order accurate scheme has been used here for integration on a uniform grid by using the solver discussed by Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020).

When tearing modes are destabilised in a large aspect ratio, periodic current sheet, so that also the $\varDelta '\delta \sim 1$ modes are excited, several magnetic islands emerge and mutually interact during the nonlinear evolution, e.g. via the coalescence instability (cf. § 2.1), by moving along the neutral line. This dynamics is due to the superposition of several unstable modes with different wavenumbers and growth rates, among which the fastest mode is obviously the dominant one. In general, an aspect ratio $L/a$ of the order of ${\sim }20$ is sufficient to grant the instability of a wavenumber sufficiently close to the fastest growing mode Velli & Hood (Reference Velli and Hood1989) – see also figure 1 of Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020). In this scenario, the ‘soliton-like’ behaviour of magnetic islands, noted already by Biskamp (Reference Biskamp1982) in his early simulations, has earned them the name of ‘plasmoids’ (see, e.g. Biskamp & Welter Reference Biskamp and Welter1989; Biskamp Reference Biskamp1996; Shibata & Tanuma Reference Shibata and Tanuma2001; Loureiro et al. Reference Loureiro, Schekochihin and Cowley2007; Bhattacharjee et al. Reference Bhattacharjee, Huang, Yang and Rogers2009), by analogy with the similarly behaving ‘plasma blobs’, also termed ‘plasmoids’, that have been identified and referenced in astrophysical plasma literature since the end of the sixties (Cowling Reference Cowling1967). Plasmoid structures displaying ‘soliton-like’ features have been observed in space plasmas, in particular, in magnetotail reconnection processes (see, e.g. early works such as those by Hones (Reference Hones1979) and Birn & Hones (Reference Birn and Hones1981) or later works such as Hautz & Scholer (Reference Hautz and Scholer1987), Scholer (Reference Scholer1987) and also Lin, Cranmer & Farrugia (Reference Lin, Cranmer and Farrugia2008), for a review of the usage of the term plasmoid in this context) and – although in a different geometrical setting – they have been also observed in connection to coronal mass ejections, arguably induced by reconnection processes in the solar corona (see, e.g. early works such as Pneuman Reference Pneuman1983; Gosling & McComas Reference Gosling and McComas1987). In this sense, in the recent literature, the term ‘plasmoid regime’ of reconnection (Uzdensky, Loureiro & Schekochihin Reference Uzdensky, Loureiro and Schekochihin2010; Huang & Bhattacharjee Reference Huang and Bhattacharjee2010; Comisso et al. Reference Comisso, Huang, Lingam, Hirvijoki and Bhattacharjee2018; Zhou et al. Reference Zhou, Wu, Loureiro and Uzdensky2021) is typically used to identify both tearing reconnection in the $\varDelta '\delta \sim 1$ wavelength limit (cf. the early numerical study by Biskamp (Reference Biskamp1986) on the instability of a Sweet–Parker current sheet to tearing modes) and the generation of magnetic islands in current sheets developed as a consequence of a turbulent motion. Here, the formation of ‘plasmoids’ is observed, either numerically (see Biskamp & Welter Reference Biskamp and Welter1989; Biskamp & Bremer Reference Biskamp and Bremer1994; Wei et al. Reference Wei, Hu, Feng and Schwen2000; Servidio et al. Reference Servidio, Matthaeus, Shay, Cassak and Dmitruk2009, Reference Servidio, Dmitruk, Greco, Wan, Donato, Cassak, Shay, Carbone and Matthaeus2011; Wan et al. Reference Wan, Matthaeus, Servidio and Oughton2013; Franci et al. Reference Franci, Cerri, Califano, Landi, Papini, Verdini, Matteini, Jenko and Hellinger2017 and several more recent works) or by experimental measurements (see, e.g. Nishizuka et al. Reference Nishizuka, Karlicky, Janvier and Bárta2015; Kumar et al. Reference Kumar, Karpen, Antiochos, Wyper and DeVore2019; Yan et al. Reference Yan, Xue, Jiang, Priest, Kliem, Yang, Wang, Kong, Song and Feng2022). We recall in this regard that also the term ‘plasmoid instability’ has been used in recent literature, with reference to the tearing instability associated with the fastest growing mode, which is destabilised on a Sweet–Parker current sheet of length L, and for which a scaling $\gamma _M\sim S_{L}^{1/4}$ is obtained (Tajima & Shibata Reference Tajima and Shibata1997; Loureiro et al. Reference Loureiro, Schekochihin and Cowley2007; Bhattacharjee et al. Reference Bhattacharjee, Huang, Yang and Rogers2009). The diverging scaling with respect to a vanishing non-ideal parameter – otherwise impossible for a spontaneous reconnecting instability – is here due to the fact that the fastest growing mode develops in this case as a ‘secondary’ reconnection process to a primary steady reconnection scenario (cf. figure 2): for $S_{L}^{-1}=0$, the Sweet–Parker reconnection does not occur, by thus forbidding the onset of the tearing modes. Due to the violently unstable nature of the tearing mode developing in the Sweet–Parker scenario, the possibility to measure, in Nature, a steady, Sweet–Parker-like reconnecting current sheet has been first questioned and discussed by Pucci & Velli (Reference Pucci and Velli2014).

Interpreting the magnetic islands observed in current sheets developed by turbulence as due to tearing mode instabilities is an hypothesis which is grounded on the assumption that a standard Fourier representation can be used in a Wentzel–Kramers–Brillouin (WKB)-sense also for reconnecting modes in a non-periodic, large aspect ratio current sheet. In this case, the fastest growing mode can be assumed to be, in an asymptotic sense, the dominant reconnecting instability. Although a proof of the correctness of this assumption for finite length current sheets has not been provided yet, this kind of approach has attracted since the beginning much attention in the context of both space plasmas (see, e.g. Cross & van Hoven Reference Cross and van Hoven1971; Van Hoven & Cross Reference Van Hoven and Cross1971) and solar physics (see, e.g. Priest Reference Priest1976; Velli & Hood Reference Velli and Hood1989), and has been more recently been considered for applications to secondary instabilities to primary reconnection processes (see, e.g. Tajima & Shibata Reference Tajima and Shibata1997; Shibata & Tanuma Reference Shibata and Tanuma2001; Tanuma et al. Reference Tanuma, Yokoyama, Kudoh and Shibata2001; Loureiro et al. Reference Loureiro, Cowley, Dorland, Haines and Schekochihin2005; Daughton & Scudder Reference Daughton and Scudder2006; Drake et al. Reference Drake, Swisdak, Schoeffler, Rogers and Kobayashi2006; Loureiro et al. Reference Loureiro, Schekochihin and Cowley2007; Bhattacharjee et al. Reference Bhattacharjee, Huang, Yang and Rogers2009; Landi et al. Reference Landi, Del Zanna, Papini, Pucci and Velli2015; Tenerani et al. Reference Tenerani, Velli, Rappazzo and Pucci2015; Del Sarto et al. Reference Del Sarto, Pucci, Tenerani and Velli2016; Del Sarto & Ottaviani Reference Del Sarto and Ottaviani2017; Papini, Landi & Del Zanna Reference Papini, Landi and Del Zanna2019b; Singh et al. Reference Singh, Pucci, Tenerani, Shibata, Hillier and Velli2019) and to reconnection in turbulence (see, e.g. Pucci & Velli Reference Pucci and Velli2014; Loureiro & Uzdensky Reference Loureiro and Uzdensky2016; Tenerani et al. Reference Tenerani, Velli, Pucci, Landi and Rappazzo2016; Comisso et al. Reference Comisso, Huang, Lingam, Hirvijoki and Bhattacharjee2018; Papini et al. Reference Papini, Franci, Landi, Hellinger, Verdini and Matteini2019a; Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020; Kowal et al. Reference Kowal, Falceta-Gonçalves, Lazarian and Vishniac2020; Pucci et al. Reference Pucci, Velli, Shi, Singh, Tenerani, Alladio, Ambrosino, Buratti, Fox and Jara-Almonte2020; Schekochihin Reference Schekochihin2020; Tenerani & Velli Reference Tenerani and Velli2020; Franci et al. Reference Franci, Papini, Del Sarto, Hellinger, Burgess, Matteini, Landi and Montagud-Camps2022) in the context of the so-called ‘turbulent (or turbulence-mediated or turbulence-driven) reconnection’ scenario (Matthaeus & Lamkin Reference Matthaeus and Lamkin1986; Strauss Reference Strauss1988; Loureiro et al. Reference Loureiro, Uzdensky, Schekochihin, Cowley and Yousef2009; Matthaeus & Velli Reference Matthaeus and Velli2011; Schekochihin Reference Schekochihin2020) – not to be confused with the study of ‘turbulent-driven magnetic island’ in tokamaks (cf. end of § 2.1). This topic is not of concern here, where we consider strictly periodic current sheets only. We therefore direct the interested reader to the aforementioned references. A survey of the scalings of the fastest growing mode in different reconnection regimes (among which the collisionless regimes we consider in the following) for periodic current sheets can be found in (Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020).

In the remainder of the article, we just focus on the formal solution of the eigenvalue problems in a slab current sheet in the $\varDelta '\delta \ll 1$ and $\varDelta '\delta \gg 1$ limits. In this regard, we note that, in a few cases, a closed form of the integral (3.7) has been provided: in resistive and collisionless regimes, formulae involving integration in the Fourier space have been provided by Pegoraro & Schep (Reference Pegoraro and Schep1986), Pegoraro et al. (Reference Pegoraro, Porcelli and Schep1989), Schep et al. (Reference Schep, Pegoraro and Kuvshinov1994), Basu & Coppi (Reference Basu and Coppi1981) and Porcelli (Reference Porcelli1991). These calculations provide the available analytical formulae for the growth rates in the ‘internal-kink’ and ‘constant-$\psi$’ RMHD, collisionless regimes.

Table 1 summarises the known results of the boundary layer analysis for the regimes of reconnection of interest in this article. The general dispersion relation for RMHD tearing modes, from which the large- and small-$\varDelta '$ limits can be obtained, has been written as in Del Sarto & Ottaviani (Reference Del Sarto and Ottaviani2017, appendix C) in terms of a characteristic scale length $\delta _L$ that coincides with the relevant layer width $\delta$ in the large- and small-$\varDelta '$ regimes. In §§ 57, we will detail the analytical calculations that allow the relevant limits of these dispersion relations to be recovered.

Table 1. Known scalings available in the previous literature as obtained from a boundary layer analysis in the collisionless and resistive regimes. The general dispersion relation of Porcelli (Reference Porcelli1991) and Ottaviani & Porcelli (Reference Ottaviani and Porcelli1995) for $d_e^2\gg \rho _s^2$ and that of Porcelli (Reference Porcelli1991) and Comisso et al. (Reference Comisso, Grasso, Waelbroeck and Borgogno2013) for $\rho _s>d_e\neq 0$ are reported. In both cases, we have adopted the notation used in Del Sarto & Ottaviani (Reference Del Sarto and Ottaviani2017, appendix C), from which the small and large wavelength limits can be obtained once $\delta _L\to \delta$ is assumed and some estimation is made for $\delta =\delta _{LD}$ or $\delta =\delta _{SD}$. The general dispersion relation valid in all wavelength limits of the cold-electron resistive regimes (Ara et al. Reference Ara, Basu, Coppi, Laval, Rosenbluth and Waddell1978) and of the warm-electron resistive regimes (Pegoraro & Schep Reference Pegoraro and Schep1986) is not reported explicitly but it can be obtained from the collisionless cases previous substitution $d_e\to (S\gamma )^{-1/2}$. Note that, although available since the early solutions of the boundary layer calculations, the scalings of $\delta _{LD}$ and of $\delta _{SD}$ we have reported here in the different regimes have not been always pointed out in the related, reference works. These scalings correspond indeed to the characteristic width of the innermost boundary layer (i.e. $\delta _1$) in each reconnection regime, but the identification of the latter as the reconnecting layer width requires indeed further hypotheses which are discussed in § 8.

Finally, we draw attention to the fact that the definition (3.8) in a slab geometry changes meaning as soon as the condition $\varDelta '\delta \gtrsim 1$ is achieved, since application of (3.11) for $ka\ll 1$ and $\varDelta '\delta \sim 1$ formally implies a dependence of $k$ on $\delta$. In general, indeed, $(\varDelta ')^{-1}$ becomes a microscopic scale as soon as $\varDelta 'L_0\gg 1$, and for $\varDelta '\delta \gtrsim 1$, the condition $(\varDelta '(k))^{-1}\ll \delta$ is satisfied by a range of wavenumbers that depend on the non-ideal parameters which define $\delta$.

3.3 Boundary layer integration of collisionless tearing in the previous literature

In the following, we will discuss the method that can be used to find the analytical solution of the boundary layer problem associated with the scalings of table 1.

Differently from Pegoraro & Schep (Reference Pegoraro and Schep1986) and Porcelli (Reference Porcelli1991), who tackled the boundary layer calculations in the Fourier space, we are going to address the integration in the coordinate space: although a little more cumbersome, if one's interest is limited to obtaining the eigenvalue scaling alone, the analysis performed in the coordinate space allows a more intuitive understanding of the physics of the problem. Moreover, it spares one from the sometimes non-trivial task of reversing the Fourier transform so as to obtain the eigensolutions in the coordinate space. To the best of our knowledge, these kinds of calculations in the coordinate space for the warm-electron regime ($\rho _s\gtrsim d_e$) have never been reported in the literature, before.

Only in the work by Zocco & Schekochihin (Reference Zocco and Schekochihin2011) have some details of the analysis for the double boundary layer integration in the presence of two matching layers been discussed in the coordinate space (see appendix B therein). However, in that work, a different analytical approach has been taken for the integration of a different analytical reduced-MHD model including ion-FLR effects in the warm-collisionless regime. In particular, starting from gyrokinetic equations, these authors derived a reduced model consisting of three equations corresponding: to the scalar potential related to the ${\boldsymbol E} \times {\boldsymbol B}$ drift velocity ($\varphi$); to the parallel component of the vector potential ($A_{\|} = -\psi$); and to the ‘reduced’ electron distribution function ($g_e$, therein), which accounts for the moments of the electron distribution function, except for the density and the mean parallel velocity. The equation describing the evolution of $g_e$ is there coupled to the other two equations via the perturbation of the parallel electron temperature ($\delta T_{e,\|}$). In the linear limit, the latter can be written in terms of $\varphi$: this allowed the authors to obtain the customary system of two slab-geometry tearing equations for $\varphi$ and $\psi$, although modified with respect to (2.1)(2.2) we consider here, because of the contribution of the gyrokinetic ion operator that appears in the vorticity equation after integration of the ion gyrokinetic equation (Pegoraro & Schep Reference Pegoraro and Schep1981); this contribution clearly vanishes in the limit of null ion temperature in which we are interested. In this way, Zocco and Schekochihin looked for asymptotic analytical solutions of the boundary layer problem in the coordinate space by identifying two distinct layers: the electron and the ion regions whose widths are defined by the reconnection layer width $\delta$ and by the ion-sound Larmor radius ($\rho _s$), respectively. However, differently from the approach we are going to develop below, in which we will look for a direct integration of the asymptotic solutions in the warm-collisionless regime, these authors solved the linear equations in the two asymptotic regions for the warm-collisionless regime by using perturbation methods (Zakharov & Rogers Reference Zakharov and Rogers1992). By matching the asymptotic solutions, they then recovered the scaling laws first evaluated in the Fourier space (Porcelli Reference Porcelli1991).

In the remainder of the article, we will treat both the collisionless and resistive limits in a unified way, by separately considering the warm ($\rho _s\gtrsim d_e$) and cold ($d_e\gg \rho _s$) regimes. Then, we will specify the results of the dispersion relation in each reconnection regime. We will thus reobtain the resistive scaling laws first computed by Furth et al. (Reference Furth, Killeen and Rosenbluth1963) in the cold-electron regime and by Pegoraro & Schep (Reference Pegoraro and Schep1986) in the warm-electron regime, and the collisionless scaling laws first evaluated by Coppi (Reference Coppi1964c,Reference Coppia) in the cold-electron limit and by Porcelli (Reference Porcelli1991) in the warm one.

In Appendix C, we make a brief review of previous works which have approached boundary layer calculations in the Fourier space (included are some more recent works which explicitly made a comparison between eigenmode solutions in the real and in the Fourier space – Connor et al. Reference Connor, Hastie and Zocco2012b), and we compare some key points of that procedure with the integration procedure that we follow in this work.

4 Boundary layer solution in the outer ideal region

We now turn our attention to solve the linearised equations in the ideal MHD region.

In the outer region and sufficiently far from the reconnection layer, a steady force balance condition in ideal MHD can be assumed. Therefore, the terms weighted by the non-ideal parameters can be neglected. In this case, using $\psi _1 =\psi _{\text {out}}, \ \varphi _1=\varphi _{\text {out}}$, (2.1)(2.2) become

(4.1)\begin{gather} \varphi_{\text{out}} ={-}\frac{{\rm i}\gamma}{k\psi_0'}\psi_{\text{out}}, \end{gather}
(4.2)\begin{gather} \psi''_{\text{out}} = \left( k^2 + \frac{\psi^{'''}_{0}}{\psi'_0} \right) \psi_{\text{out}}. \end{gather}

Equation (4.1) says that $\psi _1$ and $\varphi _1$ have opposite spatial parity and are de-phased by ${\rm \pi} /2$, i.e. $\psi _1\sim \cos (ky)$ versus $\varphi _1\sim \sin (ky)$ if $y=0$ is assumed at an $O$-point.

It is also evident that solving (4.2) gives the outer solution ($\psi _1=\psi _{\text {out}}$) which is unequivocally determined by the equilibrium profile $\psi _0$ and by $k$. Using (4.1), one can so obtain also the eigenmodes $\varphi _{\text {out}}$. The number of inner regions depends instead on the number of non-ideal parameters involved in the problem.

A class of magnetic equilibria of particular interest is provided by those fulfilling the condition $\lim _{|x|\to \infty } (\psi ^{'''}_{0}/\psi '_0)={\rm constant}$, for which a quite general integration procedure can be devised, as we are going to discuss in the specific example below.

4.1 A specific example of evaluation of $\varDelta '(\psi _0;k)$

Having in mind the numerical results to be presented in the next sections, let us solve the outer equations for the equilibrium profile

(4.3)\begin{equation} \psi'_0(x) = \frac{\tanh(x)}{\cosh^2(x)}, \end{equation}

which is the one used by Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) to obtain the scalings to which we will make comparison. This profile was first proposed by Porcelli et al. (Reference Porcelli, Borgogno, Califano, Grasso, Ottaviani and Pegoraro2002). Substituting the previous relation into (4.2), one finds

(4.4)\begin{equation} \psi''_{\text{out}} - \left(\alpha^2 - \frac{12}{\cosh^2(x)}\right) \psi_{\text{out}} = 0, \end{equation}

where $\alpha ^2 = k^2 + 4$ and ${\psi ^{'''}_0}/{\psi ^{'}_0} = 4 - {12}/{\cosh ^2(x)}$. Grasso first evaluated the corresponding $\varDelta '$, reported by Porcelli et al. (Reference Porcelli, Borgogno, Califano, Grasso, Ottaviani and Pegoraro2002), by looking for a polynomial solution expressed in powers of $\tanh (x)$ (D. Grasso, private communication 2017). Here, we formally revise the problem, and we reduce (4.4) to a Legendre-type equation that can be shown to be valid for different kinds of equilibria, which we will discuss elsewhere.

First, we notice that $\lim _{x\to \infty } \cosh ^{-2}(x)=0$. This represents a main requirement if one seeks to apply periodic boundary conditions on the problem by considering $\psi _0$ as a periodic function in $x$ with spatial period of infinite extension. Therefore, far enough from the neutral line where the reconnection event takes place, the hyperbolic term in (4.4) can be neglected and (4.4) becomes

(4.5)\begin{equation} \psi''_{f} - \alpha^2 \psi_f = 0, \end{equation}

whose solution is

(4.6)\begin{equation} \psi_f(x) = A e^{-\alpha x}, \end{equation}

with $A$ an integration constant. One then expects the solution of (4.4) to take the form

(4.7)\begin{equation} \psi_{\text{out}}(x) = \psi_f(x) f(x) = A e^{-\alpha x} f(x). \end{equation}

Substituting the previous equation in (4.4), one obtains

(4.8)\begin{equation} f'' - 2 \alpha f' + \frac{12}{\cosh^2(x)} f = 0. \end{equation}

Motivated by the fact that $\tanh '(x) =\cosh ^{-2}(x)$, we perform the change of variables $z=\tanh (x)$. Then, (4.8) for $f(z)$ reads

(4.9)\begin{equation} (1 - z^2) f'' - 2 (z + \alpha) f' + 12 f = 0, \end{equation}

where the ‘$'$’ refers to the derivation with respect to $z$. The neutral line is now at $z = \tanh (x) = 0$. When $\alpha =0$, (4.9) becomes a Legendre equation of order three (we recall that a Legendre equation of order $p$ reads $(1-z^2)f''-2zf'+p(p+1)f=0$ – see Abramowitz & Stegun Reference Abramowitz and Stegun1964, § 8.1.1). This equation has two regular singular points occurring at $z=\pm 1 \ (x \to \pm \infty )$, and one of its two linearly independent solutions is a polynomial of third degree. This reads

(4.10)\begin{equation} f = a + b z + c z^2 + d z^3, \end{equation}

where the coefficients can be found by substituting the polynomial solution into (4.9) and equating to zero the coefficients with equal powers of $z$. After obtaining these coefficients, using $z = \tanh (x)$ and (4.7), the outer solution of (4.2) becomes

(4.11)\begin{align} \psi_{\text{out}} = A e^{-\alpha |x|} \left\{ 1 + \frac{6\alpha^2 - 9}{\alpha (\alpha^2 - 4)} \tanh(|x|) + \frac{15}{(\alpha^2 - 4)} \tanh^2(|x|) + \frac{15}{\alpha (\alpha^2 - 4)} \tanh^3(|x|) \right\}. \end{align}

The equation in the outer region accepts both an even and an odd solution. The even solution is the one required for tearing modes (the magnetic island shape is symmetric with respect to the neutral line – cf. also (2.8)). This solution displays a discontinuity in $\psi '_{\text {out}}$ at $x = 0$. Taking the limit of the first derivative of (4.11),

(4.12)\begin{equation} \lim_{x\to\pm 0} \psi'_{\text{out}} = A \left \{{\mp} \sqrt{k^2 + 4} \pm \frac{6k^2 + 15}{k^2 \sqrt{k^2+4}} \right \}, \end{equation}

we obtain, using (3.8),

(4.13)\begin{equation} \varDelta' = 2 \frac{-k^4 + 2k^2 + 15}{k^2 \sqrt{k^2 + 4}}. \end{equation}

It follows that $\varDelta ' > 0$ when $k \in [0, \sqrt {5}]$, which represents the spectrum of unstable wavenumbers to tearing-type instabilities for the equilibrium profile given by (4.3).

The outer solution for the case in which the magnetic equilibrium is given by Harris’ profile $\psi _0'(x) = \tanh (x)$ (Harris Reference Harris1962) can be obtained in the same manner, as already noted by Furth (Reference Furth1963) (§ 4, therein). In this case, the equivalent of (4.9) is a Legendre equation of degree $p=2$. Therefore, $f(z) = a + b z$, which leads to an even solution $\psi _{\text {out}}(x) = \psi _0 e^{-k|x|} \{ k + \tanh (|x|) \}$. Following the same arguments as before, one recovers the result $\varDelta ' = 2 ({1}/{k} - k)$, where the instability condition $\varDelta ' > 0$ is met when $k\in [0,1]$ (Furth et al. Reference Furth, Killeen and Rosenbluth1963). The point we underline here is that the solution of the outer equation and the evaluation of $\varDelta '$ by splitting the problem in (4.5)(4.6) can be, in principle, used for a quite large class of magnetic equilibria.

As anticipated, the condition $\varDelta '>0$ determines the spectrum of unstable wavenumbers for a given equilibrium profile. This can be formally seen as related to the change of concavity that the solution $\psi _1$ must undergo while moving from the outer to the inner region, for the singular eigenfunction obtained in the ideal-MHD limit to become ‘regular’, thanks to the presence of non-ideal effects that allow reconnection (Furth et al. Reference Furth, Killeen and Rosenbluth1963). In general, exponentially growing solutions of the linear problem are obtained when $\varDelta '>0$ (cf. figure 7).

Figure 7. Profile of the eigenfunction $\psi _1$ in the purely resistive regime, computed via the numerical solver of Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) for $S^{-1}=10^{-5}$ and $k=10$ (green curve), $k=1.7$ (orange curve) and $k=0.1$ (blue curve). These three values are respectively exemplificative of: the stable regime ($\varDelta '=-19.19<0$, green curve); of the small-$\varDelta '$ wavelength limit ($\varDelta '=3.2766$, red curve); and of the large-$\varDelta '$ wavelength limit ($\varDelta '=48.9$, blue curve). In the stable case ($\varDelta '<0$), the eigenfunction does not change concavity as $x$ approaches the neutral line, whereas for the unstable tearing-type solutions obtained for $\varDelta '>0$, the value $c_0\equiv \lim _{x\to 0}\psi _1(x)$ decreases approaching zero as the value of $\varDelta '(k)$ increases approaching $+\infty$.

5 Equations in the non-ideal region

After having solved the linearised equations in the ideal MHD region, we are now interested in finding their solutions in the non-ideal region. We thus begin this section by introducing a ‘generalised resistivity’, which allows us to cast the linear equations in a unique form that can be applied to both the collisionless and resistive regimes. For analytical purposes that will become evident in the following, in this section, we also normalise the linear system with respect to an arbitrary characteristic length – say $\ell$ – which will be later specified in each subdomain and in each regime. Then, we will introduce an auxiliary function $\chi$ related to both scalar fields $\psi _1$ and $\varphi _1$, which allows us to combine the inner layer equations into a single equation. It is this equation that, in some reconnection regimes, will be later approximated in each subdomain of the boundary layer approach and which will be solved analytically according to the strategy detailed in § 3.

For the analytical integration of the linear problem, it is first useful to introduce a parameter that includes both the resistive and inertia-related non-ideal terms in Ohm's law by thus highlighting their ‘almost symmetric’ contributions in (2.1). This can be equivalently done in terms of the ‘generalised resistivity’, as first suggested by Furth (Reference Furth1963) (§ 5 therein), which, in the normalised units we use, reads

(5.1)\begin{equation} \bar{S}^{{-}1}\,\equiv \, S^{{-}1}+\gamma d_e^2, \end{equation}

or in the form of a ‘generalised electron skin depth’, as done by Porcelli (Reference Porcelli1991) (note that Porcelli (Reference Porcelli1991) uses the symbol $\varDelta$ in place of the $\bar {d}_e$ we adopt here), by expressing the Lundquist number in terms of the normalised electron–ion collision rate, $\nu _{ei}$, as $S^{-1}=d_e^2\nu _{ei}/2$, so as to write

(5.2)\begin{equation} \bar{d}_e^2\,\equiv d_e^2+ \frac{S^{{-}1}}{\gamma}=d_e^2\left(1+ \frac{1}{2} \frac{\nu_{ei}}{\gamma}\right). \end{equation}

In either case, the second left-hand side term and the last right-hand side term of (2.1) can be re-absorbed into a single contribution that formally accounts for both resistive and inertial effects, the latter being therefore interpretable, in the form $\gamma d_e^2$, as a ‘collisionless resistivity’. Using for example (5.2), we rewrite (2.1) as

(5.3)\begin{equation} \gamma [\psi_1-\bar{d}_e^2(\psi_1''-k^2\psi_1)] -{\rm i}k\varphi_1 (\psi_0'-{d}_e^2\psi_0''')={-}{\rm i}k{\rho}_s^2(\varphi_1''-k^2\varphi_1)\psi_0'. \end{equation}

In the following, when we will analytically integrate the linear problem in the non-ideal region, it will be more convenient to start from (2.2) and (5.3), of which we will then take the appropriate collisional or collisionless limits. Specialising the parameters $\bar {S}^{-1}$ or $\bar {d}_e$ to the purely resistive or purely collisionless limits is meaningful for linear tearing modes, because of the extremely narrow region of the parameter space in which both $S^{-1}$ and $d_e$ appreciably contribute to the asymptotic scalings (Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020). A transition from resistive to inertia-dominated regimes can instead be relevant to the onset of secondary collisionless modes over primary resistive tearing modes (Del Sarto & Ottaviani Reference Del Sarto and Ottaviani2017), provided the applicability of the WKB approximation we spoke of in § 3.2 for secondary tearing instabilities developing on finite length, large-aspect ratio current sheets generated by the primary reconnection event. In this case, the possibility of having a transition of regime depends on both the rescaling of the shear length and of the magnetic field amplitude of reference (see Del Sarto et al. Reference Del Sarto, Ottaviani, Pucci, Tenerani and Velli2018).

A further distinction will be made between the ‘cold’ regime $\bar {d}_e^2\gg \rho _s^2$ (discussed in § 6) and the opposite ‘warm’ regime $\rho _s^2\gg \bar {d}_e^2$ (discussed in § 7). The reason why we are going to treat the two cases separately is due to the fact that the normalisation scales that define the extent, in an asymptotic sense, of the innermost boundary layer domain (i.e. $\delta _1$ of figure 4b) turn out to be different, depending on whether $\rho _s$ is negligible or not with respect to $d_e$ (or to $(S^{-1}/\gamma )^{1/2}$). The asymptotic scaling of the layer width, $\delta _1$, also determines, via the matching conditions, the scalings of the reconnection rate $\gamma$. It follows that the ‘cold’ collisionless solution cannot be obtained as a trivial limit $\rho _s\to 0$ of the solution obtained for the ‘warm’ case.

Operationally speaking, we will solve the inner equations using the generalised electron skin depth of (5.3), so we will first formally recover the collisionless results of Coppi (Reference Coppi1964c), Coppi (Reference Coppi1964a) and Porcelli (Reference Porcelli1991). Then, we will comment about the correspondence of these results with those obtained by Furth et al. (Reference Furth, Killeen and Rosenbluth1963), Coppi et al. (Reference Coppi, Galvao, Pellat, Rosenbluth and Rutherford1976), Ara et al. (Reference Ara, Basu, Coppi, Laval, Rosenbluth and Waddell1978), Pegoraro & Schep (Reference Pegoraro and Schep1986) and Pegoraro et al. (Reference Pegoraro, Porcelli and Schep1989) when resistivity dominates over electron inertia.

5.1 Differential equations in the non-ideal region

The fact non-ideal effects are important in a region which is microscopic with respect to the equilibrium shear scale $a$, allows us to use (3.2) so as to write $\psi _0'(x)\approx 2x C_2= x\, \psi _0''|_{x=0}\equiv x J_0$, where we have named $J_0 \equiv \psi _0''(x)|_{x=0}$ the value (normalised to $B_0$ and $L_0=a$) of the second derivative of the equilibrium flux function at the neutral line. Using the ordering given by (3.4), (5.2)(2.1) become

(5.4)\begin{gather} \gamma \psi_1 - \gamma \bar{d_e}^2 \psi_1'' = {\rm i} k J_0 x \varphi_1 - {\rm i} k J_0 \rho_s^2 x \varphi_1'', \end{gather}
(5.5)\begin{gather} \gamma \varphi_1'' = {\rm i} k J_0 x \psi_1''. \end{gather}

It can be noted that a direct comparison between the relative weight of the parameters $\bar {d_e}^2$ and $\rho _s^2$ in (5.4) can be done after eliminating $\varphi _{1}''$ in (5.4) via (5.5). Doing so, (5.4) takes the form

(5.6)\begin{equation} \gamma \psi_1 = {\rm i} k J_0 x \varphi_1 + \gamma\left(\bar{d_e}^2 + \rho_s^2\frac{ k^2 J_0^2}{\gamma^2} x^2 \right)\psi_1''. \end{equation}

The coefficient multiplying $\psi _1''$ at the right-hand side of (5.6) can be read as a generalised, space-dependent, ‘Lundquist number’, i.e. it can be assimilated to a space-dependent resistivity/conductivity. (In more complex reconnection models based on a gyrokinetic modelling and including also ion-FLR effects, it can be shown (Cowley et al. Reference Cowley, Kulsrud and Hahm1986; Zocco & Schekochihin Reference Zocco and Schekochihin2011) that the generalised conductivity, which rules the collisionless reconnection process inside the innermost layer dominated by electron dynamics, takes a rational-polynomial form of the kind $\sigma _e= ({a+b (x/ \delta _1)})/({c+d (x/ \delta _1)^2 + e(x/ \delta _1)^4})$, where $\delta _1$ is the innermost reconnecting layer width and $a, b, c, d$ and $e$ are coefficients that depend on the plasma parameters. This form has been used by Connor et al. (Reference Connor, Hastie and Zocco2012b) to obtain, via Fourier-space integration, a general dispersion relation encompassing drift-tearing modes, kinetic Alfvén modes and the internal kink mode at low values of the plasma $\beta$ parameter.) To have magnetic reconnection, i.e. to allow for the existence of tearing-type unstable modes, the coefficient $\bar {d_e}^2$ must be non-null since $\rho _s^2$ alone does not allow the relaxation of the topological constraints forbidding the intersection and breaking of initially distinct magnetic lines (cf. § 2.1). Once the importance of the $\bar {d_e}^2$ term in the non-ideal region is established, one sees that the condition for which the last term in parenthesis of (5.6) is negligible in a subdomain of the integration domain, even for a non-vanishing value of $\rho _s$, is

(5.7)\begin{equation} \frac{\rho_s^2}{\bar{d_e}^2}\ll \left(\frac{\gamma}{ k x J_0}\right)^2. \end{equation}

This assumption will be later done in some regimes and will be heuristically verified.

5.2 Normalisation and distinction of the boundary layers

More in general, to solve the system (5.4)(5.5) with the boundary layer approach, further hypotheses are done about the relative magnitude of the different contributions inside of the inner layer, and some auxiliary function (traditionally noted as $\chi$) relating $\varphi _1$ to $\psi _1$ via (5.5) is introduced, depending on the $\varDelta '$-regime, so as to bring the system to a single ordinary differential equation (ODE). In each regime, appropriate normalisation of the spatial scales can be therefore chosen so as to better identify the extension of the sub-intervals of the inner region, where some term dominates over or is negligible with respect to the others. This is why, in the following, in the reconnection regimes and $\varDelta '$ limits which will be of interest, at each time, we will perform several changes of variable based on different normalisation choices.

This is an important step in the boundary layer procedure, since it is at this level that one introduces the layer widths $\delta _1$ or $\delta _2$ (cf. figure 4), although only implicitly: in practice, a ‘stretched’ coordinate is introduced by referring the coordinate $x$ to a characteristic length, say $\ell$, which replaces $x$ in different subdomains of the non-ideal region. It is when the limits $|x|/\ell \gg 1$ or $|x|/\ell \ll 1$ will be taken for the purpose of finding solutions of the differential equations within the boundary layer approach that the length $\ell$ will be identified as the layer width, e.g. $\delta _2$ or $\delta _1$. Also note that the length $\ell$ may initially depend on some unknown parameter like the eigenvalue $\gamma$: its scaling, and therefore that of $\ell$ and subsequently that of the layer width to which it corresponds, will be therefore determined a posteriori, from the conditions on the solutions obtained by integrating and matching the equations for small and large values of these stretched variables. Note that since the initial choice of each length scale $\ell$ is essentially arbitrary, it must be a posteriori verified for consistency that, asymptotically, $\delta _1\ll \delta _2 \ll 1$ (when both $\delta _2$ and $\delta _1$ are expressed in units of $L_0=a$).

It is also worth stressing that the identification of each boundary layer is subordered to the identification of the negligibility of some term of the non-ideal equations. This means that we must have knowledge of the relative ordering of the non-ideal parameters at play (for example, had we wanted to include viscosity, a third scale length – let us say $\lambda _{\nu }$ – would have appeared. If the hypothesis $\lambda _{\nu }\gg \rho _s \gtrsim \bar {d}_e$ had been fulfilled, one could have a priori expected three subdomains in the non-ideal region: one where ${\nu }$ alone dominates, one where both $\nu$ and $\rho _s$ in principle dominate, and one where all three non-ideal parameters are in principle important. Then, one should evaluate if, in the two innermost regions, specific conditions can be satisfied so that the dominant contribution of one or two non-ideal terms alone can be isolated: this possibly refines further the subdomains of interest for the purpose of the integration. It is at this level that we finally define the boundary layers, with respect to which the matching conditions will be fixed), or, if this is not the case, specific assumptions must be done on them by separately considering the different possible combinations, until in each case, the ‘thinnest’ sub-domain is this way singled out: this is the innermost region where the eigenmode solution must be first integrated by imposing the boundary conditions from outside.

As an example of general interest for the calculations that we are going to develop next, we can consider a normalisation of (5.4)(5.5) to an arbitrary length $\ell$, which we will specify later, in each case examined. This will also let us get rid of some ‘superfluous’ parameters in the equations. We thus define

(5.8)\begin{equation} \zeta = \frac{x}{\ell}, \quad \mathcal{G}=\frac{\gamma}{k \ell J_0},\quad \tilde{\varphi}_1 ={-}{\rm i}\frac{\gamma }{k \ell J_0} \varphi_1(\zeta), \quad \tilde{\psi}_1 = \psi_1(\zeta), \end{equation}

which allow us to re-write (5.4)(5.5) as

(5.9)\begin{gather} \tilde{\psi}_1 - \left(\frac{\bar{d}_e^2}{\ell^2} + \frac{\rho_s^2}{\ell^2} \frac{\zeta^2}{\mathcal{G}^2}\right)\tilde{\psi}_1'' ={-}\frac{\zeta}{\mathcal{G}^2}\, \tilde{\varphi}_1, \end{gather}
(5.10)\begin{gather} \tilde{\varphi}_1'' = \zeta \tilde{\psi}_1'', \end{gather}

where both functions $\tilde {\varphi }_1$ and $\tilde {\psi }_1$ depend now on $\zeta$.

We note that $\mathcal {G}$ depends on $\gamma$, which generally makes both quantities complex numbers: they are real for usual tearing-type instabilities, which, in the absence of diamagnetic effects related to equilibrium density gradients, do not propagate; instead, they are purely imaginary if the magnetic profile is stable. Because of this feature, in the following, it will be useful to perform some integration in the complex plane.

Also note that if $\ell =1$ (in units of $L_0=a$), then, in dimensional units, we can write $\mathcal {G} =\gamma \tau _0/(ka)$, where $\gamma \tau _0$ expresses the growth rate of the unstable mode measured with respect to the ‘natural timescale’ $\tau _0\sim \sqrt {mn_0/(4{\rm \pi} )}(c/J_0)$, which equals here the transition time of a shear Alfvén wave across the shear length $a$ (cf. Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020). Therefore, $\mathcal {G}|_{\ell =1} \ll 1$ in an asymptotic sense, as long as the wavenumber is fixed and then $ka$ is unordered with respect to $\gamma \tau _0$. If, instead, the scale $\ell$ is microscopic, the ordering of $\mathcal {G}$ with respect to unity must be a posteriori evaluated.

5.3 Auxiliary function ‘$\chi$’ and large- and small-$\varDelta '$ limits

The strategy by which the inner equations are integrated depends on mathematical features that we are going to discuss in detail in each specific regime and wavelength limit on which they depend (see also § 5.5). In general, some approximations and hypotheses are required, since they allow us to combine (5.9)(5.10) into a single equation. An integration strategy which is particularly efficient, especially in the large-$\varDelta '$ limit, consists in casting the equations for $\tilde{\psi}_1$ and $\tilde {\varphi }_1$ into an equation for an auxiliary variable, even with respect to $\zeta$, which is typically noted $\chi (\zeta )$, after the notation first chosen for it by (Coppi et al. Reference Coppi, Galvao, Pellat, Rosenbluth and Rutherford1976; Ara et al. Reference Ara, Basu, Coppi, Laval, Rosenbluth and Waddell1978)

(5.11)\begin{equation} \chi(\zeta) \equiv \zeta \tilde{\psi}_1'(\zeta) - \tilde{\psi}_1(\zeta). \end{equation}

This is directly connected to (5.10), of which it is an integral. Indeed, we have

(5.12)\begin{equation} \tilde{\varphi}_1''(\zeta) = \chi'(\zeta) \quad \implies \quad \tilde{\varphi}_1'(\zeta) = \chi(\zeta) - \lim_{\zeta\to\infty}\chi(\zeta) = \chi(\zeta) - \chi_\infty. \end{equation}

In the last passage, we have named $\chi _\infty$ the value of $\chi (\zeta )$ as $\zeta \to \infty$. Note that in boundary layer calculations, when $\ell$ will be identified as $\delta _2$ or $\delta _1$, the limit $\zeta \to \infty$ for which $\chi _\infty$ is defined must be taken in the matching layer outside of the innermost layer, but still inside the non-ideal region, i.e. for $\delta _1\ll x\ll 1$ (in units of $L_0=a$): since $\tilde {\varphi }_1'$ represents the $y-$component of the velocity, it approaches zero as $\zeta \to \infty$ inside the ideal region. Determining the asymptotic behaviour of $\chi _\infty$ at both limits of the unstable wavelength spectrum is important since the scaling laws of the eigenvalue problem depend on it (Ara et al. Reference Ara, Basu, Coppi, Laval, Rosenbluth and Waddell1978). In practice, depending on whether $\chi _\infty$ is zero or not, the differential equation for $\chi$ is homogeneous or not (see (5.19) below). These topics are discussed in the remainder of this section by considering, for simplicity, the case of a single boundary layer (i.e. assuming $\ell =\delta _1$).

5.3.1 Asymptotic estimate of $\chi _\infty$ in the small-$\varDelta '$ limit

In the case of the small-$\varDelta '$ limit and when $x \to 0$, according to the definition of $\chi (\zeta )$ and to the matching condition $\lim _{\zeta \to \infty }\tilde {\psi }_1(\zeta )\sim \lim _{x\to 0}\psi _{\text {out}}(x)$, one finds

(5.13)\begin{equation} \chi_\infty=\lim_{\zeta\to\infty}\chi(\zeta) \sim \lim_{x\to 0}x \psi_{\text{out}}'(0) - \psi_{\text{out}}(0) \simeq{-} c_0, \quad \text{(constant-$\psi$ ordering)}. \end{equation}

In this case, we have used the ‘constant-$\psi$’ approximation (Furth et al. Reference Furth, Killeen and Rosenbluth1963), i.e. the fact that in the small-$\varDelta '$ regime, the inner solution tends to a constant value $\psi _1(0)$ when it approaches the innermost region. Such constant value is approximatively obtained already from the $x\to 0$ limit of the outer solution $\psi _{\text {out}}(x)$, so that in this wavelength limit, we can write $\lim _{x\to 0}\psi _{\text {out}}(x) = c_0$.

The constant-$\psi$ ordering can be assumed and heuristically verified after integration of the eigenmode problem for the single boundary layer limit (cf. Furth et al. Reference Furth, Killeen and Rosenbluth1963; White Reference White1983), and in the small-$\varDelta '$ regime, it can be shown to be equivalent to the condition $\delta _1 \varDelta '\ll 1$. In this case, we can also estimate $\lim _{x\to 0}x \psi _1'(x)\sim \lim _{x\to 0} x\varDelta ' \psi _{\text {out}}(x)\sim \delta _1 \varDelta 'c_0$, where in the last passage, we have (over)estimated $\lim _{x\to 0}x$ by the extent of the innermost layer width: owing to the $\varDelta ' \delta _1\ll 1$ condition, the last passage of (5.13) is thus justified, whence one deduces

(5.14)\begin{equation} \chi_\infty ={-}c_0, \quad \text{for } \delta_1 \varDelta' \ll 1. \end{equation}

5.3.2 Asymptotic estimate of $\chi _\infty$ in the large-$\varDelta '$ limit

In the large-$\varDelta '$ limit for $x\to 0$, instead, one finds that $\chi _\infty =0$. This has been first discussed by Coppi et al. (Reference Coppi, Galvao, Pellat, Rosenbluth and Rutherford1976) and Ara et al. (Reference Ara, Basu, Coppi, Laval, Rosenbluth and Waddell1978), where it has been shown that the vanishing of $\chi (\zeta )$ far out from the resonant region (equivalent to the innermost layer in our notation) is consistent with the rigid plasma displacement that characterises the internal kink mode in a cylindrical tokamak: in the notation and slab geometry assumption we use in this work, this corresponds to write $\tilde {\varphi }_1(\zeta )=-\gamma \zeta$ for $|\zeta |\leq 1$ and $\tilde {\varphi }_1(\zeta )=0$ for $|\zeta |> 1$ (cf. also Porcelli Reference Porcelli1987; Del Sarto & Ottaviani Reference Del Sarto and Ottaviani2017), whence the condition $\chi _\infty =0$ immediately follows by direct comparison with the definition in (5.12). An alternative way to look at the consistency of this result is to consider the Taylor expansion of the outer solution given by (4.7):

(5.15)\begin{equation} \psi_{\text{out}}(x) \sim c_0 + c_1 |x| \implies \psi_{\text{out}} \sim c_0 + c_1 |x| \implies \varDelta' \sim \frac{2c_1}{c_0}. \end{equation}

This solution formally holds in the outermost matching layer, where the instability parameter $\varDelta '$ is defined by neglecting non-ideal effects, i.e. by taking the limit $x\to 0$ of the eigenmode in the ideal region. Therefore, the use of $\lim _{x\to 0}\psi _{\text {out}}(x)$ to match the solution $\lim _{\zeta \to \infty }\psi _{1}(\zeta )$ is a priori not always justified. This however is not the case of the large-$\varDelta '$ limit, in which it can be a posteriori shown, from heuristic arguments (see Ottaviani & Porcelli (Reference Ottaviani and Porcelli1995) for the ‘cold’ reconnection regime) or by numerical integration, that the discontinuity in the derivative of $\psi _{\text {out}}$ occurs in a position which gets progressively closer to the neutral line, as long as the numerical value of $\varDelta '(k)$ increases (cf. also §§ 810). In this limit, using therefore $\lim _{\zeta \to \infty }\tilde {\psi }_1\simeq \lim _{x\to 0}\psi _{\text {out}}(x)$ combined with (5.15) in the definition of (5.11), one finds

(5.16)\begin{equation} \lim_{\zeta\to\infty}\chi(\zeta) \sim c_1 |x| - c_0 - c_1 |x| ={-}c_0 \sim \frac{c_1}{\varDelta'}. \end{equation}

The fact that $c_0\to 0$ as $\varDelta '$ increases is shown in figure 7, where the spatial profile of $\psi _1$, obtained after numerical integration of the purely resistive regime (for $S^{-1}=10^{-5}$), is shown for the values of $\varDelta '(k)= -19.19$ (green curve), $\varDelta '(k)= 48.9$ (blue curve) and $\varDelta '(k)= 3.2766$ (orange curve), evaluated using (4.13) for $k=10, 0.1$ and $1.7$, respectively. Thus, in the large-$\varDelta '$ limit where $\varDelta '\to \infty$, (5.16) implies that

(5.17)\begin{equation} \chi_\infty = 0 \quad \text{for } \varDelta' \to \infty. \end{equation}

The fact that condition $\varDelta ' \to \infty$ can be replaced by $\varDelta '$ larger than the inverse of some characteristic scale length will be discussed in § 10. As already anticipated (§ 3.2), this condition can be generally read as $\varDelta '\delta _1\gg 1$.

5.4 Auxiliary equation for the function $\chi (\zeta )$ in the non-ideal region

Definitions in (5.11)(5.12) can be used to cast (5.9) in an equation for $\chi (\zeta )$. For analytical convenience, i.e. to facilitate the substitution of variables, it is preferable to evaluate the derivative of (5.9) after it has been divided by $\zeta$. After opportunely regrouping the terms of the equation so obtained, the result reads

(5.18)\begin{equation} (\zeta\tilde{\psi}_1' - \tilde{\psi}_1) - \frac{\bar{d}_e^2}{\ell^2} (\zeta\tilde{\psi}_1'''+\tilde{\psi}_1'' -2\tilde{\psi}_1'') -\frac{1}{\mathcal{G}^2}\frac{\rho_s^2}{\ell^2}\, \zeta^2\, (\zeta \tilde{\psi}_1'''+\tilde{\psi}_1'')={-}\frac{1}{\mathcal{G}^2}\,\zeta^2 \tilde{\varphi}_1', \end{equation}

which can be rewritten as

(5.19)\begin{equation} \left(\frac{\bar{d}_e^2}{\ell^2} +\frac{\rho_s^2}{\ell^2} \frac{\zeta^2 }{\mathcal{G}^2} \right) \chi'' -2 \,\frac{\bar{d}_e^2}{\ell^2} \frac{\chi'}{\zeta} - \left(1+ \frac{\zeta^2 }{\mathcal{G}^2}\right) \chi ={-}\frac{\zeta^2}{\mathcal{G}^2}\, \chi_{\infty}. \end{equation}

In the following, we will take different limits of this equation, depending on the tearing regime and wavelength limit considered. We emphasise that to give to the right-hand side term $\chi _\infty$ of (5.19) the geometrical (and physical) meaning we have previously discussed in terms of the boundary layer theory, the length $\ell$ must correspond to the layer width, $\delta _2$ or $\delta _1$ (depending on whether we are in the case of figure 4), to which the solution $\psi _{\text {out}}$ obtained in the ideal region is matched.

5.5 Wavelength limit and choice of integration via the equations for $\chi$ or for $\psi$ and $\varphi$

The choice of performing the integration by using (5.4)(5.5) for $\psi _1$ and $\varphi _1$, rather than (5.19) for $\chi$, clearly depends on analytical convenience. In general, we note that this choice can be biased by the ease by which the boundary conditions imposed by the solution obtained in the outer, ideal region, are ‘transferred’ to the solution in the innermost region. The asymptotic matching generally poses some constraints on the integration constants in each subdomain, but it is easy to see that, operationally speaking, these conditions take different and quite ‘appealing’ forms in the small- and large-$\varDelta '$ limits, depending on whether one looks at the equation for $\psi _1$ or at the auxiliary equation for $\chi$, respectively. In particular, the following are observed.

  1. (i) In the small- $\varDelta '$ limit, it is usually more convenient to perform the integration on the equation for $\psi _1$, for which the boundary conditions imposed from the ideal region are directly implemented by making explicit the dependence on $\varDelta '$ via (3.7) and (3.10). In this case, the constraint imposed by the solution $\psi _{\text {out}}$ in the ideal region can be directly used on the integration of the innermost equation in the form of the constant-$\psi$ condition. Indeed, if we define $c_0\equiv \lim _{x\to 0}\psi _{\text {out}}(x)$ and we heuristically assume the constant-$\psi$ condition to be valid, the latter implies that $\psi _1\simeq c_0$ both in the innermost region $|x|\leq \delta _1$ and in the possible intermediate region $\delta _1 < |x|\leq \delta _2$. Therefore, also when two boundary layers are present, the constant-$\psi$ condition can be directly implemented in the integration of the innermost equation, by thus practically allowing one to ‘bypass’ the procedure of matching between the solutions in the ideal and in the intermediate region, and then between the solutions in the intermediate and in the innermost regionFootnote 2 (at least for the purpose of establishing the dispersion relation, i.e. to find the asymptotic scalings of the eigenvalue). However, as the matching occurs in the boundary layer with the ideal region, it is generally more convenient to maintain the normalisation to the intermediate layer width, also when two layers are present. This approach will be used in §§ 6.2 and 7.5.

  2. (ii) In the large- $\varDelta '$ limit, it is instead usually more convenient to perform the integration on the auxiliary equation, since in this case, the constraint imposed by the outer solution $\psi _{\text {out}}$ on the solutions in the non-ideal region is simply expressed as $\chi _\infty = 0$ (cf. § 5.3.2). From an operational point of view, this realises the matching between the ideal and the intermediate region. The dispersion relation is therefore directly obtained by solving the non-ideal equation for $\chi _\infty = 0$ in case a single boundary layer is present, or by matching the innermost solution and the intermediate solution (both sought under the constraint $\chi _\infty = 0$) if two boundary layers are present. This approach will be used in §§ 6.1 and 7.4.

6 Solutions in the inner region: cold regime

Let us first discuss the reconnection regime with both cold electrons and ions. The linear problem has been first solved in the small-$\varDelta '$ wavelength limit by Furth et al. (Reference Furth, Killeen and Rosenbluth1963) in the resistive regime and by Coppi (Reference Coppi1964c,Reference Coppia) in the inertial regime, using boundary layer calculations in the coordinate space. A solution of the linear problem in the large-$\varDelta '$ wavelength limit has been first provided in the resistive regime by Coppi et al. (Reference Coppi, Galvao, Pellat, Rosenbluth and Rutherford1976) and has been extended in the collisionless regime by Porcelli (Reference Porcelli1991). In this latter work, the calculations in the Fourier space, with which Pegoraro & Schep (Reference Pegoraro and Schep1986) also recovered the resistive results of Furth et al. (Reference Furth, Killeen and Rosenbluth1963) and Coppi et al. (Reference Coppi, Galvao, Pellat, Rosenbluth and Rutherford1976), have been extended to the purely collisionless regime. We will come back to these more recent works relying on the integration in the Fourier space in § 7, where we will discuss the ‘warm’ reconnection regime including FLR effects related to parallel electron compressibility (i.e. a kind of diamagnetic effect).

The reason for which we start from the case of cold species is that, from the point of view of the boundary layer theory, the ‘cold’ regime in which we neglect $\rho _s$-related terms in (2.1)(2.2) is both conceptually and analytically simpler than the ‘warm’ tearing regime (cf. figure 4). Since we consider the cold limit $\bar {d}_e^2\gg \rho _s^2$ which encompasses the case $\rho _s=0$, to ensure convergence of the solution with the $\rho _s=0$ limit, we assume the validity of the condition (5.7) in the whole integration domain. We can verify a posteriori its validity, once the eigenvalue problem is solved.

We also mention here that the analytical results we are going to obtain below have been numerically verified by Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) to be valid for $\bar {d_e}^2> \rho _s^2/100$, i.e. for $\bar {d_e}\gtrsim \rho _s/10$.

6.1 Solution for $\bar {d}_e^2 / \rho _s^2 \gg 1$: large-$\varDelta '$ limit

Assuming heuristically the validity of (5.7), we neglect the second term in parentheses in (5.9). The restriction to the large $\varDelta '$-limit, in which the condition in (5.17) holds, suggests that we make use of the auxiliary function $\chi (\zeta )$ defined by (5.11), which leads us to consider the appropriate limit of (5.19):

(6.1)\begin{equation} \chi''-\frac{2}{\zeta}\chi'- \frac{\ell^2}{\bar{d}_e^2}\left(1+\frac{\zeta^2}{\mathcal{G}^2}\right)\chi = 0. \end{equation}

The structure of the equation suggests that we take $\ell \equiv \bar {d}_e$ as the normalisation length to be used in the definitions of (5.8). This means that we postulate

(6.2)\begin{equation} \delta_1=\bar{d}_e, \end{equation}

which, at least for $\bar {d_e}=d_e$, is evidently consistent with the asymptotic condition $\delta _1\ll 1$. A solution to (6.1) can be sought in the form

(6.3)\begin{equation} \chi_1 = e^{-\alpha \zeta^2}. \end{equation}

Substituting it in (6.1) yields

(6.4)\begin{equation} 2\alpha = 1, \quad 4 \alpha^2 = \frac{1}{\mathcal{G}^2}. \end{equation}

Using the first part of (5.8) and combining the two conditions above, gives

(6.5)\begin{equation} \gamma = k {\bar{d}_e} J_0. \end{equation}

This result can be specialised both to the fully collisionless ($\bar {d}_e^2=d_e^2$) and to the fully collisional ($\bar {d}_e^2=S^{-1}/\gamma$) internal-kink regime. We recall in this regard that numerical analysis (Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) confirms that the parameter space interval where both inertial and resistive effects non-negligibly combine is very narrow, so that tearing modes can be usually considered either as fully collisionless or as fully resistive.

In the collisionless limit, we thus recover the result of (Porcelli Reference Porcelli1991)

(6.6)\begin{equation} \gamma = k d_e J_0\,\quad \delta_1\sim d_e. \end{equation}

In the resistive case, (6.6) specialises to the result of (Coppi et al. Reference Coppi, Galvao, Pellat, Rosenbluth and Rutherford1976)

(6.7)\begin{equation} \gamma = k^{{2}/{3}} J_0^{{2}/{3}} S^{-{1}/{3}}, \quad \delta_1 \sim k^{-{1}/{3}} J_0^{-{1}/{3}} S^{- {1}/{3}}. \end{equation}

Substitution of the results of (6.6) and (6.7) for $x\sim \delta _1$ in (5.7) makes it equivalent to the condition $\rho _s^2/\bar {d}_e^2\ll 1$ from which we started, thus verifying a posteriori its correctness.

The scalings in (6.6) and (6.7)), in contrast to those of the large-$\varDelta '$ limit of the warm-collisionless regime, can be obtained using a heuristic-type derivation based on dimensional estimates, as we are going to prove in § 9.3.

Finally, some words about the second solution of (6.1), which we have neglected, since it can be shown that it is not of interest to the tearing instability problem. To this purpose, we can look for a solution of the form $\chi _2(\zeta ) = u(\zeta )\, \chi _1(\zeta )$. Substituting $\chi =\chi _2$ in (6.1) and using the fact that $\chi _1$ already solves it, one obtains, after a little algebra,

(6.8)\begin{equation} u''\chi_1 + 2 u'\left(\chi_1' -\frac{\chi_1}{\zeta}\right)+u\chi_1''=0. \end{equation}

Further substitution of (6.3) into the equation above yields

(6.9)\begin{equation} u''-\left(4\alpha \zeta + \frac{2 }{\zeta}\right)u'=0, \end{equation}

which, defining $w = u'$, can be brought to a directly integrable form:

(6.10)\begin{equation} \frac{{\rm d}w}{w} = 4 \alpha \zeta \,{\rm d}\zeta + \frac{2 \,{\rm d}\zeta}{\zeta} \implies w(\zeta) = w_0\, \zeta^2 e^{2\alpha\zeta^2}, \end{equation}

$w_0$ being a constant of integration. Specialising the result for $\alpha =1/2$, integrating $w$ once to find $u$ and then substituting the result in $\chi _2 = u \chi _1$, one obtains

(6.11)\begin{equation} \chi_2 = {\rm i} \sqrt{\frac{\rm \pi}{2}} {\rm erf}\left( \frac{{\rm i} \zeta}{\sqrt{2}} \right) e^{-({\zeta^2}/{2})} + \zeta, \end{equation}

which is odd and unbounded when $\zeta \to \infty$. Therefore, for our purposes, the solution $\chi _2$ can be disregarded since its spatial parity is not compatible with that of tearing modes (i.e. it is not compatible with the formation of $X$-points).

6.2 Solution for $d_e^2/ \rho _s^2 \gg 1$: small-$\varDelta '$ limit

In the small-$\varDelta '$ limit, the boundary layer analysis mirrors that of the paper by Furth et al. (Reference Furth, Killeen and Rosenbluth1963) if the resistive limit $\bar {d}_e^2\to S^{-1}/\gamma$ is considered. In general, regardless of the form of $\bar {d}_e$, we can here use the ‘constant-$\psi$’ approximation stating that $\tilde {\psi }_1(\zeta )\to c_0$ as $\zeta \to 0$ (cf. (5.15)). However, instead of using the auxiliary equation (5.19), in this case, it is more convenient to consider (5.9) in which $\tilde {\psi }_1''$ has been eliminated by using (5.10) since, after combination with the constant-$\psi$ limit, this yields an equation for the function $\tilde {\varphi }_1(\zeta )$ that can be brought to an integrable form more easily than the equation for $\chi (\zeta )$:

(6.12)\begin{equation} \tilde{\varphi}_1'' - \frac{\zeta^2}{\mathcal{G}^2} \frac{\ell^2}{\bar{d}_e^2}\tilde{\varphi}_1 = c_0\frac{\ell^2}{\bar{d}_e^2}\, \zeta. \end{equation}

Note that in this equation, the length $\ell$ is so far left unspecified. In practice, at this stage of the calculations, it can be taken $\ell =1$, which would be still consistent with the constant-$\psi$ limit with respect to which the substitution $\tilde {\psi }_1(0)\simeq c_0$ has been done. The identification of a proper normalisation length defining the boundary layer width can be done after a further change of variable that leads (6.12) to an integrable form in which the boundary layer integration can be more easily performed. Equation (6.12) for $\ell =1$ can be indeed brought to the form

(6.13)\begin{equation} \varPhi'' - z^2 \varPhi = z \end{equation}

after the change of variables and coordinates

(6.14)\begin{equation} z=\left({\mathcal{G}}{\bar{d}_e}\right)^{-{1}/{2}}\zeta,\quad {\varPhi} = \left(\frac{\bar{d}_e}{\mathcal{G}^3}\right)^{{1}/{2}} \left( \frac{\tilde{\varphi}_1}{c_0}\right). \end{equation}

The normalisation of Eq. (6.12) is detailed in Appendix D as a didactical example of how to generally proceed to find the normalisation coefficients of a differential equation. In the limit $\bar {d}_e^2\to S^{-1}/\gamma$, (6.13) is the equation of the resistive tearing mode studied by Furth et al. (Reference Furth, Killeen and Rosenbluth1963), whose solution has been provided by Basu & Coppi (Reference Basu and Coppi1976), Coppi et al. (Reference Coppi, Galvao, Pellat, Rosenbluth and Rutherford1976) and Ara et al. (Reference Ara, Basu, Coppi, Laval, Rosenbluth and Waddell1978) (cf. (III.26) and (III.27) and appendix A in the latter) in the closed integral form:

(6.15)\begin{equation} \varPhi={-} \frac{1}{2}z\int_0^1 (1-t^2)^{-{1}/{4}}e^{-({1}/{2})tz^2}\,{\rm d}t. \end{equation}

The scaling of the eigenvalue $\gamma$ and its dependence on the amplitude of the instability parameter $\varDelta '$ can be made explicit by comparing the eigenfunction $\varPhi (z)$ to ${\varphi }_1(x/\delta _2)$ in the matching condition expressed by (3.7), which also allows us to identify the layer width $\delta _2$: combining (3.7) and (3.10) (where we re-introduce the quantity $J_0$) yields

(6.16)\begin{equation} \psi_{\text{out}}(0) \varDelta' ={-}{\rm i}\frac{\gamma}{kJ_0}\int_{-\bar{X}'}^{+\bar{X}'} \frac{{\varphi}_1''(x')}{x'}\, {{\rm d} x}', \end{equation}

where $\varDelta '$ is defined by (3.8), $x'=x/\delta _1$ and $\bar {X}'=\bar {X}/\delta _1$ is indicative of some point of the matching region such that $\delta _1 \ll \bar {X}\ll 1$ in units of $L_0=a$. Connection with (6.13) follows from noticing that (6.14) implies:

(6.17)\begin{equation} \left.\frac{{\rm d}^2\tilde{\varphi}_1}{{\rm d}\zeta^2}\right|_{\ell=1} = c_0 \left.\left(\frac{1}{\bar{d}_e\mathcal{G}^3}\right)\right|_{\ell=1}^{{1}/{2}} \left( \frac{{\rm d}z}{{\rm d}\zeta} \right)^2 \frac{{\rm d}^2 \varPhi}{{\rm d}z^2} = c_0 \left.\left(\frac{\mathcal{G}}{\bar{d}_e^3}\right)\right|_{\ell=1}^{{1}/{2}} \frac{{\rm d}^2 \varPhi}{{\rm d}z^2}. \end{equation}

This authorises us to identify

(6.18)\begin{equation} x'=\frac{x}{\delta_1} \longrightarrow z=\frac{x}{(\mathcal{G}\bar{d}_e)|_{\ell=1}^{{1}/{2}}}, \quad -{\rm i}\frac{\gamma}{kJ_0}\frac{{\rm d}^2{\varphi(x')}_1}{{{\rm d} x}'^2} \longrightarrow c_0 \left.\left(\frac{\mathcal{G}}{\bar{d}_e^3}\right)\right|_{\ell=1}^{{1}/{2}} \frac{{\rm d}^2 \varPhi}{{\rm d}z^2}, \end{equation}

which now fixes the boundary layer width to be

(6.19)\begin{equation} {\delta_1}=\left.(\mathcal{G}\bar{d}_e)\right|_{\ell=1}^{{1}/{2}} =\sqrt{\frac{\gamma\bar{d}_e}{k}}, \end{equation}

whose consistency with the condition $\delta _1\ll 1$ must be verified once $\mathcal {G}|_{\ell =1}$ is explicitly obtained. Substituting then $\psi _{\text {out}}(0)\simeq \lim _{\zeta \to 0}\tilde {\psi }_1(\zeta )=c_0$ and using (6.18) makes (6.16) become

(6.20)\begin{equation} \varDelta' = \left.\left(\frac{\mathcal{G}}{\bar{d}_e^3}\right)\right|_{\ell=1}^{{1}/{2}} \int_{-\bar{Z}}^{+\bar{Z}} \frac{\varPhi''}{z} \,{\rm d}z, \end{equation}

where $\bar {Z} = \bar {X}/ (\mathcal {G}\bar {d}_e)|_{\ell =1}^{1/2}$. The integral in the previous equation is convergent, let us call its value $I\equiv \int _{-\bar {\infty }}^{+\infty } (\varPhi ''/{z})\, {\rm d}z \simeq \int _{-\bar {Z}}^{\bar {Z}} (\varPhi ''/{z}) \,{\rm d}z$. Then, using the definition in (5.8) for $\ell =1$, one obtains

(6.21)\begin{equation} \gamma = k{\bar{d}_e}^3 (\varDelta')^2 \left(\frac{J_0}{I^2}\right). \end{equation}

In the inertia-dominated regime, $\bar {d}_e=d_e$, one recovers the well-known result by Coppi (Reference Coppi1964c,Reference Coppia), later re-obtained by Porcelli (Reference Porcelli1991):

(6.22)\begin{equation} \gamma = k (\varDelta')^2 {{d}_e}^3 \left(\frac{J_0}{I^2}\right), \quad \delta_1 = \frac{\varDelta' d_e^2}{I}. \end{equation}

In taking the resistive limit, $\bar {d}_e= (S^{ -1}/\gamma )^{1/2}$, one obtains instead the classical result of Furth et al. (Reference Furth, Killeen and Rosenbluth1963), later recovered by Pegoraro & Schep (Reference Pegoraro and Schep1986):

(6.23)\begin{equation} \gamma = k^{{2}/{5}}(\varDelta')^{{4}/{5}} S^{-{3}/{5}} \left(\frac{J_0}{I^2}\right)^{{2}/{5}}, \quad \delta_1 = \frac{k^{-{2}/{5}}(\varDelta')^{{1}/{5}} S^{-{2}/{5}}}{({J_0}{I^2})^{{2}/{5}}}. \end{equation}

In both cases (6.22) and (6.23), the asymptotic condition $\delta _1\to 0$ as $d_e\to 0$ or $S^{-1}\to 0$ is verified.

7 Solutions in the inner region: warm-electron regime

Let us now consider the boundary layer problem in the ‘warm’ reconnecting regime by limiting our attention to the case of cold ions, consistent with (2.1) and (2.2).

Extending the analysis to include ion temperature effects related to FLR corrections in the fluid description of these low frequency modes is possible, but generally it requires relying on reduced models (i.e. ‘gyrofluid’ models), in which the ion response related to the separation between the motion of particles and of gyrocentres is obtained by starting from the gyrokinetic Vlasov equation. Fluid models of reconnection, in which this has been done, typically differ because of the way the gyrokinetic operator is approximated and because of the number of scalar fields that are retained in the description (see, e.g. Pegoraro & Schep Reference Pegoraro and Schep1981, Reference Pegoraro and Schep1986; Pegoraro et al. Reference Pegoraro, Porcelli and Schep1989; Porcelli Reference Porcelli1991; Schep et al. Reference Schep, Pegoraro and Kuvshinov1994; Loureiro & Hammett Reference Loureiro and Hammett2008 for essentially two-field models, and, e.g. Aydemir (Reference Aydemir1992), Zakharov & Rogers (Reference Zakharov and Rogers1992), Dorland & Hammett (Reference Dorland and Hammett1993), Smolyakov, Pogutse & Hirose (Reference Smolyakov, Pogutse and Hirose1995), Pogutse, Smolyakov & Hirose (Reference Pogutse, Smolyakov and Hirose1998), Waelbroeck, Hazeltine & Morrison (Reference Waelbroeck, Hazeltine and Morrison2009), Zocco & Schekochihin (Reference Zocco and Schekochihin2011), Waelbroeck & Tassi (Reference Waelbroeck and Tassi2012), Connor et al. (Reference Connor, Hastie and Zocco2012b) and Tassi, Sulem & Passot (Reference Tassi, Sulem and Passot2016) for other gyrofluid models retaining more than two scalar fields). A common feature of linear tearing analysis build on these gyrofluid models, as well as on other descriptions retaining kinetic features at a more fundamental level (see Drake & Lee Reference Drake and Lee1977; Cowley et al. Reference Cowley, Kulsrud and Hahm1986; Rogers et al. Reference Rogers, Kobayashi, Ricci, Dorland, Drake and Tatsuno2007), is the ‘symmetric’ way by which ion and electron temperature effects enter in the dispersion relation in regimes where an estimate of the growth rate of tearing-type modes is analytically available. This symmetry is expressible via the formal substitution $\rho _s^2 \to \rho _s^2+\rho _i^2$ in the dispersion relations which we are going to discuss in the remainder of this work. In regimes of the parameter space that go beyond the applicability of the analytical predictions by Pegoraro & Schep (Reference Pegoraro and Schep1986), Pegoraro et al. (Reference Pegoraro, Porcelli and Schep1989) and Porcelli (Reference Porcelli1991), a departure from the symmetry between $\rho _s$ and ion-Larmor-radius effects associated with the parameter $\rho _i=\rho _s \sqrt {T_i/T_e}$ has been numerically observed in the growth rate of the internal-kink mode (Del Sarto et al. Reference Del Sarto, Marchetto, Pegoraro and Califano2011). Dealing with these subjects would include some complicacy about the build of the linear model, which are unnecessary with respect to the purpose of the present work. This is why we neglect them here.

We therefore start from (5.4) and (5.5). Since the problem has now two characteristic spatial scales, namely $\bar {d}_e$ and $\rho _s$, the identification of the two boundary layers will require us to perform two subsequent normalisations and approximations of the terms of the aforementioned equations. Since we are now restricting to the parameter range where $\rho _s^2$ is not negligible with respect to $\bar {d}_e^2$, we can assume $\rho _s^2 \gtrsim \bar {d}_e^2$. A natural choice for choosing a first normalisation scale defining the ‘largest’ non-ideal region of interest is therefore to pose $\ell = \rho _s$. We note that identifying this as a boundary layer width or not is a further logical step, whose appropriateness depends on the kind of approximations of the equations and integration strategies that can be possibly performed with respect to the stretched variable $x/\rho _s$. This is what we are going to discuss below.

Also note that for economy of symbols – let us say – in this section, we will use some symbols used in the changes of variables of § 6 with a different definition that will be given in each specific case (this will concern, in particular, the definition of $\zeta$, $\mathcal {G}$ and the use of the ‘$\widetilde {\cdots }$’ to label some normalised quantities).

7.1 Equations in the non-ideal region for $\rho _s\gtrsim d_e$

As we enter the non-ideal region from the ideal one, when $x< 1$, the first non-ideal characteristic scale that is encountered under the hypotheses previously done is $\rho _s > d_e$. This is the first scale length with respect to which we can try to identify a boundary layer: aiming to single out a subdomain in the $x\ll 1$ range where only $\rho _s$ possibly matters, we start by fixing $\ell =\rho _s$ in (5.8), so that (5.9) and (5.10) become

(7.1)\begin{equation} \tilde{\psi}_1 - \left(\frac{\bar{d}_e^2}{\rho_s^2} + \frac{\tilde{\zeta}^2}{\tilde{\mathcal{G}^2}}\right)\tilde{\psi}_1'' ={-}\frac{\tilde{\zeta}}{\tilde{\mathcal{G}^2}}\tilde{\varphi}_1,\quad \tilde{\varphi}_1'' = \tilde{\zeta} \tilde{\psi}_1'', \end{equation}

with

(7.2)\begin{equation} \tilde{\zeta}=\frac{x}{\rho_s}, \quad\tilde{\mathcal{G}}=\frac{\gamma}{k\rho_s J_0}, \quad \tilde{\varphi}_1={-}i\tilde{\mathcal{G}}{\varphi}_1(\tilde{\zeta}), \quad \tilde{\psi}_1={\psi}_1(\tilde{\zeta}). \end{equation}

The auxiliary equation (5.19) specialises to

(7.3)\begin{equation} \left(\frac{\bar{d}_e^2}{\rho_s^2} + \frac{\tilde{\zeta^2} }{\tilde{\mathcal{G}}^2} \right) \tilde{\chi}'' -2 \frac{\bar{d}_e^2}{\rho_s^2} \frac{\chi'}{\tilde{\zeta}} - \left(1+ \frac{\tilde{\zeta}^2 }{\tilde{\mathcal{G}}^2}\right) \tilde{\chi} ={-}\frac{\tilde{\zeta}^2}{\tilde{\mathcal{G}}^2}\, \tilde{\chi}_{\infty}. \end{equation}

Equation (7.3) can be shown to be equivalent to (B42) of Zocco & Schekochihin (Reference Zocco and Schekochihin2011), although the latter has been obtained from a four-field system which also accounts for the dynamics of the reduced electron guiding centre contribution and for the parallel electron temperature perturbation ($g_e$ and $\delta T_{\|,e}$, respectively, therein). These authors also solved the boundary layer problem in the coordinate space, but – also because of the different system of equations they started from – they took a different path with respect to the one we take here, in which we tackle the integration of the inner equation directly in the different wave length limits of interest. In this sense, the comparison between the boundary layer integration procedure of the aforementioned article and the one we are going to outline here is not immediate, although possible, since some different operational assumptions in the definition of the boundary layers and in the analytical technique of integration have been used in that work and in the present one. In their work, Zocco & Schekochihin (Reference Zocco and Schekochihin2011) identified two non-ideal integration regions where ion physics and electron physics were respectively dominating. In these two regions, they sought the solutions of their auxiliary equation, from which they computed the magnetic flux function $\psi$ in the electron and ion region ($A_{\|,e}$ and $A_{\|,i}$ in their notation, respectively, which, in this work, map into the solution $\psi$ that we will compute in the innermost and outermost non-ideal regions) by assuming in the large- and small-wavelength limit some heuristic orderings (a posteriori verified) in terms of $\varDelta '$ and what they named the ‘reconnection region’ ($\delta _{\text {in}}$, in their notation). As we will later discuss (§ 8), the characteristic width of their ‘reconnection region’, $\delta _{\text {in}}$, differs from that of the innermost boundary layer that we are going to identify ($\delta _1$). However, in some regimes, $\delta_{in}$ can be interpreted in the light of a further characteristic scale length, which we will introduce in § 10 (namely, what we will call $(\varDelta _{v_y}')^{-1}$). In the following, we are going to adopt an integration procedure based on the integral representation of hypergeometric functions, whereas, in their work, (Zocco & Schekochihin Reference Zocco and Schekochihin2011) used a perturbative method to obtain a closed form solution for $A_{\|,e}$. Note that, so far, the boundary layers have not yet been identified here.

7.2 Boundary layers in the non-ideal region for $\rho _s^2/d_e^2\gg 1$

We now specialise the condition $\rho _s\gtrsim d_e$ by making the assumption $\rho _s^2/\bar {d}_e^2\gg 1$, which requires a sufficiently large scale separation between $\bar {d}_e$ and $\rho _s\geq \bar {d}_e$. Note that this is henceforth the strongest assumption about the relative ordering of the non-ideal parameters at play, which is done in this analysis for tearing modes in the warm-electron regime. Nevertheless, the quantitative results obtained in this way are usually applied also for $\rho _s\gtrsim \bar {d}_e$: numerical calculations in a wide interval of the parameter space (Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) suggest that the condition $\rho _s> \bar {d}_e/10$ can be assumed for the validity of the results we are going to obtain below. In the interval $\rho _s/10\lesssim \bar {d}_e \lesssim 10 \rho _s$, a transition from power-law scalings of the cold limit of § 6 to those of the warm limit that we are going to discuss here is observed.

7.2.1 Outermost boundary layer

Assuming $\rho _s^2/\bar {d}_e^2\gg 1$, (7.1) becomes

(7.4)\begin{equation} \tilde{\psi}_1 - \frac{\tilde{\zeta}^2}{\bar{\mathcal{G}^2}}\tilde{\psi}_1'' ={-}\frac{\tilde{\zeta}}{\bar{\mathcal{G}^2}}\tilde{\varphi}_1,\quad \tilde{\varphi}_1'' = \tilde{\zeta} \tilde{\psi}_1'' \end{equation}

and (7.3) reads

(7.5)\begin{equation} \tilde{\chi}'' - \left(1+\frac{\tilde{\mathcal{G}}^2 }{\tilde{\zeta}^2}\right) \tilde{\chi} ={-} \tilde{\chi}_\infty. \end{equation}

These equations must be assumed to be valid in an interval where spatial scales are comparable to $\rho _s$ but larger than $\bar {d}_e$, that is, $\bar {d}_e/\rho _s\lesssim \tilde {\zeta }\ll 1$, where the first inequality is meant in the sense discussed above. Therefore, restriction to this interval formally corresponds to take $\bar {d}_e = 0$ in (5.4). The integration of (7.8) and the matching of their solutions with $\psi _{\text {out}}$ of the ideal region via (3.5) and (3.6) formally identifies the outermost boundary layer of width

(7.6)\begin{equation} \delta_2 =\rho_s. \end{equation}

7.2.2 Innermost boundary layer

Of course, since the formation of the $X$-point via reconnection requires $\bar {d}_e\neq 0$, a second, innermost boundary layer can be defined in a region where both scales $\bar {d}_e$ and $\rho _s$ matter. The width $\delta _1$ of this second boundary layer does not necessarily correspond to the scale $\bar {d}_e$ (if it were so, the condition $\delta _1\ll \delta _2$ would require $\bar {d}_e\ll \rho _s$, which is an even stronger restriction than that defining the validity of (7.6)). To identify this inner boundary layer, we can look back to (7.1) and (7.3) or to ((5.9) and (5.10) and (5.19)) and notice that the two terms in parentheses, in which there is an explicit dependence on both $\rho _s$ and $\bar {d}_e$, are comparable (regardless of the definition of $\ell$ and $\mathcal {G}$) if $(\zeta \rho _s)^2/(\mathcal {G}\bar {d}_e)^2\sim 1$. In dimensional units, this means $x k\rho _sJ_0/(\gamma \bar {d}_e)\sim 1$, which suggests that we introduce the normalisation scale $\ell = {\gamma \bar {d}_e}/(k\rho _s J_0).$ Using then in (5.9) and (5.10) the change of coordinates and variables expressed by

(7.7)\begin{equation} \zeta=\frac{x }{\ell}= \frac{x k\rho_sJ_0}{\gamma \bar{d}_e}, \quad{\mathcal{G}}=\frac{\rho_s}{\bar{d}_e }, \quad {\varphi}_1={-}{\rm i}{\mathcal{G}}{\varphi}_1(\zeta), \quad {\psi}_1={\psi}_1(\zeta), \end{equation}

we write

(7.8)\begin{equation} {\psi}_1 - \frac{\bar{d}_e^2}{\ell^2}\left(1 + {\zeta^2}\right){\psi}_1'' ={-}\frac{\zeta}{{\mathcal{G}^2}}\varphi_1,\quad {\varphi}_1'' = \zeta {\psi}_1''. \end{equation}

Using the definition of ${\mathcal {G}}$ in (7.7) and the fact that $\rho _s^2/\bar {d}_e^2\gg 1$, one sees that in this interval, $\zeta {\varphi }_1 /{\mathcal {G}}^2\sim \zeta ^2 {\psi }_1 \bar {d}_e^2/(J_0^2)\rho _s^2 \ll {\psi }_1$ so that (7.8) can be approximated as

(7.9)\begin{equation} {\psi}_1 - \frac{\bar{d}_e^2}{\ell^2} \left(1 + \zeta^2\right) {\psi}_1'' =0,\quad {\varphi}_1'' = \zeta {\psi}_1'', \end{equation}

and with the same normalisation, (7.3) (or (5.19)) becomes

(7.10)\begin{equation} (1+\zeta^2)\chi''-\frac{2}{\zeta}\chi'-\frac{\ell^2}{\bar{d}_e^2}\chi={-}\zeta^2\chi_\infty. \end{equation}

Note that in the definitions above, we have used the same symbols for $\ell$, $\mathcal {G}$ and $\zeta$, previously introduced in (5.8), since the different regime of interest here should not induce any confusion with those symbols used in § 6. Also note that to avoid unnecessary burdens to notation in the following calculations, we have used for $\psi _1(\zeta )$ and $\varphi _1(\zeta )$ the same symbol that we use everywhere else for $\psi _1(x)$ and $\varphi _1(x)$. This, also, should not cause any confusion as the argument of the function will be later specified in the few cases where the two meanings could be confounded.

Later in this section, we will proceed with the integration of these solutions and with their matching. It is therefore reasonable to postulate the innermost layer width to be

(7.11)\begin{equation} \delta_1= \frac{\gamma \bar{d}_e}{k\rho_s J_0}, \end{equation}

which we will later show to be the case, indeed, although we remind that the identification $\delta _1\to \ell$ formally requires us to a posteriori verify that $\delta _1\ll \rho _s\ll 1$. From now on, we will therefore write $\delta _1$ defined by (7.11) in place of $\ell$ defined by (7.7).

7.3 Integration strategy in the warm-electron regime for $\rho _s^2/\bar {d}_e^2\gg 1$

We can now better establish the integration strategy for this problem, which has been generally outlined at the end of § 3. We have identified three overlapping spatial intervals of the integration domain in which different limits of the eigenvalue equations hold. For practical reasons, these can be identified in terms of the variable $\bar {\zeta }$ introduced in (7.1) and (7.2).

  1. (i) The innermost region $x\leq \delta _1$, where $\delta _1={\gamma \bar {d}_e}/(k\rho _s J_0)$, to be yet quantified, and (7.9)- and (7.10) are valid. This integration domain applies for $\tilde {\zeta }\ll 1$, i.e. for $x\ll \rho _s$.

  2. (ii) The intermediate region $\delta _1 \ll x\leq \delta _2$ where $\delta _2=\rho _s$ and (7.4) and (7.5) hold. This integration domain applies for $\tilde {\zeta }\sim 1$, i.e. for $\delta _1 \ll x \ll a$.

  3. (iii) The outer region valid for $x$ such that $\delta _2\ll x$, whose governing equations are those of ideal MHD. Here, (4.2) is used. This integration domain applies for $\tilde {\zeta }\gg 1$, i.e. for $x \sim a$.

Two matching regions can be therefore identified.

  1. (a) In the interval $\delta _1 < x< \delta _2$, the solution ${\chi }({\zeta })$ of (7.10) is matched with solution $\tilde {\chi }(\tilde {\zeta })$ of (7.5) according to $\lim _{{\zeta }\to \infty }{\chi }({\zeta })= \lim _{\tilde {\zeta }\to 0}\tilde {\chi }(\tilde {\zeta })$.

  2. (b) In the interval $\delta _2< x< a$, the solution $\tilde {\chi }(\tilde {\zeta })$ of (7.5) (corresponding to $\tilde {\psi }_1(\tilde {\zeta })$ and $\tilde {\varphi }_1(\tilde {\zeta })$ of (7.4)) is matched with the solutions $\psi _{\text {out}}(x)$ and $\varphi _{\text {out}}(x)$ of (4.1) and (4.2) according to $\lim _{\tilde {\zeta }\to \infty }\tilde {\psi }_1(\tilde {\zeta }) = \lim _{{x}\to 0}{\psi _{\text {out}}}(x)$ and $\lim _{\tilde {\zeta }\to \infty }\tilde {\varphi }_1(\tilde {\zeta }) = \lim _{{x}\to 0}{\varphi _{\text {out}}}(x)$.

Notice that the matching conditions also generally depend on the wavelength range, that is, on the value of $\varDelta '$, which rules the solution in the ideal region. This can influence the integration strategy, as outlined in § 5.5.

Before solving (7.1) or (7.3) in both the inner and the outer regions, and then matching the corresponding solutions to obtain the general solution of the problem, it is useful to discuss (see § 7.3.1) a general method that can be used to solve equations having the form of (7.9). This method is based on the integration in the complex plane via representation by means of hypergeometric functions.

7.3.1 Solution of the equation in the innermost interval ($x\ll \delta _1$) via integral representation of hypergeometric functions

Equation (7.9) is a second-order ordinary differential equation that in the complex plane has three singular points at $\zeta = \pm i$ and $\zeta = \infty$. Therefore, employing some transformations, it can be written in the standard form of a hypergeometric equation. This suggests that we look for the solution of (7.9) and (7.10) in terms of hypergeometric functions. Inspired by the integral representation of hypergeometric functions, we can thus look for a solution of (7.9) of the form

(7.12)\begin{equation} \psi_1 = \frac{1}{2{\rm \pi} {\rm i}} \oint_C F(s) (1 + \zeta^2)^s \,{\rm d}s, \end{equation}

where $C$ is a closed contour in the complex plane which starts at $-\infty$ of the real axis, goes around the singularities of the function $F(s)$ (it will turn out that these singularities are given by an infinite number of simple poles of $F(s)$) and then returns back to $-\infty$. This representation will allow us to employ the residue theorem which states that $\psi _1$ equals the sum of residues enclosed by the contour $C$. The second derivative of $\psi _1$ becomes

(7.13)\begin{equation} \psi_1'' = \oint_C 2s (2s - 1) F(s) (1 + \zeta^2)^{s - 1} \,{\rm d}s - \oint_C 4s(s - 1) F(s) (1 + \zeta^2)^{s - 2} \,{\rm d}s. \end{equation}

Shifting the operator in the second term by taking $s' = s + 1$, (7.13) reads

(7.14)\begin{equation} \psi_1'' = \oint_C 2s (2s - 1) F(s) (1 + \zeta^2)^{s - 1} \,{\rm d}s - \oint_C 4s(s + 1) F(s + 1) (1 + \zeta^2)^{s - 1} \,{\rm d}s. \end{equation}

Substituting the previous relation in (7.9), one obtains an expression stating that the integral of $(1+\zeta ^2)^s$ times a function of $s$ equals zero. A sufficient condition for its validity is to set to zero the function of $s$, which reads

(7.15)\begin{equation} [2s (2s - 1) - (\delta_1/\bar{d}_e)^2] F(s) - 4s (s + 1) F(s + 1) = 0. \end{equation}

To get an insight on the form of $F(s)$, let us write $F(s+k)$ in terms of $F(s)$ by using (7.15). After a little algebra, we obtain

(7.16)\begin{equation} F(s + 1) = \frac{\left(s - \dfrac{\alpha}{2}\right) \left(s - \dfrac{\beta}{2}\right)}{s (s + 1)} F(s), \end{equation}

where $\alpha =\frac {1}{2} - \nu$, $\beta = \frac {1}{2} + \nu$ and $\nu = ({1}/{4} + (\delta _1/\bar {d}_e)^2)^{{1}/{2}}$. Then,

(7.17)\begin{equation} F(s + j) = \frac{\left(s - \dfrac{\alpha}{2}\right)_j \left(\dfrac{s - \beta}{2}\right)_j}{(s)_j (s + 1)_j} F(s), \end{equation}

where

(7.18)\begin{equation} (q)_j \equiv q (q + 1) \cdots (q + j - 1) =\frac{\varGamma(q + j)}{\varGamma(q)} \end{equation}

is the so-called ‘Pochhammer symbol’ defined in terms of the $\varGamma$-function (see Abramowitz & Stegun Reference Abramowitz and Stegun1964, § 6.1.22). Motivated by the form of (7.17) one can look for a function $F(s)$ expressed as a combination of Gamma functions, that is,

(7.19)\begin{equation} F(s) = \frac{\varGamma(s + a) \varGamma(s + b)}{\varGamma(s + c) \varGamma(s + d)}. \end{equation}

If, by comparison with (7.17), we choose $c = 0$ and $d = 1$, (7.15) becomes

(7.20)\begin{equation} \left(a + b + \frac{1}{2}\right) s + ab + \left(\frac{\delta_1}{2 \bar{d}_e}\right)^2 = 0. \end{equation}

The previous equation should be satisfied for any value of $s$. Therefore, one obtains

(7.21)\begin{equation} a + b + \frac{1}{2} = 0, \quad ab + \left(\frac{\delta_1}{2 \bar{d}_e}\right)^2 = 0, \end{equation}

which has the following solutions:

(7.22)\begin{equation} a ={-}\frac{\alpha}{2}, \quad b ={-}\frac{\beta}{2}, \quad \alpha = \frac{1}{2} - \nu, \quad \beta = \frac{1}{2} + \nu, \quad \nu = \left( \frac{1}{4} + \frac{\delta_1^2}{\bar{d}_e^2} \right)^{{1}/{2}}. \end{equation}

Here, $a$ and $b$ are real numbers because $(\delta _1/\bar {d}_e)^2$ is also a real number, whence it is evident that the singularities of $F(s)$ are points on the real axis. Using (7.22), $F(s)$ reads

(7.23)\begin{equation} F(s) = \frac{\varGamma\left(s - \dfrac{\alpha}{2}\right) \varGamma\left(s - \dfrac{\beta}{2}\right)}{\varGamma(s) \varGamma(s + 1).} \end{equation}

One sees that $F(s)$ has two series of simple poles at

(7.24)\begin{equation} s = \frac{\alpha}{2} - m, \quad s = \frac{\beta}{2} - m, \quad m = 0, 1, 2, 3,\ldots. \end{equation}

It is also clear that the previous poles are associated with two series of residues. The first one is obtained by taking the limit $s = {\alpha }/{2} - m$ in the evaluation of the residue of $F(s) (1 + \zeta ^2)^s$,

(7.25)\begin{align} \psi_\alpha &= \text{Res}_{\alpha} \equiv \sum_{m = 0}^{\infty} \lim_{s \to {\alpha}/{2} - m} \left[ (s + m) \varGamma\left(s - \frac{\alpha}{2}\right) \right] \frac{\varGamma(-\nu - m) (1 + \zeta^2)^{{\alpha}/{2} - m}}{\varGamma\left(\dfrac{5}{4} - \dfrac{\nu}{2} - m\right) \varGamma\left(\dfrac{1}{4} - \dfrac{\nu}{2} - m\right)} \nonumber\\ &= \sum_{m = 0}^{\infty} \frac{({-}1)^{m}}{m!} \frac{\varGamma(-\nu - m)}{\varGamma\left(\dfrac{5}{4} - \dfrac{\nu}{2} - m\right) \varGamma\left(\dfrac{1}{4} - \dfrac{\nu}{2} - m\right)} (1 + \zeta^2)^{{\alpha}/{2} - m}, \end{align}

where the limit in the previous equation is obtained using the property $\varGamma (s + 1) = s\varGamma (s)$, which leads to

(7.26)\begin{align} &\lim_{s\to -m} (s+m)\varGamma(s) = \lim_{s\to -m} (s+m)\frac{\varGamma(s + 1)}{s} \nonumber\\ &\qquad = \lim_{s\to -m} (s+m)\frac{\varGamma(s + m + 1)}{s (s + 1) \cdots. (s + m - 1)(s+m)} = \frac{({-}1)^{m}}{m!}. \end{align}

The second series of the residues can be calculated taking the limit $s = {\beta }/{2} - m$. Then,

(7.27)\begin{equation} \psi_\beta = \text{Res}_\beta = \sum_{m = 0}^{\infty} \frac{({-}1)^{m}}{m!} \frac{\varGamma(\nu - m)}{\varGamma\left(\dfrac{5}{4} + \dfrac{\nu}{2} - m\right) \varGamma\left(\dfrac{1}{4} + \dfrac{\nu}{2} - m\right)} (1 + \zeta^2)^{{\beta}/{2} - m}. \end{equation}

Therefore, using the residue theorem ($\oint _C f(z) \,{\rm d}z = 2 {\rm \pi}i \sum _{z \in C} \text {Res}[ f(z)]$), and substituting in (7.12), $\psi _1$ becomes

(7.28)\begin{equation} \psi_1 = \psi_\alpha + \psi_\beta. \end{equation}

We recall that this solution is valid as long as $x\ll \delta _2$, i.e. $x \ll \rho _s$.

The same result could be found by using the Frobenius method (see, e.g. Bender & Orszag Reference Bender and Orszag1978, § 3.3) by assuming a power series solution of the form $\psi _1 = \sum _{m = 0}^{\infty } g_m (1 + \zeta ^2)^{\lambda - m}$: after substituting $\psi _1$ in (7.9) and using $g_0\neq 0$, the indicial equation leads to $\lambda _1 = {\alpha }/{2}$ and $\lambda _2 = {\beta }/{2}$, where $\alpha$ and $\beta$ are given by (7.22). The coefficients $g_m$ for both $\lambda _{1}$ and $\lambda _{2}$ can be found by equating the coefficients of the powers of $\zeta$ to zero. In this way, one obtains again (7.25)(7.28).

A discussion about both the convergence and the independence of solutions (7.25) and (7.27) is in Appendix E.

7.4 Solution of the auxiliary equation for $\rho _s^2/d_e^2\gg 1$: large-$\varDelta '$ limit

We now solve the auxiliary equation (7.10) in the large-$\varDelta '$ limit. Its detailed analytical investigation becomes essential in this wavelength limit since, as we will see in § 9.3, the ‘standard’ heuristic derivation here (surprisingly) fails to recover the correct scaling laws.

In this wavelength limit, the solution of the auxiliary equation in the innermost region can be matched directly to the solution obtained in the outermost non-ideal region by setting $\chi _\infty =0$, as discussed in § 5.5. From this matching, we will obtain the scaling laws for both the growth rate and the reconnecting layer width.

7.4.1 Solution in the innermost layer $x\leq \delta _1$

We can solve (7.10) by following the same line of thought used to solve (7.9), as described in § 7.3.1: we thus assume a solution of the form

(7.29)\begin{equation} \chi(\zeta) = \frac{1}{2{\rm \pi} i} \oint_C G(s) (1 + \zeta^2)^s \,{\rm d}s. \end{equation}

After calculating the first and second derivative of (7.29) with respect to $\zeta$, and substituting them into (7.10), one obtains the condition

(7.30)\begin{equation} [2s(2s - 1) - (\delta_1/\bar{d}_e)^2 ] G(s) - 4(s + 1)^2 G(s + 1) = 0. \end{equation}

By noticing the similarity of the previous relation with (7.15), we look for a $G(s)$ of the form

(7.31)\begin{equation} G(s) = \frac{\varGamma(s + a) \varGamma(s + b)}{\varGamma^2(s + 1)} = \frac{F(s)}{s}. \end{equation}

We recall that the previous equation has been written by assuming $c = d = 1$ in (7.19), and this allows us to eliminate the $(s+1)^2$ coefficient in (7.30). In (7.31), we have also used $\varGamma ^2(s+1) = s \varGamma (s)\varGamma (s+1)$. This choice leads us again to (7.21) and, therefore, to the same values for $a$ and $b$ which are given by (7.22), meaning that both (7.19) and (7.31) have the same infinite series of poles, as can be expected from the definition of $\chi$ expressed by (7.29). Therefore, the general solution of (7.10) reads

(7.32)\begin{equation} \chi_{\text{in}}(\zeta) = \chi_\alpha(\zeta) + \chi_\beta(\zeta), \end{equation}

where

(7.33)\begin{gather} \chi_\alpha(\zeta) = \sum_{m=0}^{\infty} \frac{({-}1)^m}{m!} \frac{\varGamma(-\nu - m)}{\varGamma^2\left(\dfrac{5}{4} - \dfrac{\nu}{2} - m\right)} (1 + \zeta^2)^{{\alpha}/{2} - m}, \end{gather}
(7.34)\begin{gather} \chi_\beta(\zeta) =\sum_{m = 0}^{\infty} \frac{({-}1)^{m}}{m!} \frac{\varGamma(\nu - m)}{\varGamma^2\left(\dfrac{5}{4} + \dfrac{\nu}{2} - m\right)} (1 + \zeta^2)^{{\beta}/{2} - m}. \end{gather}

We close the discussion in this section by noticing that the leading terms in (7.32) are $(1 + \zeta ^2)^{{\alpha }/{2}}$ and $(1 + \zeta ^2)^{{\beta }/{2}}$. Therefore, we can look for an approximate asymptotic solution of the form

(7.35)\begin{equation} \chi_{\text{in}}(\zeta) \approx \chi_{\text{in}}^{\text{appr}}(\zeta) = A (1 + \zeta^2)^{{\alpha}/{2}} + B (1 + \zeta^2)^{{\beta}/{2}} + C, \end{equation}

where $A$, $B$ and $C$ are constants. Their values will be estimated by matching the solutions in the different regions and by applying the constraints at $\zeta =0$, which we are going to obtain in § 7.4.2. Note that for limit $\zeta \to \infty$, (7.35) can be further approximated to

(7.36)\begin{equation} \chi_{\text{in}}(\zeta) \approx \chi_{\text{in}}^{\text{appr}}(\zeta) = A \zeta^\alpha + B \zeta^\beta + C. \end{equation}

7.4.2 Constraints at the origin

Any solution of the auxiliary equation must satisfy the constraints below that follow from the definition of $\chi (\zeta )$ in (5.11) and from (7.9) at $\zeta = 0$:

(7.37)\begin{equation} \psi_1''(0) = \frac{\delta_1^2}{\bar{d}_e^2} \psi_1(0) = \chi''(0) ={-} \frac{\delta_1^2}{\bar{d}_e^2} \chi(0), \end{equation}
(7.38)\begin{equation} \psi_1^{iv}(0) = \frac{\delta_1^2}{\bar{d}_e^2} \psi_1''(0) - 2 \frac{\delta_1^2}{\bar{d}_e^2} \psi_1(0) ={-} 2 \left(1 - \frac{1}{2}\frac{\delta_1^2}{\bar{d}_e^2}\right) \psi_1''(0) ={-}2\frac{\delta_1^2}{\bar{d}_e^2} \left(1 - \frac{1}{2}\frac{\delta_1^2}{\bar{d}_e^2} \right) \psi_1(0), \end{equation}
(7.39)\begin{equation} \chi^{iv}(0) = 3 \psi_1^{iv}(0) ={-}6\left(1 - \frac{1}{2}\frac{\delta_1^2}{\bar{d}_e^2}\right)\psi_1''(0) ={-}6\left(1 - \frac{1 }{2}\frac{\delta_1^2}{\bar{d}_e^2}\right) \chi''(0) = 6\frac{\delta_1^2}{\bar{d}_e^2} \left(1 - \frac{1}{2}\frac{\delta_1^2}{\bar{d}_e^2}\right) \chi(0). \end{equation}

These expressions will be later used in the asymptotic matching, so as to determine the free coefficients in the expression of $\chi ^{\text {aprr}}_{\text {in}}$ in (7.35).

7.4.3 Solution in the intermediate region, $\delta _1 \ll x \leq \delta _2$

In this region, we look for a solution $\tilde {\chi }(\tilde {\zeta })$ of (7.3). Being interested in the large-$\varDelta '$ limit, we can pose $\tilde {\chi }_\infty = 0$. Equation (7.3) then becomes the homogeneous equation

(7.40)\begin{equation} \tilde{\chi}'' - \left(1 + \frac{\tilde{\mathcal{G} }^2}{\tilde{\zeta}^2}\right) \tilde{\chi} = 0, \end{equation}

where ‘$'$’ refers to the derivative with respect to $\tilde {\zeta }=x/\rho _s$. This equation belongs to the family of modified Bessel equations (see Abramowitz & Stegun Reference Abramowitz and Stegun1964, § 9.6.1). In fact, it can be transformed into the standard modified Bessel equation via the change of variable $\chi \tilde {\zeta } = \tilde {\zeta }^{{1}/{2}}u(\tilde {\zeta })$ that leads to

(7.41)\begin{equation} \tilde{\zeta}^2 u'' + \tilde{\zeta} u' - \left(\tilde{\zeta}^2 + \nu^2\right) u = 0, \end{equation}

where $\nu$ is given by the last of (7.22). The only solution of the previous equation that is bounded when $\tilde {\zeta } \to \infty$ (or equivalently, when $x \to \infty$) is the modified Bessel function of the second kind, that is, $u(\tilde {\zeta }) = \mathcal {K}_\nu (\tilde {\zeta })$. Therefore, (7.40) has the solution

(7.42)\begin{equation} \tilde{\chi}_{\text{mid}}(\tilde{\zeta}) = \tilde{\zeta}^{{1}/{2}} \mathcal{K}_\nu(\tilde{\zeta}), \end{equation}

the label ‘$_{\text {mid}}$’ standing for the intermediate region domain $\delta _1 \ll x \leq \delta _2$ where it is defined. Its asymptotic behaviour as $\tilde {\zeta } \to 0$ is given by

(7.43)\begin{align} \tilde{\chi}^{\text{appr}}_{\text{mid}} & \equiv \lim_{\tilde{\zeta}\ll 1} \tilde{\chi}_{\text{out}}(\tilde{\zeta}) = \lim_{\tilde{\zeta} \ll 1} \tilde{\zeta}^{{1}/{2}} \mathcal{K}_\nu(\tilde{\zeta}) = \lim_{\tilde{\zeta} \ll 1} \tilde{\zeta}^{{1}/{2}} \left(\frac{\rm \pi}{2}\right) \frac{\mathcal{I}_{-\nu}(\tilde{\zeta}) - \mathcal{I}_{\nu}(\tilde{\zeta})}{\sin({\rm \pi} \nu)}\nonumber\\ &\sim \frac{2^\nu}{\varGamma(1 - \nu)} \tilde{\zeta}^\alpha - \frac{2^{-\nu}}{\varGamma(1 + \nu)} \tilde{\zeta}^\beta, \end{align}

where we used the relation in § 9.6.2 of Abramowitz & Stegun (Reference Abramowitz and Stegun1964). Here, $\alpha$ and $\beta$ are once more given by (7.22), and $\mathcal {I}_{\nu }(\tilde {\zeta })$ is the modified Bessel function of the first kind,

(7.44)\begin{equation} \mathcal{I}_{\nu}(\tilde{\zeta})\equiv \left(\frac{\tilde{\zeta}}{2}\right)^\nu \sum_{m = 0}^{\infty} \frac{\displaystyle{\left(\frac{\tilde{\zeta}^2}{4}\right)^m}}{m! \varGamma(\nu + m + 1)}. \end{equation}

7.4.4 Matching of solutions: scaling laws and eigenfunctions

The scaling laws for both the growth rate $\gamma$ and the width of the innermost layer, $\delta _1$, are obtained from matching the solutions in the innermost and intermediate regions of the domain. This formally closes the eigenvalue problem.

To accomplish this, we start by estimating the second and the fourth derivative of $\chi _{\text {in}}^{\text {appr}}(\zeta )$ at $\zeta = 0$. Then we will apply the constraints at the origin written in § 7.4.2 and the condition $\lim _{\zeta \to \infty }\chi (\zeta )=\lim _{\tilde {\zeta }\to 0}\tilde {\chi }(\tilde {\zeta })$, that is, $\lim _{\zeta \to \infty }\chi ^{\text {appr}}_{\text {in}}(\zeta )=\lim _{\tilde {\zeta }\to 0}\tilde {\chi }^{\text {appr}}_{\text {mid}}(\tilde {\zeta })$, so as to obtain the equations that will allow us to find $A$, $B$ and $C$, and, finally, the scaling laws.

First of all, from the matching of (7.43) with (7.36), we obtain

(7.45)\begin{equation} A = \frac{2^\nu}{\varGamma(1 - \nu)} \left(\frac{\delta_1}{\rho_s}\right)^\alpha, \quad B ={-}\frac{2^{-\nu}}{\varGamma(1 + \nu)} \left(\frac{\delta_1}{\rho_s}\right)^\beta. \end{equation}

Taking the second and the fourth derivatives of $\chi _{\text {in}}^{\text {appr}}$ (we here omit superscript ‘${\text {appr}}$’ to simplify the notation), one finds

(7.46)\begin{gather} \chi_{\text{in}}(0) = A + B + C, \end{gather}
(7.47)\begin{gather} \chi_{\text{in}}''(0) = A \alpha + B \beta, \end{gather}
(7.48)\begin{gather} \chi_{\text{in}}^{iv}(0) = 6 A \alpha \left(\frac{\alpha}{2} - 1\right) + 6 B \beta \left(\frac{\beta}{2} - 1\right). \end{gather}

Then, using the constraint on the second derivative given by (7.37) and (7.47),

(7.49)\begin{equation} C ={-} A \left(1 + {\alpha}\left(\frac{\bar{d}_e}{\delta_1}\right)^2\right) - B \left(1 + {\beta}\left(\frac{\bar{d}_e}{\delta_1}\right)^2\right). \end{equation}

The constraints on the fourth derivative in (7.39) and on (7.48) instead lead to

(7.50)\begin{equation} A \alpha \left(\alpha - \left(\frac{\delta_1}{\bar{d}_e}\right)^2\right) + B \beta \left(\beta - \left(\frac{\delta_1}{\bar{d}_e}\right)^2 \right) = 0. \end{equation}

Combining the previous equation with (7.45), one obtains

(7.51)\begin{equation} -\frac{A}{B} = \frac{\beta \left(\beta - \left(\delta_1/\bar{d}_e\right)^2\right)}{\alpha \left(\alpha - \left(\delta_1/\bar{d}_e\right)^2\right)} = 2^{2\nu} \left(\frac{\delta_1}{\rho_s}\right)^{{-}2\nu} \frac{\varGamma(1 + \nu)}{ {\varGamma(1 - \nu)}}, \end{equation}

where we used $\alpha - \beta = - 2\nu$. For small values of $(\delta _1/\bar {d}_e)$, (7.22) gives $\nu \approx 1/2$, $\alpha \approx - (\delta _1/\bar {d}_e)^2$ and $\beta \approx 1$; therefore, (7.51) becomes

(7.52)\begin{equation} \frac{1}{2}\left(\frac{\bar{d}_e}{\delta_1}\right)^4 \left(1- \left(\frac{\delta_1}{\bar{d}_e}\right)^2\right) = 2 \frac{\varGamma\left(\dfrac{3}{2}\right)}{\varGamma\left(\dfrac{1}{2}\right)} \left(\frac{\rho_s}{\delta_1}\right). \end{equation}

Using $\varGamma (3/ 2)=\varGamma (1/ 2)/ 2 = \sqrt {{\rm \pi} }/ 2$ and $(\delta _1/\bar {d}_e)^2\ll 1$ (to be verified a posteriori), the equation above implies

(7.53)\begin{equation} \delta_1 \sim \rho_s^{-{1}/{3}} \bar{d}_e^{{4}/{3}} \end{equation}

whence, using the definition in (7.11), one obtains

(7.54)\begin{equation} \gamma \sim k\rho_s^{{2}/{3}} \bar{d}_e^{{4}/{3}}. \end{equation}

Specialising (7.53) and (7.54) to the purely resistive regime, $\bar {d}_e=(S^{-1}/\gamma )^{1/2}$, one recovers the result of Pegoraro & Schep (Reference Pegoraro and Schep1986) for the growth rate (cf. (33) therein and also (20) in Pegoraro et al. Reference Pegoraro, Porcelli and Schep1989) and of Zocco & Schekochihin (Reference Zocco and Schekochihin2011) for $\delta _1$ (cf. $\delta _\eta$ in (B100), therein),

(7.55)\begin{equation} \gamma \sim k^{{6}/{7}} \rho_s^{{4}/{7}} S^{-{1}/{7}},\quad \delta_1 \sim k^{-{4}/{7}} \rho_s^{-{5/}{7}} S^{-{4}/{7}}. \end{equation}

Specialising the same equations to the purely collisionless regime, $\bar {d}_e=d_e$, one recovers instead the result of Porcelli (Reference Porcelli1991) for both for the growth rate ((8), therein) and for the innermost layer width (quantity $\sigma$, defined therein above (7); see also Bhattacharjee, Germaschewski & Ng Reference Bhattacharjee, Germaschewski and Ng2005),

(7.56)\begin{equation} \gamma \sim k \rho_s^{{2}/{3}} d_e^{{1}/{3}},\quad \delta_1 \sim \rho_s^{-{1}/{3}} d_e^{{4}/{3}}. \end{equation}

In both the purely resistive and the collisionless case, the conditions $\delta _1^2\ll \bar {d}_e^2$ and $\delta _1\ll \delta _2\sim \rho _s$ we had previously heuristically assumed, are a posteriori verified.

Finally, by looking at the definition of $\chi$, it is obvious that $\psi _1 \approx \chi _{\text {in}}^{\text {appr}}$ when $x \ll \delta _1$, since $x\psi '_1 \sim {x}\psi _1/{\delta _1} \ll 1$. Therefore, substituting the values of $\nu$, $\alpha$, $\beta$, $\mathcal {G}$ and $\delta _1$ in (7.35), one obtains for the inner layer

(7.57)$$\begin{gather} \lim_{x\ll \delta_1}\psi_{1} \simeq \frac{2^{1/2}}{\varGamma(1/2)} \left( \frac{\delta_1}{\rho_s} \right)^{- {\delta_1^2}/{\bar{d}_s^2}} \left( 1 + \frac{x^2}{\delta_1^2} \right)^{- {\delta_1^2}/{\bar{d}_s^2}} - \frac{2^{{-}1/2}}{\varGamma(3/2)} \displaystyle{\left( \frac{\delta_1}{\rho_s} \right) \left( 1 + \frac{x^2}{\delta_1^2} \right)^{{1}/{2}} } \nonumber\\ + \left( \frac{\delta_1}{\rho_s} \right) \left( 1 + \frac{\bar{d}_s^2}{\delta_1^2}\right). \end{gather}$$

Integrating (5.12) and using (7.35), one finds for $x \ll \delta _1$,

(7.58)$$\begin{gather} \lim_{x\ll \delta_1} \varphi_{1} \simeq \sum_{m = 0}^{\infty} \left[ \frac{2^{1 /2}}{\varGamma({1}/{2})} \left( \frac{\delta_1}{\rho_s} \right)^{- {\delta_1^2}/{\bar{d}_s^2} } \left( \begin{matrix} -{\delta_1^2}/{\bar{d}_s^2} \\ m \end{matrix} \right)_b - \frac{2^{{-}1 /2}}{\varGamma({3}/{2})} \left( \frac{\delta_1}{\rho_s} \right) \left( \begin{matrix} {-1}/{2} \\ m \end{matrix} \right)_b \right] \left( \frac{x}{\delta_1} \right)^{2\,m + 1} \nonumber\\ + \left( \frac{\delta_1}{\rho_s} \right) \left( 1 + \frac{\bar{d}_s^2}{\delta_1^2} \right) \left( \frac{x}{\delta_1} \right), \end{gather}$$

where

(7.59)\begin{equation} \left( \begin{matrix} r \\ m \end{matrix} \right)_b = \displaystyle{\frac{r (r - 1) \cdots (r - m + 1)}{m!}} = \displaystyle{\frac{\left( r \right)_m}{m!}} \end{equation}

is the usual binomial coefficient, generalised for real values of the argument $r$.

Notice that for ${x} \leq \delta _1$, the dominant term in (7.58) is the one with $m=0$. Therefore, to the lowest order, $\varphi$ can be approximated as $\varphi \sim x$. Figure 8 shows the spatial profile of $\psi _1$, $\chi _{\text {in}}^{\text {appr}}$ and $x\psi _1'$ in blue, black and orange colours, respectively, in a sub-interval of the integration domain. These profiles have been computed numerically by solving the eigenvalue problem for $\psi _1$ and $\varphi _1$ with the solver of Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) and using the definitions in (7.36). We can this way numerically verify that $\chi _{\text {in}}^{\text {aprr}} \approx \psi _1$ within the innermost layer with $|x|\ll \delta _1$, in agreement with the previous discussion.

Figure 8. Profiles of $\psi _1$ (blue), $\chi _{\text {in}}^{\text {appr}}$ of (7.46) (black-dashed) and $x \psi _1'$ (orange) in the reconnection layer around the neutral line: these profiles are obtained by numerical integration of the complete eigenvalue problem and should be compared to the approximated analytical solutions ((7.57) and (7.58)). The width of reconnection layer, evaluated as it will be specified in § 7, is $\delta = \delta _1 \simeq 10^{-6}$. The physical parameters of the numerical integration performed with the solver of Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) in the large-$\varDelta '$ limit are shown in the figure.

The eigenfunctions (7.57) and (7.58) agree with the behaviour of the perturbed current density profile provided by Pegoraro & Schep (Reference Pegoraro and Schep1986) in the inner layer for the warm-resistive regime ($\bar {d}_e=(S^{-1}/\gamma )^{1/2}$), which had been already numerically verified by Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020).

7.5 Solution of the linear equations for $\rho _s^2/d_e^2\gg 1$: small-$\varDelta '$ limit

In the small-$\varDelta '$ limit, we use again the constant-$\psi$ approximation, as we already did in the integration of the cold-electron limit in § 6.2. Figure 9 shows the appropriateness of this assumption also in the warm-electron regime: it displays the profile of $\psi _1$ numerically computed with the eigenvalue solver of Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) in the collisionless regime for $d_e=2\times 10^{-2}$, $\rho _s=5\times 10^{-3}$ and for a value of $k=1.7$, which corresponds to $\varDelta '(k)=3.3$ and $\delta _1\simeq 2.4\times 10^{-6}$ (evaluated according to the definition in (8.3) – see later).

Figure 9. Spatial profile of $\psi _1$, numerically computed with the eigenvalue solver of Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) in the collisionless regime for a choice of the parameters $d_e=2\times 10^{-2}$, $\rho _s=5\times 10^{-3}$, and for a value of $k=1.7$. These values correspond to $\varDelta '(k)=3.3$ and to $\delta =\delta _1\simeq 2.4\times 10^{-6}$, evaluated according to the definition in (8.3) introduced next. The smaller box inserted at the top right-hand corner of the figure shows a zoom of the solution inside the innermost layer of width $\sim \delta =\delta _1$, in which the appropriateness of the constant-$\psi _1$ condition is evidenced.

Therefore, we write $\psi _{1} = c_0$ in the whole non-ideal region and, as discussed in § 5.5, this allows us to more conveniently perform the integration by looking at the equations for $\psi _1$ and $\varphi _1$ rather than the auxiliary equation. In this way, (7.1) can be combined in the form

(7.60)\begin{equation} \left(1 + \frac{\rho_s^2}{\bar{d}_e^2} \frac{\tilde{\zeta}^2}{\tilde{\mathcal{G}^2}}\right) \tilde{\varphi}_1''\, =\, \frac{\rho_s^2}{\bar{d}_e^2} \frac{\tilde{\zeta}^2}{\tilde{\mathcal{G}^2}}\tilde{\varphi}_1 + c_0\frac{\rho_s^2}{\bar{d}_e^2}\tilde{\zeta}. \end{equation}

The results of § 8.3 in the small-$\varDelta '$ cold-electron limit are recovered by neglecting the $\rho _s^2\tilde {\zeta }^2\tilde {\varphi }_1''/(\bar {d}_e^2\tilde {\mathcal {G}^2})$ contribution in the left-hand side term of (7.60). Its presence here makes an intermediate matching region also appear in this wavelength limit.

This can be seen in figure 10, where the results of a numerical integration performed in the small-$\varDelta '$ limit of the warm collisionless regime ($\bar {d}_e=d_e$) with the solver of Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) are shown. The blue curve corresponds to the $\rho _s^2\tilde {\zeta }^2\tilde {\varphi }_1''/(\bar {d}_e^2\tilde {\mathcal {G}^2})$ term in (7.60). Its strong variation close to the neutral line is concentrated in a narrow region much smaller than $\delta _2=\rho _s$, which is of the order of $\delta _1$ (evaluated from the numerically computed eigenfunction, as it will be discussed in § 8). The green curve corresponds instead to the $\tilde {\varphi }_1''$ term in (7.60), i.e. to the first left-hand side contribution between parentheses, whereas the orange curve corresponds to the first term on the right-hand side of (7.60), $\rho _s^2\tilde {\zeta }^2\tilde {\varphi }_1/(\bar {d}_e^2\tilde {\mathcal {G}^2})$. This term can be neglected with respect to the one corresponding to the blue curve. For completeness, the contribution of the last right-hand side term of (7.60), $c_0\rho _s^2\tilde {\zeta }^2/\bar {d}_e^2$, is shown as the black, dashed line. From the comparison of these contributions, one verifies the appropriateness of the hypotheses justifying (7.4).

Figure 10. Comparison of the spatial profiles of the terms $\rho _s^2/(\bar {d}_e^2\tilde {\mathcal {G}}^2)\times \tilde {\zeta }^2\tilde {\varphi }_1''$ (blue curve), of $\rho _s^2/(\bar {d}_e^2\tilde {\mathcal {G}}^2)\times \tilde {\zeta }^2\tilde {\varphi }_1$ (orange curve) and of $\tilde {\varphi }_1$ (green curve) of (7.60), in a region of the domain with width $|x|<\delta _2=\rho _s$. The parameters of the numerical integration performed with the eigen-solver of Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) in the small-$\varDelta '$ limit are shown at the top of the figure. The value of $\delta =\delta _1$ in the bottom right corner has been evaluated according to the definition in (8.3).

Thus, an intermediate region, corresponding to the interval $\delta _1 < |x|\lesssim \delta _2$ of § 7.2.1, can be recognised: here, (7.4) hold, which we can combine in the form

(7.61)\begin{equation} \tilde{\zeta}\tilde{\varphi}_1'' = \tilde{\zeta}\tilde{\varphi}_1+c_0\bar{\mathcal{G}^2}. \end{equation}

However, as discussed in § 5.5, it is not necessary here to first perform the integration of (7.61), since the constant-$\psi$ condition holds in this case also in the innermost region.

7.5.1 Matching and scalings in the non-ideal region $x\ll 1$

Combining (7.8) is in practice equivalent – except for the chosen normalisation – to neglecting the $\sim \tilde {\zeta }^2\tilde {\varphi }_1/$ term in (7.60). Using the latter, where we recall lengths are normalised to $\rho _s$, we write

(7.62)\begin{equation} \tilde{\varphi}_1'' = {c_0}\frac{\rho_s^2}{\bar{d}_e^2} \frac{\tilde{\zeta}}{1 +{\rho_s^2} \tilde{\zeta}^2/(\bar{d}_e^2\tilde{\mathcal{G}}^2)}. \end{equation}

The equivalent of (3.10), obtained after integration of the second of (7.1) (cf. it also with (6.16)) reads

(7.63)\begin{equation} \int_{-\tilde{Z}}^{+\tilde{Z}} \tilde{\psi}_1''\, {\rm d}{\tilde{\zeta}} = \int_{-\tilde{Z}}^{+\tilde{Z}} \frac{\tilde{\varphi}_1''}{\tilde{\zeta}} \, {\rm d}\tilde{\zeta}, \end{equation}

where $\tilde {Z}=X/\rho _s\gg 1$ is a matching point in the interval $\delta _1 \ll X \ll \delta _2$ in units of $L_0=a$. Use of (3.9) allows one to establish the matching conditions with the outer solution by making explicit the dependence on $\varDelta '$: writing therefore $\int _{-\tilde {Z}}^{+\tilde {Z}} \tilde {\psi }_1'' = c_0\varDelta '$ and substituting (7.62) in (7.63), one obtains

(7.64)\begin{equation} c_0 \varDelta' = \int_{-\tilde{Z}}^{+\tilde{Z}} \frac{\tilde{\varphi}_1''}{\tilde{\zeta}} \,{\rm d}\tilde{\zeta}=c_0 \frac{\rho_s^2}{\bar{d}_e^2} \int_{-\tilde{Z}'}^{+\tilde{Z}'} \frac{{\rm d}\tilde{\zeta}}{1 +{\varrho}^2 \tilde{\zeta}^2} = c_0 \frac{\rho_s^2}{\varrho\bar{d}_e^2} \left.\arctan(\varrho\tilde{\zeta})\right|_{-\tilde{Z}}^{+\tilde{Z}}, \end{equation}

where $\varrho \equiv \rho _s/(\bar {d}_e\tilde {\mathcal {G}})$. Taking the asymptotic limit $\tilde {Z}\to \infty$ and using $\arctan (\bar {\varrho } \tilde {\zeta })|_{-\infty }^{+\infty }={\rm \pi}$, (6.16) reads

(7.65)\begin{equation} c_0 \varDelta' = c_0 {\rm \pi}\frac{\rho_s^2\tilde{\mathcal{G}}}{\bar{d}_e} \implies \gamma = \sqrt{\rm \pi} J_0 k\rho_s \bar{d}_e \varDelta' \implies \delta_1 \sim \varDelta'\bar{d}_e^2, \end{equation}

where in the last passage, we have used the definition in (7.11) of $\delta _1=\gamma \bar {d}_e/(k\rho _s J_0)$. We can thus explicate the scalings above in the purely collisionless and purely resistive limits. In the former, taking $\bar {d}_e=d_e$, one obtains

(7.66)\begin{equation} \gamma \sim J_0 k \varDelta' \rho_s d_e, \quad \delta_1 \sim \varDelta' d_e^2, \end{equation}

that is, the result of Porcelli (Reference Porcelli1991) (cf. the discussion between (8) and (9), therein). Taking instead $\bar {d}_e=S^{-1/2}\gamma ^{-1/2}$, one obtains

(7.67)\begin{equation} \gamma \sim J_0^{2/3} (k \varDelta')^{2/3} \rho_s^{2/3} S^{{-}1/3}, \quad \delta_1 \sim J_0^{{-}1/3} k^{{-}2/3} (\varDelta')^{1/3} \rho_s^{{-}2/3} S^{{-}2/3}, \end{equation}

that is, the result first obtained by Pegoraro & Schep (Reference Pegoraro and Schep1986) for the growth rate (cf. (39), therein) and by Zocco & Schekochihin (Reference Zocco and Schekochihin2011) for $\delta _1$ (which maps into $\delta _\eta$ of (B.96), therein).

7.5.2 Approximated eigenmodes in the non-ideal region $x\ll 1$

Approximated expressions for the eigenmodes can be obtained by integrating (7.62) twice, and by calculating the second of these integrals by parts: one finds the following approximate formula for the eigenfunction $\tilde {\varphi }_1(x /\delta _2)$ in a spatial interval $x\lesssim \delta _2\simeq \rho _s$ that covers the whole non-ideal region:

(7.68)\begin{equation} {\varphi}_1(x) \simeq \frac{c_0 \delta_1^3}{2\bar{d}_e^2} \left\{ \left( \frac{x}{\delta_1} \right) \ln\left( 1 + \left( \frac{x}{\delta_1} \right)^2 \right) + 2 \arctan\left( \frac{x}{\delta_1} \right) - 2 \left( \frac{x}{\delta_1} \right)\right\}. \end{equation}

Substituting (7.68) into (7.62), using the second part of (7.8) and integrating twice by parts, gives an approximated expression for $\psi _1$:

(7.69)\begin{equation} \psi_1(x) \simeq c_0 + \frac{c_0 \delta_1^2}{\bar{d}_e^2} \left\{ \left( \frac{x}{\delta_1} \right) \arctan\left( \frac{x}{\delta_1} \right) - \frac{1}{2} \ln\left( 1 + \left( \frac{x}{\delta_1} \right)^2 \right) \right\}. \end{equation}

From the expressions above, it is straightforward to find closed forms for both the magnetic field components and the current density perturbations, using their definitions ${B}_x(x) \equiv k \psi _{1}$, ${B}_y(x)\equiv -{\partial \psi _{1}}/{\partial x}$, and ${J}_z(x) \equiv - {\delta ^2\psi _{1}}/{\partial x^2}$ and (7.69). These lead to

(7.70)\begin{equation} {B}_x(x) ={-} k \sin(ky) \psi_{1}(x), \quad {B}_y(x) ={-}\frac{k\delta_1 c_0}{\bar{d}_e^2} \arctan\left( \frac{x}{\delta_1} \right), \end{equation}
(7.71)\begin{equation} {J}_z(x) ={-} \frac{c_0}{\bar{d}_e^2} \frac{1}{1 + \displaystyle{\left( \frac{x}{\delta_1} \right)^2}}. \end{equation}

In the appropriate limits, this equation corresponds to (74) of Pegoraro & Schep (Reference Pegoraro and Schep1986) and to (31) of Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020).

8 Operational definition of the reconnecting layer width, $\delta$, and other microscopic scales related to the eigenmodes

We have seen (§§ 6 and 7) that the boundary layer calculations allow one to determine the asymptotic scaling of the width of the layers, where approximate solutions of the eigenfunctions can be analytically evaluated. These characteristic widths appear as normalisation scales $\delta _1$ and $\delta _2$, which, although sometimes can be seen as ‘natural’, being suggested by the comparison of some terms in the equations (as in the cases we previously considered), are a priori arbitrary and are essentially determined by the algebra. Recognising some of these scales as indicative of the reconnecting layer width requires instead further insight (and hypothesis) of physical nature.

This is the subject we discuss in this section, where we are going to provide an operational definition allowing the measurement of the reconnecting layer width, $\delta$, which is in agreement with previous theoretical assumptions (i.e. theoretical definitions based on some physical insight that have been used in the previous literature), and which can be useful for both numerical and experimental quantitative estimates. The appropriateness of this definition is then shown by direct comparison with boundary layer calculations of §§ 6 and 7 and by numerical verification of its asymptotic scaling.

By analysing some local properties of the derivatives of the eigenfunctions close to or on the neutral line, we also identify the asymptotic scalings of some microscopic scales, which are associated with the inverse of these spatial gradients. The relevance of these scale lengths will be shown again by comparison between the results of the boundary layer calculations and the numerical results.

To this purpose, we use the adaptive multi-precision solver discussed by Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020), which had been specifically developed to address the generalised eigenvalue problem in a slab periodic box of dimension $[ -L_x/2, L_x/2]$, for $k$ assigned. This solver uses a compact finite difference scheme of tunable precision for the derivatives on a non-uniform grid along the $x$-direction. The use of a non-uniform grid and of tunable precision in the calculation allow us to resolve the inner layer even for microscopic, realistic values of the non-ideal parameters: the grid spacing in the inner layer can be so chosen to be much smaller than in the outer layer to save computing time without loosing accuracy. In that work, we verified the numerical scalings predicted by boundary layer calculations in different reconnection regimes. Although the scalings in the small-$\varDelta '$ regime are insensitive of the magnetic equilibrium profile, this is not the case for the large-$\varDelta '$ limit, as it had been already noted in a series of works by Cross & van Hoven (Reference Cross and van Hoven1971), Van Hoven & Cross (Reference Van Hoven and Cross1971), Van Hoven & Cross (Reference Van Hoven and Cross1973b) and Van Hoven & Cross (Reference Van Hoven and Cross1973a): in both (Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) and in the present article, the numerical results refer to the equilibrium profile (4.3) in a numerical box with $L_x = 4{\rm \pi}$.

8.1 Notions of reconnecting layer and estimates of its width in previous literature

The reconnecting layer width $\delta$, meant as the extension of the interval around the neutral line, where the reconnection process takes place and is mostly localised, is not per se unequivocally defined via boundary layer calculations. Its identification requires further ansatz based on physical assumptions. Because of this, in most of the early reference papers about boundary layer calculations performed in the warm fluid–electron regime, the notion of ‘reconnecting layer width’ has not been explicitly used (cf. Pegoraro & Schep Reference Pegoraro and Schep1986; Pegoraro et al. Reference Pegoraro, Porcelli and Schep1989; Porcelli Reference Porcelli1991). In this context, an early notion of ‘layer width’, identifiable as $\delta$, has just been used in some kinetic models for tearing modes (cf. Drake & Lee Reference Drake and Lee1977; Mahajan et al. Reference Mahajan, Hazeltine, Strauss and Ross1978, Reference Mahajan, Hazeltine, Strauss and Ross1979; Cowley et al. Reference Cowley, Kulsrud and Hahm1986). In particular, in some of these works (Drake & Lee Reference Drake and Lee1977; Cowley et al. Reference Cowley, Kulsrud and Hahm1986), it has been referenced as the ‘electron layer width’, because of its identification with the microscopic region dominated by electron dynamics, in contraposition to the broader ‘ion layer width’, where non-ideal effects are important, but at scales larger than those of electrons. In this sense, $\delta$ has been made to actually correspond to the innermost layer width, which we have named $\delta _1$ (the ‘ion layer’ being practically correspondent to $\delta _2$). This identification, which also agrees with the assumption made in cold-electron regimes (practically since the first work of Furth et al. (Reference Furth, Killeen and Rosenbluth1963) – cf. also Ottaviani & Porcelli Reference Ottaviani and Porcelli1995) where a single non-ideal layer can be identified, has been made more explicit in later works (Bhattacharjee et al. Reference Bhattacharjee, Germaschewski and Ng2005; Zocco & Schekochihin Reference Zocco and Schekochihin2011; Connor et al. Reference Connor, Hastie and Zocco2012b).

The physical argument which is at the basis of the identification $\delta \to \delta _1$, and which has been more or less explicitly stated in different works on tearing mode analysis, grounds on the idea that it is the innermost subdomain which contains the essential non-ideal physics allowing the reconnection process. For example, for $\rho _s> \bar {d}_e$, the non-ideal region extending up to $x/\rho _s\sim 1$ is not of interest in this sense, coherently with the fact that $\rho _s$ does not allow, per se, the onset of tearing instabilities: it is only in the innermost layer of width $\delta _1$ that both electron inertia ($d_e$) or resistivity ($S^{-1}$) and ion-sound FLR effects ($\rho _s$) are important. Because of this, $\delta _1$ is the natural candidate to be identified as the characteristic microscopic region where magnetic lines dragged there by the flow can intersect, by thus violating the frozen-in condition. And since this cannot occur as long as magnetic lines are frozen in the electron flow (cf. comment on (2.7)), this layer is arguably dominated by electron physics, whence the appropriateness of naming it the ‘electron layer.’

Alternative, yet similar, expressions have been frequently used in the literature to characterise the layer where the electron frozen-in condition is violated. In this regard, it is worth mentioning the ‘dissipation region (or layer)’, named this way with reference to the dissipation of magnetic flux in resistive reconnection (see, e.g. among many others, Parker Reference Parker1973; Mandt, Denton & Drake Reference Mandt, Denton and Drake1994; Shay et al. Reference Shay, Drake, Denton and Biskamp1998), and the probably even more common expression, ‘(electron) diffusion region’. Introduced with initial reference to the resistive magnetic diffusion in Sweet–Parker-like steady reconnection processes – see, e.g. Sonnerup (Reference Sonnerup1973) and Vasyliunas (Reference Vasyliunas1975) – the latter expression is currently used to identify the reconnecting layer in any reconnection processes, and thus also in all regimes of tearing-type modes (see, e.g. Drake & Kleva Reference Drake and Kleva1991; Hesse, Forbes & Birn Reference Hesse, Forbes and Birn1999; Le et al. Reference Le, Egedal, Ohia, Daughton, Karimabadi and Lukin2013) and several other works, especially connected to astrophysical plasma research). All these notions fit with the idea of identifying the extension of the reconnecting layer with that of the reconnecting current sheet. In practically all works on magnetic reconnection explicitly touching on the subject, the scale length $\delta$ is more or less explicitly assimilated also to the characteristic width of the ‘(reconnecting) current layer (or sheet)’.

However, in spite of the vast scientific literature, spanning almost seven decades, which addresses the subject of characterising the reconnecting layer in different reconnection scenarios and regimes, no precise and generally acquired operational definition seems to exist for it. Identifying the reconnecting region and measuring the spatial profile of physical quantities inside of it is a crucial element for the modelling of reconnection processes observed in Nature or in experiments (cf. e.g. Bratenahl & Yeates Reference Bratenahl and Yeates1970; Yamada et al. Reference Yamada, Ji, Hsu, Carter, Kulsrud, Ono and Perkins1997; Vaivads et al. Reference Vaivads, Khotyaintsev, André, Retino, Buchert, Rogers, Décréau, Paschmann and Phan2004; Yamada et al. Reference Yamada, Kulsrud and Ji2010). Even identifying its position and extension from measurements may be a non-trivial task, and indeed specific proxies are sought for this purpose in different reconnection regimes (e.g. the quadrupolar pattern of the magnetic field for Hall-reconnection, etc.). Also, when reconnection is known or is expected to occur because of tearing-type modes, for which quite accurate analytical estimates are available, quantitative information on $\delta$ obtained from measurements can give insight on important features of the reconnection process (we will dedicate §§ 9 and 10 of this work to this point and to the heuristic interpretation of the boundary layer calculation). Moreover, once compared with theoretical estimates, this quantitative information can give indication on the dominant non-ideal effects at play. We recall indeed that the scale $\delta _1$ obtained from boundary layer calculations is not ‘trivial’ (cf. §§ 9 and 10): for example, in the large-$\varDelta '$ limit of the $\rho _s > \bar {d}_e$ regime, we have seen the scale $\delta _1$ to be asymptotically smaller than both $\bar {d}_e$ and $\rho _s$. At the same time, despite this information being available since the first boundary layer calculations performed in the Fourier space in the warm electron regime, its interpretation in physical terms has not been univocal: while practically all early works agreed on recognising the non-ideal ion region to have extension of the order of $\rho _s$, the width of the current layer has been instead differently identified, on the basis of slightly different physical arguments, often depending on further hypotheses or heuristic estimates. For example, the current width in the linear, large-$\varDelta '$, warm-collisionless regime has been also estimated to be larger than $d_e$ (see, e.g. Drake & Kleva Reference Drake and Kleva1991; Ottaviani & Porcelli Reference Ottaviani and Porcelli1995), on the basis of the heuristic arguments developed by Drake & Lee (Reference Drake and Lee1977), Cowley et al. (Reference Cowley, Kulsrud and Hahm1986) and then re-discussed by Zocco & Schekochihin (Reference Zocco and Schekochihin2011), or of the order of $\sim d_e$ (see, e.g. Grasso et al. Reference Grasso, Pegoraro, Porcelli and Califano1999), based on the simple argument that $d_e$ is the characteristic scale related to the reconnection process in collisionless regimes (see, e.g. Vasyliunas Reference Vasyliunas1975). The distinction between current layer width and reconnecting layer width, which was done by Zocco & Schekochihin (Reference Zocco and Schekochihin2011), is a subtle point to which we will return later, at the end of § 10.5.

From all these argument follows the interest in seeking a priori estimates of $\delta$, which can be of general acceptance and do not ground on boundary layer calculations but can be directly implemented by starting from experimental or numerical data.

In the context of nonlinear numerical simulations of tearing mode reconnection, different examples of the estimate of $\delta$ have been proposed in the literature. Sometimes, $\delta$ has been identified in terms of the global profile of the eigenmode superposed on the equilibrium function. For example, Ali, Li & Kishimoto (Reference Ali, Li and Kishimoto2014) have measured the width of the current sheet as the distance between the local minima of $J_{z,1}(x,y_0)+J_{z,0}(x)$ with respect to the coordinate $x$, and for $y_0$ that corresponds to the ordinate of the X-point. The current layer width has also been identified (Tenerani et al. Reference Tenerani, Velli, Rappazzo and Pucci2015) by evaluating the distance between the local minima of what here we would name $B_{x,1}(x,y_0)\sim \psi _1(x)\cos (k(y-y_0))$. It can be noted that the estimate of Tenerani et al. (Reference Tenerani, Velli, Rappazzo and Pucci2015) relates to the numerical evaluation we later give of the inverse scale length $D'$ (cf. (10.10) and figure 14) rather than of $\delta$. In the context of the aforementioned work, this is however coherent since in the simulations of Tenerani et al. (Reference Tenerani, Velli, Rappazzo and Pucci2015), the instability of a large aspect ratio current sheet is studied, where a fastest growing mode exists (as evidenced already by Furth et al. Reference Furth, Killeen and Rosenbluth1963) for which $\varDelta '\sim 1/\delta$ (cf. Loureiro et al. Reference Loureiro, Schekochihin and Cowley2007; Bhattacharjee et al. Reference Bhattacharjee, Huang, Yang and Rogers2009; Del Sarto et al. Reference Del Sarto, Pucci, Tenerani and Velli2016; Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020). In the nonlinear simulations of Papini et al. (Reference Papini, Landi and Del Zanna2019b), instead, the reference value for $\delta$ has been taken by evaluating the full width (with respect to, let us say, the $x$ coordinate) at half-maximum of the total current $J_z$ minus its average background value, that is, $\delta$ has been evaluated as the width at half-height of a local estimation of $J_{z,1}(x)$.

8.2 Operational definition of the reconnecting layer width, $\delta$, and its scalings

Here, we propose a quantitative, operational definition for the measurement of $\delta$, which differs with respect to those previously suggested in the way we estimate the current layer width, at least for tearing-type reconnection. We start indeed by noting that, in agreement with the well-known identification ‘reconnecting layer $\leftrightarrow$ current sheet’, this layer can be identified as the region around the neutral line in which the current density related to the perturbation is concentrated. In our notation, this current density is $J_{z,1}=-\nabla ^2\psi _1$. Inside of this region, a velocity field is also concentrated, corresponding to an outflow parallel to the neutral line outwardly directed from the X-point, which is a hyperbolic point of the flow (see sketch in figure 11). In the notation used here, such a velocity field is $v_{y,1}=-\varphi _1'$.

Figure 11. Sketch of the hyperbolic pattern of the velocity field associated with linear perturbations around the X-point. The yellow stripe represents the reconnecting layer in the vicinity of the neutral line, where a current sheet $J_{z,1}$ of width $\sim \delta$ is concentrated (the width of the magnetic island, instead, may well trespass the width of this region, during its nonlinear evolution – cf. caption of figure 2). Note that the estimate $v_{x,1}\sim \gamma \delta$, with $v_{x,1}=-{\rm i}k\varphi _1$, holds inside of this region, whereas the $v_{y,1}=\varphi _1'$ component exhibits a change of sign on the neutral line and therefore displays a strong gradient across the $\varDelta x\simeq \delta$ interval.

The profiles of $J_{z,1}(x)$ and of $v_{y,1}(x)$ close to the neutral line (cf. figure 12) are qualitatively analogous, so that both their respective characteristic ‘thicknesses’ could be taken as candidates for $\delta$. Therefore, by referring to the eigenfunctions only, we here consider the distance from the neutral line of the local maxima (or minima) of the gradient of the current density ($J_z'$) and to the distance from the neutral line of the local maxima (or minima) of the vorticity ($\varphi ''$). Since these two distances can in principle differ, we respectively name them ${\delta }_\psi$ and ${\delta }_\varphi$:

(8.1)\begin{equation} {\delta}_{\psi}\,:\quad \left.{J_{z,1}''}\right|_{x=\delta_{\psi}}= \left.{\psi_1^{iv}}\right|_{x=\delta_{\psi}}=0, \quad {\delta}_{\varphi}\, :\quad \left.{v_{y,1}''}\right|_{x=\delta_{\phi}}= \left.{\varphi_1'''}\right|_{x=\delta_{\varphi}}=0. \end{equation}

As shown in figure 12, ${\delta }_\psi$ and ${\delta }_\varphi$ can be easily calculated once the profile of the corresponding eigenfunction has been computed.

Figure 12. Examples of evaluation of the half-width of the current layer as the distance of the local maxima (minima) of $\psi _1'''=-J_{z,1}'$ (a) and of $\varphi _1''$ (b) from the neutral line (the parameters of the numerical calculation are indicated at the top of the panels).

A numerical scan performed in the whole parameter range of the regimes considered by Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) indicates that ${\delta }_\psi$ and ${\delta }_\varphi$ approximatively display the same asymptotic scaling, as shown in figure 13, as an example, for the warm-collisionless reconnection regime. We can so write ${\delta }_{\psi }\sim {\delta }_{\varphi }$, even if a proportionality factor not much different from unity is present (for example, we measured ${\delta }_{\varphi }\simeq 2{\delta }_{\psi }$ in warm-collisionless RMHD with $\rho _s/d_e\simeq 10$), which seems to display a weak dependence on the non-ideal parameters involved, at least in the parameter range we have numerically investigated.

Figure 13. Scalings of $\delta _\varphi$ (a,c) and $\delta _\psi$ (b,d) calculated using the definitions given by (8.1). The scalings, shown on the different panels, prove that these two quantities follow the same scaling laws.

Based on the coherence with the numerical results that show the correspondence of the scalings of $\delta _1$ with those of $\delta _\psi$ in all the reconnection regimes considered (also when electron–electron viscosity is included – see Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020),

(8.2)\begin{equation} we\ identify\ \delta\ as\ twice\ the\ distance\ from\ the\ neutral\ line\ of\ the inflection\ point\ of\ J_z. \end{equation}

Accordingly, we propose to operationally define the layer width as

(8.3)\begin{equation} \delta\equiv {\delta}_{\psi}\sim \delta_1. \end{equation}

The symbols $\delta _{SD}$ and $\delta _{LD}$ can be therefore used to indicate $\delta \equiv {\delta }_{\psi }\sim \delta _1$ in the small- and large-$\varDelta '$ regimes.

8.3 Further micro-scales related to the gradients of the eigenmodes on the neutral line

A further ensemble of characteristic spatial scales of the system is provided by the normalised derivatives of the eigenfunctions evaluated on the neutral line. These can be shown to be related to combinations of powers of the non-ideal parameters and of $\delta _1$. We define

(8.4)\begin{equation} {\delta}_{\psi}^{(N)}\equiv \left|\frac{\psi_1^{(N)}}{\psi_1}\right|^{-({1}/{N})}_{x=0}, \quad N=2, 4, 6,\ldots, \end{equation}

where $N$ is the order of the derivative with respect to the shear variable $x$.

These length scales can be related to the local expansion of the eigenfunctions in a neighbourhood of $x=0$. Using the fact that both $\psi _1$ and $v_{y,1}(x)=\varphi _1'(x)$ are even and are continuous, at least up to the third derivative with respect to $x$, we can write

(8.5)\begin{gather} \psi_1(x)\simeq c_0 \left\{1 \pm \frac{1}{2}\left(\frac{x}{\delta_\psi^{(2)}}\right)^2 \pm \frac{1}{4!}\left(\frac{x}{\delta_\psi^{(4)}}\right)^4 \pm \ldots\right\}, \end{gather}
(8.6)\begin{gather} v_y(x)=\varphi_1'(x)\simeq q_0 \left\{1 \pm \frac{1}{2}\left(\frac{x}{\delta_{v_y}^{(2)}}\right)^2 \pm \frac{1}{4!}\left(\frac{x}{\delta_{v_y}^{(4)}}\right)^4 \pm \ldots\right\}. \end{gather}

Note that the coefficients of (8.6) are related to those of (8.5). Using indeed ${\rm d}\varphi _1(\zeta )/{\rm d}\zeta =\zeta {\rm d}\psi _1(\zeta )/{\rm d}\zeta -\psi _1(\zeta )-\chi _\infty$ (cf. (5.11) and (5.12)) for $\zeta =x/\delta _1$, whence ${{\rm d} x}/{\rm d}\zeta =\delta _1$, and taking into account the normalisation of $\varphi _1(\zeta )$, i.e. $\varphi _1(\zeta )=A\varphi _1(x)$ with $A=-i\gamma /(k\delta _1 J_0)$ in the cold-electron regime and $A=-i\rho _s/\bar {d}_e$ in the warm-electron regime, one obtains

(8.7)\begin{equation} \frac{{\rm d}\varphi_1}{{\rm d} x}=\frac{x}{A\delta_1}\frac{{\rm d}\psi_1}{{\rm d} x}-\frac{\psi_1}{A\delta_1}-\frac{\chi_\infty}{A\delta_1}. \end{equation}

Direct substitution of ((8.5) and (8.6)) into (8.7) and comparison of equal powers of $x$ yields

(8.8)\begin{equation} q_0\sim{-}\frac{c_0}{A\delta_1}-\frac{\chi_\infty}{A\delta_1}, \quad \delta_{v_y}^{(N)}\sim \left(1+\frac{\chi_\infty}{c_0} \right)^{{1}/{N}} \left(N-1 \right)^{{1}/{N}} \delta_{\psi}^{(N)}. \end{equation}

Using then (5.14) and (5.17), we can write

(8.9)\begin{equation} q_0\sim \frac{O(\varDelta'\delta_1)}{A\delta_1} \quad \text{for}\ \varDelta'\delta_1\ll 1, \quad q_0\sim{-}\frac{c_0}{A \delta_1}\quad \text{for}\ \varDelta'\delta_1\gg 1. \end{equation}

Notice that the first of (8.9) is of scarce utility here, since not accurate enough: the estimate (5.14), which identifies $\chi _{\infty }\sim -c_0$ by neglecting possible corrections roughly estimable as being of the order of $\sim O(\varDelta '\delta _1)$, cannot be indeed used for the present calculations, as it would always yield $v_{y}(0)=0$. This does not generally agree with the analytical results of the boundary layer calculations in the small-$\varDelta '$ limit (cf. e.g. figure 14c obtained via numerical integration of the eigenvalue problem). Therefore, when needed, depending on the reconnecting regime and wavelength limit considered next, we will rely on a more accurate estimate obtained from the boundary layer solution found in the case of interest.

Figure 14. (a) Example of numerical evaluation of the coefficients $c_0$ and $c_1$ of (10.10), once the profile of $\psi _1(x)$ and of its derivatives $\psi _1'(x)$ have been computed for $|x|\geq \delta$ in a neighbourhood of $x=\delta$. (b) Analogous example of evaluation of $q_0$ and $q_1$ of (10.10) from $\varphi _1'$ and $\varphi _1''$, calculated once the eigenfunction $\varphi _1$ has been computed; in this example, $\varDelta '_{v_y}\approx 73.82$. (c) A zoomed version of panel (b) in which the inner region, here of width $\delta \approx 0.00184$ (computed according to the method outlined in § 8), is shaded in light-blue colour. All the values required to estimate $\varDelta '_{v_y}$ using (10.2) are here shown; it is this way that $\varDelta '_{v_y}\approx 74.48$ is obtained, in very good agreement with the result computed with the alternative method sketched in panel (b).

The scalings of some of the length scales $\delta ^{(N)}_\psi$ are easily found to be related to those of ${\delta }_1$ (and hence of $\delta \sim \delta _\psi$). This can be proven analytically. A possibility, quite generally applicable, consists in combining ((2.1) and (2.2)) by using $\varphi _0(0)=0$ with the local expansion

(8.10)\begin{equation} \psi_0(x) \simeq C_0 + \frac{C_2}{2} x^2 + \frac{C_4}{4!}x^4 + \mathcal{O}(x^4), \end{equation}

and in differentiating the eigenfunctions the number of times which is needed. Note the correspondence $C_2=J_0$ in the notation used in previous sections. The sought scalings can be obtained by balancing the dominant terms of the equations while taking the limit $x\to 0$, and by using the scalings of $\gamma$ and $\delta _1$ obtained from boundary layer calculations. An alternative procedure consists in the direct evaluation of these derivatives from the approximated eigenmode solutions close to the neutral line, once they are obtained by integration of the boundary layer problem.

Here, below (§ 8.3.1), we use the second procedure to compute $\delta ^{(2)}_\psi$ and the former to obtain $\delta ^{(4)}_\psi$.

8.3.1 Asymptotic scalings of $\delta ^{(2)}_\psi$ and of $\delta ^{(4)}_\psi$

The scalings that can be obtained for $\delta ^{(2)}_\psi$ and $\delta ^{(4)}_\psi$ in the different regimes are summarised in table 2.

Table 2. Scaling of the spatial scales associated with the second- and fourth-order derivative of $\psi _1(x)$, evaluated on the neutral line $x=0$ in different reconnection regimes and wavelength limits. In each regime, $\delta _{SD}$ and $\delta _{LD}$ are those of table 1 and correspond to the different values of $\delta _1$ evaluated in §§ 6 and 7.

The estimate of $\delta _{\psi }^{(2)}$ is of interest, since it provides a direct estimate of the peak amplitude of the current density on the neutral line during the linear stage of the tearing mode evolution. According to the values in table 2, $J_z|_{x=0}\simeq c_0/\bar {d}_e$. Also, the scale of $\delta _{v_y}^{(2)}$ is of potential interest, since it gives the characteristic curvature with respect to $x$ of the velocity profile $v_{y}(x)$ on the neutral line. While, according to (8.8) and (8.9), $\delta _{v_y}^{(2)}\sim \bar {d}_e$ in the large-$\varDelta '$ limit, in the small-$\varDelta '$ limit, its value is typically asymptotically smaller, since it depends on the estimate of $(1+\chi _{\infty }/c_0)\ll 1$ (cf. e.g. (8.21) for the cold-electron limit). These estimates are likely to be relevant for the study of the stability of the Bickley jet (Bickley Reference Bickley1937) related to the $v_y(x)$ velocity component, which, especially in the cold collisionless, large-$\varDelta '$ regimes, nonlinearly develops along the neutral line and for $d_e^2\gg \rho _s^2$, leads to a turbulent regime via the onset of secondary Kelvin–Helmholtz instabilities (Del Sarto et al. Reference Del Sarto, Califano and Pegoraro2003, Reference Del Sarto, Califano and Pegoraro2006). A similar feature had been observed also in early nonlinear simulations of the collisionless internal-kink mode in cylindrical geometry (Biskamp & Sato Reference Biskamp and Sato1997) and had motivated dedicated studies of the stability of the Bickley jet in the presence of a background magnetic field aligned to it (Biskamp, Schwarz & Zeiler, Reference Biskamp, Schwarz and Zeiler1998). The destabilisation via Kelvin–Helmholtz of a Bickley jet developing during the nonlinear stage of the collisionless reconnection process has been confirmed also in nonlinear simulations of tearing modes in three-dimensional geometry (Grasso, Borgogno & Pegoraro Reference Grasso, Borgogno and Pegoraro2007; Grasso et al. Reference Grasso, Borgogno, Pegoraro and Tassi2009, Reference Grasso, Borgogno, Tassi and Perona2020).

The estimate of $\delta _{\psi }^{(4)}$ is also of interest, since it intervenes in the modelling of the nonlinear current sheet evolution, which in the purely collisionless regime has been shown to shrink exponentially in time (Ottaviani & Porcelli Reference Ottaviani and Porcelli1993, Reference Ottaviani and Porcelli1995).

Below, we separately evaluate $\delta ^{(2)}_\psi$ and $\delta ^{(4)}_\psi$ so to prove the scalings reported in the table.

8.3.2 Asymptotic scalings of $\delta ^{(2)}_\psi$

The scaling of $\delta ^{(2)}_\psi$ follows from (2.1), which, in the purely collisionless limit, both ‘cold’ and ‘warm’, implies the conservation of the electron canonical momentum on the neutral line because of (8.10): $\partial _t(\psi -d_e^2\nabla ^2\psi )|_{x=0}=0$. The result can be formally extended to the resistive regime by relying again on the generalised electron skin depth $\bar {d}_e$. More precisely, regardless of the value of $\rho _s$, we obtain from (2.1)

(8.11)\begin{equation} {\delta}_{\psi}^{(2)}= \frac{\bar{d}_e}{\sqrt{1+k^2\bar{d}_e^2}}\simeq \bar{d}_e, \end{equation}

since in both the large- and small-$\varDelta '$ regimes, purely resistive or purely inertial, $k^2\bar {d}_e^2\ll 1$. The above expression can also be approximatively written as

(8.12)\begin{equation} \left. \frac{\psi_1''}{\psi_1} \right|_{x=0}\,\simeq\,\frac{1}{\bar{d}_e^2}\quad \implies \quad \delta^{(2)}_\psi \simeq \bar{d}_e. \end{equation}

The corresponding scalings in table 2 descend from (8.12), after the relevant parameter and wavelength limits are considered.

8.3.3 Asymptotic scalings of $\delta ^{(4)}_\psi$

To evaluate $\delta ^{(4)}_\psi$, we should differentiate the solutions $\psi _1$ obtained in each reconnection regime and wavelength limit. Save for the warm-electron, large-$\varDelta '$ limit, where the last equality of (7.38) immediately gives $|(d^4\psi _1/d\zeta ^4)/\psi _1|_{\zeta =0}\, \simeq 2 \delta _1^2/\bar {d}_e^2 + O( \delta _1^4/\bar {d}_e^4 )$, whence $\delta ^{(4)}_\psi$ can be rapidly evaluated by using $\zeta = x/\delta _1$ and then $\delta _1={{\rm d}x}/{{\rm d}\zeta }$, in all other cases, some more algebra is required. It is therefore interesting to look if it is possible to address all the regimes and cases in a unified way. To this purpose, we can combine (8.10) and (8.5) and (8.6) with the limit $x\to 0$ of the second-order derivative of (2.1) and of the first-order derivative of (2.2) with respect to $x$. Having introduced once more $\bar {d}_e$ so as to treat the inertial and resistive cases all together, we obtain

(8.13)$$\begin{gather} \gamma\left((1+k^2\bar{d}_e^2)\psi_1''-\bar{d}_e^2\psi_1^{iv}\right)|_{x=0}\simeq {\rm i}2C_2k\left(\rho_s^2\varphi_1'''-\varphi_1'\right)|_{x=0}, \end{gather}$$
(8.14)$$\begin{gather}\gamma\varphi_1'''|_{x=0} \simeq {\rm i}k\left(C_2\psi_1''-C_4\psi_1\right)|_{x=0}. \end{gather}$$

Combining them and using (8.6),

(8.15)\begin{equation} \gamma^2\left( (1+k^2\bar{d}_e^2)\psi_1''-d_e^2\psi_1^{iv}\right)|_{x=0}\simeq{-}2C_2k^2\rho_s^2(C_2\psi_1''-C_4\psi_1)|_{x=0} -{\rm i}2C_2k\gamma q_0. \end{equation}

Finally, using (8.12),

(8.16)\begin{align} -\frac{\bar{d}_e^4}{1+k^2\bar{d}_e^2} \left. \frac{\psi_1^{iv}}{\psi_1}\right|_{x=0}\simeq{-}1 -2(C_2)^2\frac{k^2}{1+k^2\bar{d}_e^2} \frac{\rho_s^2}{\gamma^2}\,+\, 2C_2\frac{k^2}{1+k^2\bar{d}_e^2} \frac{\bar{d}_e^2}{\gamma^2}\left(\rho_s^2C_4 -{\rm i}\frac{\gamma q_0}{kc_0}\right). \end{align}

At this point, we specialise the result to the cold and to the electron limits.

  • Cold-electron limit

    In the cold limit $\rho _s=0$, using $A=-{\rm i}\gamma /(k\delta _1 J_0)$, $q_0\sim -c_0/(A\delta _1)$ or $\sim -2c_0/(A\delta _1)$ (cf. (8.9)), we can so distinguish two cases.

    1. (a) One case is for $k^2\bar {d}_e^2\delta _1/\gamma ^2\sim d_e\ll 1$, which corresponds to the large-$\varDelta '$ limit, in which we obtain the scaling

      (8.17)\begin{equation} \left|\frac{\psi_1^{iv}}{\psi_1}\right|_{x=0} \simeq \frac{1}{\bar{d}_e^4}\quad \implies \quad \delta^{(4)}_\psi \simeq \bar{d}_e\sim \delta_1\sim\sqrt{\bar{d}_e\delta_{LD}}. \end{equation}
    2. (b) The second case is for $k\bar {d}_e^2q_0/(c_0\gamma )\gg 1$, which is true in the small-$\varDelta '$ limit, for which $\gamma \bar {d}_e/k \sim \delta _1^2$ (cf. (6.19), but for which a better estimate of $q_0$, as given by the first part of (8.9), must be found. To this purpose, we can directly use the solution found for $\varPhi (z)=(\bar {d}_e k^3J_0^3/\gamma ^3)^{1/2}\varphi _1(x(kJ_0)^{1/2}/(\gamma \bar {d}_e)^{1/2})$ from boundary layer calculations performed in this regime (§ 6.2). We first take the derivative of (6.15):

      (8.18)\begin{equation} \frac{{\rm d}\varPhi}{{\rm d}z} ={-}\frac{1}{2} \int_0^1 (1 - t^2)^{-{1}/{4}} e^{-({1}/{2}) t z^2} \,{\rm d}t + z^2 \int_0^1 t (1 - t^2)^{-{1}/{4}} e^{-({1}/{2}) t z^2} \,{\rm d}t. \end{equation}
      Since $z\propto x$, we can evaluate $Q_0\equiv \varPhi '(z)|_{z=0}$ and then relate it to $q_0=\varphi _1'(x)|_{x=0}$. From (8.18), we obtain
      (8.19)\begin{equation} Q_0 ={-}\frac{1}{2} \int_0^1 \frac{{\rm d}t}{(1 - t^2)^{{1}/{4}}} ={-} \int_0^{{\rm \pi}/{4}} \sqrt{1 - 2 \sin^2(u)}\, {\rm d}u ={-} \mathcal{E}\left(\left.\frac{\rm \pi}{4} \right| 2 \right), \end{equation}
      where in the last passage, we have made the changes of variables of integration $t = \sin \theta$ and $u = \theta /2$, and where $\mathcal {E}$ is the incomplete elliptic integral of second kind (see Gradshteyn & Ryzhik Reference Gradshteyn and Ryzhik2015, (8.2) in § 8.111). Using (6.14), we thus find
      (8.20)\begin{equation} q_0 = {\rm i}\frac{c_0 \gamma}{k\bar{d}_eJ_0} Q_0 ={-} {\rm i} \frac{c_0 \gamma}{k\bar{d}_eJ_0} \mathcal{E}\left(\left.\frac{\rm \pi}{4} \right| 2 \right). \end{equation}
      Finally, using (6.21), one finds
      (8.21)\begin{equation} q_0 ={-} {\rm i} \frac{c_0 \left( \varDelta' \bar{d}_e\right)^2}{I^2} \mathcal{E}\left(\left.\frac{\rm \pi}{4} \right| 2 \right). \end{equation}
      This leads to
      (8.22)\begin{equation} \left.\frac{\psi_1^{iv}}{\psi_1}\right|_{x=0}\simeq{-}{\rm i}2\frac{C_2}{I^2}\mathcal{E}\left(\left.\frac{\rm \pi}{4} \right| 2 \right) \frac{k\varDelta'}{\gamma}, \end{equation}
      that is,
      (8.23)\begin{equation} \left|\frac{\psi_1^{iv}}{\psi_1}\right|_{x=0} \simeq\frac{1}{\varDelta'\bar{d}_e^3}\quad \implies \quad \delta^{(4)}_\psi \simeq \sqrt{\bar{d}_e\delta_{SD}}. \end{equation}

  • Warm-electron limit

    In the warm limit $\rho _s>\bar {d}_e$, we can assume instead the first term of (8.22) to be always dominant at the right-hand side and greater than unity, so as to write

    (8.24)\begin{equation} \left.\frac{\psi_1^{iv}}{\psi_1}\right|_{x=0}\simeq 2C_2^2\frac{k^2}{\gamma^2} \frac{\rho_s^2}{\bar{d}_e^4}. \end{equation}
    Using finally the definition (7.11), we obtain, in both the small- and large-$\varDelta '$ limits,
    (8.25)\begin{equation} \left|\frac{\psi_1^{iv}}{\psi_1}\right|_{x=0} \simeq \frac{1}{\bar{d}_e^2\delta_1^2}\quad \implies \quad \delta^{(4)}_\psi \simeq \sqrt{\bar{d}_e\delta_1}. \end{equation}
    Specialising $\bar {d}_e$ in the purely inertial or purely resistive regime and taking the relevant scalings in the large-$\varDelta '$ allows us to complete table 2.

9 Heuristic derivation of the scaling laws of tearing modes

In previous sections, we obtained the growth rate scaling law by analytically solving the eigenvalue problem in the warm- and cold-collisionless regimes. To solve the equations, we expanded the equilibrium profile around the neutral line using a Taylor series. Therefore, despite the cumbersome analysis developed to find them, the eigenfunctions so obtained are merely approximations and not ‘exact’ solutions of the problem. The complexity of the boundary layer approach is also evident, which leads to differential equations of hypergeometric nature.

A complementary approach is possible, which, in some cases as we are going to see, allows the estimate of the scalings of both the growth rate and of the reconnecting layer width by providing some more physical insight about the analytical assumption made in the boundary layer formalism. This method is based on some heuristic orderings of the terms in the inner layer equations, and on balancing these terms together to obtain the scaling laws via dimensional analysis – see Drake & Lee (Reference Drake and Lee1977) and Cowley et al. (Reference Cowley, Kulsrud and Hahm1986) for an application to kinetic tearing, Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020) for an application to the cold resistive and viscous-resistive regimes, and Drake & Kleva (Reference Drake and Kleva1991) for an application to secondary instabilities to a primary tearing-type mode. Although it has not been detailed, this approach has been also used to get the collisionless scalings of tearing modes in Ottaviani & Porcelli (Reference Ottaviani and Porcelli1995). An analogous approach is at the basis of the available theoretical estimates of the scalings of the reconnecting rate in the whistler-mediated reconnection scenario (Mandt et al. Reference Mandt, Denton and Drake1994) and of the reconnecting rate in Hall-MHD reconnection (Biskamp, Schwarz & Drake Reference Biskamp, Schwarz and Drake1995, Reference Biskamp, Schwarz and Drake1997). Heuristic ansatz on the scaling of the gradients of the tearing eigenfunction have been revealed to be useful also for quantitative estimates – which have been numerically verified a posteriori – about the time and spatial behaviour of the reconnecting current sheet during its nonlinear, collisionless evolution (Ottaviani & Porcelli Reference Ottaviani and Porcelli1993, Reference Ottaviani and Porcelli1995). More in general, heuristic estimates are of fundamental importance to allow insight on the physical interpretation of less evident analytical results (see, e.g. Drake & Lee Reference Drake and Lee1977; Cowley et al. Reference Cowley, Kulsrud and Hahm1986 for tearing modes and Grasso et al. (Reference Grasso, Pegoraro, Porcelli and Califano1999) for the interpretation of the physical meaning of $\rho _s$ in reduced MHD reconnection).

While this heuristic method, with some variations, is frequently presented in textbooks as a shortcut procedure to find the scalings of the cold-electron, resistive tearing mode (see, e.g. Biskamp Reference Biskamp2000, § 4.1.1; Schnack Reference Schnack2009, Lecture 34; Boyd & Sanderson Reference Boyd and Sanderson2003, § 5.3.1, to give some examples), its application cannot be clear when more than one boundary layer exists, as it is in the case of warm-electron tearing modes. Discussing this point is therefore of general interest: this is what we are going to do in this section, where we compare the heuristic approach to the boundary layer analysis presented in previous sections. In particular, we are going to show that further information is required to get consistent results from the heuristic analysis to get the correct scalings when $\rho _s\gtrsim \bar {d}_e$. From preliminary analysis, this information appears not to be immediately available from a priori arguments. This suggests that the heuristic approach should be carefully handled, when electron temperature effects (and, more in general, FLR effects) are included, since it could lead to incorrect estimates, as we are going to show below.

9.1 General hypotheses in the heuristic approach to the scaling estimate

Let us first outline the general hypotheses, which allow one to recover the correct scaling laws by dimensional analysis in the textbook-like examples of the purely resistive and of the purely inertial tearing mode analysis.

We first re-write the eigenvalue equations for $x\ll 1$ (i.e. $x/a\ll 1$ in dimensional units), in the non-ideal region:

(9.1)\begin{gather} \psi_1 - {\rm i} \frac{k x}{\gamma} J_0 \varphi_1 = \bar{d}_e^2 \psi_1'' + \frac{\rho_s^2 }{\gamma^2} k^2 x^2 J_0^2 \psi_1'', \end{gather}
(9.2)\begin{gather} \varphi_1'' ={-}{\rm i} \frac{kx}{\gamma}J_0 \psi_1'', \end{gather}

where (9.2) was used to express the second term on the right-hand side of (9.1), and resistivity has been included in the parameter $\bar {d}_e^2$.

The usual heuristic estimations, as they have been successfully used for both the purely collisionless and purely resistive regimes, are based on the following ideas.

  1. (i) There is a characteristic scale, say $l_c$, for the gradient of $\psi _1$, which we are going to quantitatively define in the following. It allows one to estimate $\psi _1' \sim \psi _1/l_c$ at some point $x$ in the neighbourhood of the neutral line.

  2. (ii) A single characteristic microscopic scale exists for both the first derivative of $\psi _1'$ and $\varphi _1$: this corresponds to the inner layer width, $\delta$, which we operationally define according to (8.3), (this can be proven via numerical integration of the equations – see later).

  3. (iii) We can generally assume $\delta \leq l_c$.

We then add a further assumption that can be a posteriori verified (also numerically), and somewhat generalises the examples for which the heuristic approach has been successfully applied in the past.

  1. (iv) $l_c$ is the largest characteristic scale length in the matching layer with the ideal-MHD solution. That is, in a neighbourhood of the neutral line, we write the estimates

    (9.3)\begin{equation} \psi_1' \sim \frac{\psi_1}{l_c}, \quad \psi_1'' \sim \frac{\psi_1}{\delta l_c}. \end{equation}

The two characteristic scales that naturally appear when a distinction has been made between the large-$\varDelta '$ and small-$\varDelta '$ limits are ($\varDelta ')^{-1}$ and $\delta$. Accordingly, in the small-$\varDelta '$ limit, $l_c = (\varDelta ')^{-1}$ with $\varDelta '\delta \ll 1$, while in the large-$\varDelta '$ limit, $l_c = \delta$ and $\varDelta '\delta \gg 1$. This argument suggests the following scalings (Ottaviani & Porcelli Reference Ottaviani and Porcelli1995):

(9.4)\begin{equation} \frac{\psi_1''}{\psi_1} \sim \frac{1}{\delta^2} \quad \text{for} \ (\varDelta' \delta \gg 1), \quad \frac{\psi_1''}{\psi_1} \sim \frac{\varDelta'}{\delta} \quad \text{for} \ (\varDelta' \delta \ll 1). \end{equation}

Note that, differently from (8.12), where the ratio $\psi _1'' /\psi _1$ is evaluated exactly on the neutral line ($x\equiv 0$), in the estimates of (9.4), it is evaluated in the neighbourhood of the line (i.e. $x\simeq 0$).

After approximating $x \sim \delta$ in the inner layer, (9.2) gives

(9.5)\begin{equation} \varphi_1'' \sim \frac{k\delta}{\gamma}J_0 \psi_1'', \end{equation}

which is true in all tearing regimes. All further estimates rely on assumptions about the relative ordering between the terms of the equations. In what follows, we will discuss the derivation of the scaling laws in the different regimes by using this method: it will prove to be successful in the cold-collisionless limit, but we will see that it fails in the warm-collisionless regime.

9.2 Heuristic derivation of the scaling laws in the cold-electron regime ($\rho _s^2\ll \bar {d}_e^2$)

In these regimes, we can take $\rho _s = 0$. Therefore, the second term on the right-hand side of (9.1) vanishes.

For reasons of ‘economy of thought’ and of convenience about the generalisability of the heuristic approach, which will be discussed next (cf. comments on (10.4) in § 10.1), we now follow a procedure, which, although practically equivalent to those that can be found in classical textbook examples, slightly differs from most of them, as far as some ansatz are concerned: in particular, we are not going to make use of the estimate $\varphi _1''\sim \varphi _1/\delta _1^2$, otherwise typically used, and which can be verified to be valid in the cold-electron regimes.

We then start by balancing the remaining terms of (9.1) and (9.2). This leads us to

(9.6)\begin{equation} \psi_1\sim \bar{d}_e^2 \psi_1'', \quad \varphi_1 \sim \frac{\gamma }{k\delta J_0} \psi_1, \quad \varphi_1 \sim \frac{\gamma \bar{d}_e^2}{k\delta J_0} \psi_1''. \end{equation}

Differentiating twice the second part of (9.6) and using (9.5), one gets

(9.7)\begin{equation} {\gamma}\sim {k\delta J_0}, \end{equation}

which is valid for both wavelength limits and in both the collisionless and resistive regimes. Substituting (9.4) in the first part of (9.6), one obtains

(9.8)\begin{equation} \delta \sim \bar{d}_e \quad \text{for} \ (\varDelta' \delta \gg 1), \quad \delta \sim \bar{d}_e^2 \varDelta' \quad \text{for} \ (\varDelta' \delta \ll 1). \end{equation}

It can be noticed that, looking at the physical aspects, the first of the conditions in (9.6) is the result of the balance between the two terms directly involved in the process of energy conversion that is related to magnetic reconnection: the magnetic potential $\psi _1$ on the one side, and, on the other side, the electron kinetic energy $d_e^2\psi _1''$ in the collisionless limit or the energy dissipated by Ohm's law in the resistive limit, $S^{-1}\psi _1''$.

At this point, it is convenient to treat the purely inertial and the purely collisionless case separately.

  1. (a) Collisionless case.

    Substituting $\bar {d}_e\to d_e$ and (9.8) into (9.7) yields

    (9.9)\begin{equation} \gamma \sim J_0 k d_e \quad \text{for} \ \varDelta' \delta \gg 1, \quad \gamma \sim J_0 k (\varDelta')^2 d_e^3 \quad \text{for} \ \varDelta' \delta \ll 1. \end{equation}
    The corresponding scalings of the width of the reconnecting layer read
    (9.10)\begin{equation} \delta \sim d_e \quad \text{for} \ \varDelta' \delta \gg 1, \quad \delta \sim \varDelta' d_e^2 \quad \text{for} \ \varDelta' \delta \ll 1. \end{equation}
    These are the scaling laws of Porcelli (Reference Porcelli1991) that we have analytically obtained in § 6.
  2. (b) Resistive case.

    Substituting instead $\bar {d}_e\to S^{-1/2}/\gamma ^{1/2}$ and proceeding as above yields the scalings of Furth et al. (Reference Furth, Killeen and Rosenbluth1963) and Coppi et al. (Reference Coppi, Galvao, Pellat, Rosenbluth and Rutherford1976) (see also Ottaviani & Porcelli Reference Ottaviani and Porcelli1995), which we have also already obtained in § 6:

    (9.11)\begin{gather} \gamma \sim J_0^{2/3} k^{2/3} S^{{-}1/3} \quad \text{for} \ \varDelta' \delta \gg 1, \quad \gamma \sim J_0^{2/5} k^{2/5} (\varDelta')^{4/5} S^{{-}3/5} \quad \text{for} \ \varDelta' \delta \ll 1), \end{gather}
    (9.12)\begin{gather} \delta \sim J_0^{{-}1/3} k^{{-}1/3} S^{{-}1/3} \quad \text{for} \ \varDelta' \delta \gg 1, \quad \delta \sim J_0^{{-}2/5} k^{{-}2/5}(\varDelta')^{1/5} S^{{-}2/5} \quad \text{for}\ \varDelta' \delta \ll 1. \end{gather}

9.3 Heuristic derivation of the scaling laws in the warm-electron regime ($\rho _s\gtrsim \bar {d}_e$)

If we now follow an analogous approach in the warm-collisionless regime, problems arise suggesting some (or all) of the hypotheses made at points (i)–(iii) and in (9.3)(9.5) to not be correct. Let us see why.

In this regime, the second term on the right-hand side of (9.1) does not vanish. Therefore, it is expected that this term will be of the same order of the first right-hand side term at the boundary of the inner layer (i.e. at $|x|\simeq \delta$). Balancing these two terms yields

(9.13)\begin{equation} \frac{\gamma}{k J_0} \sim \frac{\rho_s}{\bar{d}_e} \delta. \end{equation}

By balancing the first left-hand side terms of (9.1), one estimates $\psi _1\sim \bar {d}_e^2\psi _1''$. For the small-$\varDelta '$ limit, one has $\psi _1''/\psi _1 \sim \varDelta '/\delta \sim 1/\bar {d}_e^2$, meaning $\delta \sim \varDelta ' \bar {d}_e^2$. Therefore, for the $\varDelta ' \delta \ll 1$ limit, one obtains

(9.14)\begin{equation} \gamma \sim J_0 k \varDelta' \rho_s \bar{d}_e , \quad \delta \sim \varDelta' \bar{d}_e^2. \end{equation}

These scaling laws for $\gamma$ and $\delta$ are identical to those given by (7.65) which we analytically derived in § 7.5. That is, they allow us to recover the small-$\varDelta '$ limit of the scalings of (Pegoraro & Schep Reference Pegoraro and Schep1986).

Following the same line of thought to find the scaling laws of the large-$\varDelta '$ limit, one would expect, as discussed in § 9, that the largest scale $l_c$ equals $\delta$ since this time, $\delta \gg (\varDelta ')^{-1}$. Therefore, $\psi _1\sim \bar {d}_e^2 \psi _1''$ gives $\delta \sim \bar {d}_e$, which differs from the scaling in (7.53). Proceeding with this argument, after substituting $\delta \sim \bar {d}_e$ in (9.13), the scaling $\gamma \sim k \rho _s$ is obtained. This also differs from the scaling law obtained analytically in (7.54) and which, instead, has been numerically verified (Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020). No numerical evidence in the range $\rho _s\gtrsim \bar {d}_e$ has been found of the scalings $\delta \sim \bar {d}_e$ and $\gamma \sim k \rho _s$. Also note that $\gamma \sim k \rho _s$ does not display any explicit dependence on $\bar {d}_e$, which contains the non-ideal parameters that allow magnetic reconnection (and which should make $\gamma \to 0$ as $\bar {d}_e\to 0$). We therefore conclude the scalings $\delta \sim \bar {d}_e$ and $\gamma \sim k \rho _s$ to be wrong.

This implies that the generalisability of the heuristic approach to the warm regimes is not evident and further information about the estimates of the relevant quantities is needed.

Even if this problem is not solved yet, and a closed set of equations for the heuristic estimates is not available when $\rho _s\gtrsim \bar {d}_e$, in the next subsection, we investigate this possibility by introducing a new characteristic scale-length of the system associated with the gradient of the velocity component parallel to the neutral line, and which we postulate to be related to the gradient of the magnetic flux function at the boundaries of the ‘outer region’: although so far we must rely on its numerical estimate, we are going to show that this allows us to use a heuristic-type approach to get the correct scaling laws of the growth rate and of the inner layer width in all reconnection regimes here considered. In this sense, introducing this scale length at least allows us to generalise the heuristic procedure to warm-electron regimes, although this generalisation remains so far incomplete, due to the lack of a procedure apt to a priori estimate this scale length. Moreover, the asymptotic scaling of this quantity can be shown to correspond to that of a characteristic scale length, which in both Porcelli (Reference Porcelli1991) and Zocco & Schekochihin (Reference Zocco and Schekochihin2011) has been obtained as a normalisation length in boundary layer integration procedure. In these works, it had been related to the width $>\delta _1$ of the domain sub-interval in which the solution of the innermost equation is valid in the large-$\varDelta '$ limit.

10 An ansatz about the ‘generalisation’ of the heuristic estimates: the role of the velocity gradient in the non-ideal region

A critical ingredient of the previous analysis is the estimate of the characteristic length, $l_c$, related to the first derivative of the magnetic stream function in the non-ideal region. Let us focus on the $\rho _s\gtrsim \bar {d}_e$ regime, where heuristic estimates display problems. Using $\psi _1 \sim \bar {d}_e^2 \psi _1''$ and (9.13), we see that the scale length $l_c$ enters in the estimates of the growth rate and of the reconnecting layer width as

(10.1)\begin{equation} \gamma \sim J_0 \frac{k \rho_s \bar{d}_e}{ l_c}, \quad \delta \sim \frac{\bar{d}_e^2}{l_c}. \end{equation}

In both the collisionless and resistive regimes, the correct estimates are recovered in the small-$\varDelta '$ limit when $l_c\sim (\varDelta ')^{-1}$, whereas they are incorrect in the large-$\varDelta '$ limit if we assume $l_c\sim \delta$. This suggests looking for another reasonable estimate for $l_c$, in this case.

The strategy we pursue here is therefore to look for a third ‘effective’ scale length for $l_c$, different from $\varDelta '$ and $\delta$, that would allow us to recover the correct scaling from (10.1) in the large-$\varDelta '$ limit, and to verify its relevance and appropriateness by means of numerical calculations.

10.1 Velocity gradient in the innermost non-ideal region

The likely candidate we propose for an alternative definition of $l_c$ is the inverse of the logarithmic jump in the component of the derivative of the fluid velocity parallel to the neutral line and evaluated at $x=\pm \delta$, which we name $\varDelta '_{v_y}$, in (loose) analogy with the usual $\varDelta '$ defined for the magnetic stream function $\psi _1$:

(10.2)\begin{equation} \varDelta'_{v_y} \equiv \frac{v'_{y}(\delta) - v'_{y}(-\delta)}{v_{y}(\delta)} = \frac{\varphi''_{1}(\delta) - \varphi''_{1}(-\delta)}{\varphi_{1}'(\delta)}. \end{equation}

It must be noted that in (10.2), we have used the ‘whole’ eigenfunction $\varphi _1$, differently from what happens in the definition of $\varDelta '$, in which only the ‘outer’ eigenfunction $\psi _{\text {out}}$ is involved. This fact is important for the numerical computation of both $\varDelta _{v_y}'$ and $\varDelta '$, as it will be discussed in § 10.3. Definition (10.2) and the identification $\delta =\delta _1$ means that $1/\varDelta '_{v_y}$ represents the characteristic scale length of the velocity gradient at the boundary of the innermost layer, i.e. the ‘electron diffusion region’.

Evaluating (9.2) at $x=\delta$ and $x=-\delta$, and using the definition (10.2), we obtain

(10.3)\begin{equation} \gamma \varDelta'_{v_y} \varphi_1'|_{x=\delta} = {\rm i}k \delta J_0\,(\psi_1''|_{x=\delta}+\psi_1''|_{x={-}\delta}). \end{equation}

Using the fact that $\psi _1''(\delta )=\psi _1''(-\delta )$ and assuming the validity of condition (9.3) at $x=\delta$, we find

(10.4)\begin{equation} \left.\frac{\gamma \varphi_1'}{k \psi_1}\right|_{x=\delta} = \frac{2 {\rm i} J_0}{l_c \varDelta_{v_y}'}. \end{equation}

Equation (10.4) expresses a constraint on the product $l_c\varDelta _{v_y}'$ which depends on the scaling of $\gamma$, and on the profiles of $v_y=\varphi _1'$ and of $\psi _1$. It should be noted that in the cold-electron regimes, the correct scalings can be obtained via heuristic approach using the hypothesis $\varphi _1''|_{x\simeq \delta }\sim \varphi _1/\delta _1^2$ (see, e.g. Biskamp Reference Biskamp2000, § 4.1.1). Should this assumption be always correct, the scaling $l_c\sim (\varDelta _{v_y})^{-1}\sim \delta _1$ would be always obtained in the large-$\varDelta '$ limit regardless of the reconnection regime, but this estimate does not allow us to recover the correct scalings when $\rho _s\gtrsim \bar {d}_e$. However, although so far the asymptotic scaling of $\varDelta _{v_y}'$ is not known, considering the results that we have found to be valid in the cold-electron regimes, we can expect that, at least for $\rho _s=0$ and in the large-$\varDelta '$ limit, the quantity $\varDelta _{v_y}'$ be related to (actually, ‘be proportional to the inverse of’) $\delta _1$.

10.2 A heuristic generalisation of the definition of the scale $l_c$

Based on the remarks above, we heuristically postulate the definition:

(10.5)\begin{equation} l_c \sim\max\{(\varDelta')^{{-}1}, (\varDelta_{v_y}')^{{-}1}\}, \end{equation}

which we will later show (§ 10.4) to be indeed consistent with all other definitions and hypotheses, and thus, arguably correct. It can be a posteriori verified that the results would have been equally consistent even if we had developed the arguments which follow using the alternative definition $l_c \sim \max \{(D')^{-1}, (\varDelta _{v_y}')^{-1}\}$, based on the further inverse scale length $D'$, which we introduce below, in (10.10), and whose scaling we numerically compute later, in different regimes. Definition (10.5) has been chosen, here, since $\varDelta '$ is always a priori known, whereas the evaluation of $D'$, too, requires a numerical integration of the boundary layer problem. According to heuristic definition (10.5), in the small-$\varDelta '$ limit, we expect $l_c\sim (\varDelta ')^{-1}\gg (\varDelta _{v_y}')^{-1}$, whereas in the large-$\varDelta '$ limit, we expect $l_c\sim (\varDelta _{v_y}')^{-1}\gg \varDelta '$. From (10.5), the transition between the two limits is therefore ruled by the asymptotic scaling of the ratio $\gamma \varphi _1'/(k \psi _1)$. In particular, regardless of the reconnection regime (resistive or collisionless, warm or cold), we must have

(10.6)\begin{equation} \left.\frac{\gamma \varphi_1'}{k \psi_1}\right|_{x=\delta} \sim \frac{\varDelta'}{\varDelta_{v_y}}\ll 1 \quad(\text{for}\ \varDelta'\delta\ll 1), \quad \left.\frac{\gamma \varphi_1'}{k \psi_1}\right|_{x=\delta} \sim O(1) \quad(\text{for}\ \varDelta'\delta\gg 1). \end{equation}

The first of the conditions in (10.6) can be however a posteriori refined and re-formulated as a condition on the ratio $(\gamma \varphi _1'')/(k \psi _1)$, which reads

(10.7)\begin{equation} \left.\frac{\gamma \varphi_1''}{k \psi_1}\right|_{x=\delta} \sim O(1) \quad(\text{for}\ \varDelta'\delta\ll 1). \end{equation}

This condition follows from (9.2) combined with the second part of (9.3) and from the knowledge we have about the scalings of $\delta$ and $\gamma$ in terms of $l_c$: using the estimates $l_c\delta \sim d_e^2$ and $l_c\delta \sim S^{-1} /\gamma$ that we obtain by specialising $\bar {d}_e$ to the collisionless and resistive regimes, respectively, we obtain

(10.8)\begin{equation} \left.\frac{\gamma \varphi_1''}{k \psi_1}\right|_{x=\delta} \sim \frac{\delta}{d_e^2} \quad(\text{collisionless}), \quad \left.\frac{\gamma \varphi_1''}{k \psi_1}\right|_{x=\delta} \sim \frac{ \gamma \delta}{S^{{-}1}} \quad(\text{resistive}). \end{equation}

Equation (10.9) is then verified in all reconnection regimes once we substitute the relevant known scalings we have already evaluated from boundary layer theory into (10.8). Also, using the definition of $\varDelta _{v_y}'$ of (10.2), the two equations in (10.8) result to be compatible with the second part of (10.6). In conclusion, we can therefore write the constraints:

(10.9)\begin{equation} \left\{ \begin{array}{l} \displaystyle{ \left.\dfrac{\gamma \varphi_1''}{k \psi_1}\right|_{x=\delta} \sim O(1)} \\ \displaystyle{ \left.\dfrac{\gamma \varphi_1'}{k \psi_1}\right|_{x=\delta} \sim \dfrac{\varDelta'}{\varDelta_{v_y}}\ll 1} \end{array}\right.\quad (\text{for}\ \varDelta'\delta\ll 1),\quad \left.\frac{\gamma \varphi_1'}{k \psi_1}\right|_{x=\delta} \sim O(1) \quad(\text{for}\ \varDelta'\delta\gg 1). \end{equation}

These conditions can be taken to be generally discriminating for the transition from the small- to the large-$\varDelta '$ scaling relations in any reconnection regimes.

It must be however emphasised that, while both (10.6) and (10.8) are self-consistently deduced from the specific hypotheses we have made so far, (10.7) must be heuristically assumed, since, although a priori reasonable and compatible with the first of conditions (10.6), it does not follow from the other hypotheses, but it is just verified once the scalings of $\gamma$ and $\delta$ are found.

The discussion and the analysis we are going to develop next (§ 10.4) seems to suggest that accomplishing this latter task in the context of the heuristic derivation may not be feasible. Nevertheless, we will numerically prove the correctness of the hypotheses (i)–(iii) of § 9 and we will show that definitions (10.4) and (10.5) are consistent with the estimates of the correct scalings, provided the scaling of $\varDelta _{v_y}'$ is known. In doing so, we will elucidate (cf. Appendix E) the logical points of the heuristic approach in the different tearing regimes by pointing out when an a priori self-consistent estimate can be done or not, and, in the second case, which information is missing. Before doing so, we need however to first discuss how to numerically evaluate $\varDelta _{v_y}'$, which must be compared with the numerical value of $\varDelta '$ and of $\delta$, the latter of which has been already discussed in § 8. This is what we are going to do in § 10.3.

10.3 Numerical evaluation of $\varDelta '$ and $\varDelta _{v_y}'$

Let us first look at a numerical procedure that allows us to quantify the inverse scale lengths $\varDelta '$ and $\varDelta _{v_y}'$ defined by (3.8) and (10.2), respectively.

Of course, because of the definition in (3.8), the values of $\varDelta '$ are always independent of the non-ideal parameters. The numerical evaluation of $\varDelta '(k)$ becomes therefore trivial whenever an analytical formula that depends only on $k$ and on the equilibrium profile can be obtained (cf. (4.13)).

Then, for the evaluation of $\varDelta _{v_y}'$, the definition in (10.2) may be operationally used, although it is procedurally quite demanding: it requires to compute first the eigenfunction $\varphi _1$, then its first and second derivative, and then to evaluate them at $x=\delta$, a value which can be numerically obtained by using the definition in (8.3) and by following the procedure sketched in figure 12.

It is however possible to ‘speed up’ the calculation of the scalings of both $\varDelta '$ and $\varDelta _{v_y}$, by relying on an alternative numerical procedure.

This procedure mimics the numerical evaluation of $\varDelta '$ that is made possible only in the small- $\varDelta '$ limit, thanks to the geometrical interpretation that can be given of the instability parameter in terms of a local expansion of the outer solution (Furth et al. Reference Furth, Killeen and Rosenbluth1963): we recall that by using $\psi _{\text {out}} \approx {c}_0 + {c}_1 |x|$ as $|x| \to 0$, then $\varDelta ' = 2 {{c}_1}/{{c}_0}$ (cf. §§ 5.3.1 and 5.3.2). These two coefficients can be evaluated by measuring the value of $\psi _{\text {out}}$ and of the slope of the tangent to $\psi _{\text {out}}$ close to $x=0$.

This idea can be borrowed so as to evaluate analogous quantities defined with respect to the total solutions $\psi _1$ and $\varphi _1$: noting that, graphically speaking, both $\psi _1$ and $\varphi _1'$ still display a linear behaviour with respect to $x$ as $x\to \delta$, both in the small- and large-$\varDelta '$ limits, we write $\psi _1|_{x\to \delta } \approx {c}_0 + {c}_1 |x|$ and $v_{y}|_{x\to \delta } = \varphi '_{1}|_{x\to \delta } \approx q_0 + q_1 |x|$. Note that these approximations are consistent with the local expansions (8.5) and (8.6), which are valid, instead, for $x\ll \delta$. Therefore, once the corresponding eigenfunctions have been numerically computed, at $|x|\sim \delta$, we can evaluate (see figure 14)

(10.10)\begin{equation} D' \equiv 2 \frac{c_1}{c_0}, \quad \varDelta'_{v_y} = 2 \frac{q_1}{q_0}. \end{equation}

In particular, due to the smallness of $\delta$, the coefficients $c_1, c_0$ and $q_0, q_1$ can be numerically computed as shown in figure 14, by measuring the values of $\psi _1$ and $\varphi _1$ and of the peak values of their derivatives close to $x=0$.

We have verified that the scalings obtained for $\varDelta _{v_y}'$ in this way agree with those directly computed by first evaluating $\delta$ and by then using the definition in (10.2), as shown in figure 14(b,c), and that $D'=\varDelta '$ in the small-$\varDelta '$ limit.

Figure 15 displays the scaling laws of $(D')^{-1}$, $(\varDelta '_{v_y})^{-1}$ and $\delta$, numerically computed according to the operational definitions given by (10.10) and (8.3), respectively, in the cold-collisionless regime at $\rho _s=0$: the different characteristic lengths are shown in figure 15(a) for the large-$\varDelta '$ limit, and in figure 15(b) for the small-$\varDelta '$ limit. Figure 16 shows the corresponding scaling laws for the tearing modes in the warm-collisionless regimes at $\rho _s\neq 0$. The dependence of $(D')^{-1}$, $(\varDelta '_{v_y})^{-1}$, and $\delta$ on the electron skin depth ($d_e$) is in figure 16(a,c), the dependence on the ion-sound Larmor radius ($\rho _s$) is in figure 16(b,d); both the large-$\varDelta '$ limit (figure 16(a,b)) and the small-$\varDelta '$ limit (figure 16(c,d)) are considered.

Figure 15. Examples of scaling laws of $D'$, $\varDelta '_{v_y}$ (cf. (10.10)) and $\delta$ with respect to $d_e$, for $\rho _s= 0$. The large-$\varDelta '$ limit is in panel (a), the small-$\varDelta '$ in panel (b). The values of $D'$ and $\varDelta '_{v_y}$ are computed numerically according to the procedure sketched in figure 14. In the small-$\varDelta '$ limit, the values of $\varDelta '$, obtained from (4.13), coincide with the orange line in panel (b).

Figure 16. Examples of scaling laws of $D'$, $\varDelta '_{v_y}$ (cf. (10.10)) and $\delta$ for $\rho _s\neq 0$. Panels (a,b) correspond to the large-$\varDelta '$ limit, whereas panels (c,d) correspond to the small-$\varDelta '$ limit. The values of $D'$ and $\varDelta '_{v_y}$ are computed numerically according to the procedure sketched in figure 14. Notice that $D'$ and $\delta$ are independent of $\rho _s$ in the small-$\varDelta '$ limit (cf. panel d). In the small-$\varDelta '$ limit, the values of $\varDelta '$ obtained from (4.13) coincide with the blue lines in panels(c,d).

Numerical results prove that, as it could be expected, $D'$ of (10.10) coincides with the definition of $\varDelta '$ only in the small-$\varDelta '$ limit, where we can indeed state that $\psi _1(x)|_{x\to \delta ^+ }\simeq \psi _{\text {out}}(x)|_{x\to \delta ^+ }$ (cf. figure 15b, and figure 16c,d). In this limit, $D'$ can be therefore taken as an accurate estimate of $\varDelta '$, in spite of the fact that the latter is formally defined by evaluating the derivatives of $\psi _{\text {out}}$ at a distance from the neutral line much larger than $\delta$. This is made possible by the fact that in this large wavelength limit, the outer solution must match the inner one in an overlapping region that gets sufficiently close to $|x|=\delta$. That is, the constant-$\psi$ hypothesis holds in the whole non-ideal region, down to the innermost layer.

Different is the result in the large-$\varDelta '$ limit: in the cold-collisionless regime, $(D')^{-1}$ displays the same asymptotic scaling of $\delta$, except for a numerical factor (cf. figure 15a), and these scaling laws are the same as for $(\varDelta _{v_y}')^{-1}$; in the warm-collisionless regime, instead, both $(D')^{-1}$ and $(\varDelta _{v_y}')^{-1}$ display the same asymptotic scalings $(D')^{-1}\sim (\varDelta _{v_y}')^{-1} \sim \rho _s^{{1}/{3}} d_e^{{2}/{3}}$ (figure 16a,b) which are non-trivial, since they differ from $\delta _{LD}^{-1} \sim \rho _s^{{1}/{3}} d_e^{{4}/{3}}$. Analogous results, not shown here, are found in the warm-resistive, large-$\varDelta '$ limit, in which, for these quantities, we obtain also an explicit dependence on $k$: $(D')^{-1}\sim (\varDelta _{v_y}')^{-1} \sim k^{-{2}/{7}}\rho _s^{-{1}/{7}} S^{-{2}/{7}}$ and $\delta \sim k^{-{4}/{7}}\rho _s^{-{5}/{7}} S^{-{4}/{7}}$.

For summary, we recall here the scalings of $\varDelta _{v_y}'$ that have been numerically obtained in the different regimes:

  • $\varDelta _{v_y}'\sim \delta _{SD}^{-1}$   (warm/cold and collisionless/resistive regimes at small-$\varDelta '$);

  • $\varDelta _{v_y}'\sim \delta _{LD}^{-1}$   (cold and collisionless regimes at large-$\varDelta '$);

  • $\varDelta _{v_y}'\sim \rho _s^{-{1}/{3}}d_e^{-{2}/{3}}$   (warm-collisionless regime at large-$\varDelta '$ );

  • $\varDelta _{v_y}'\sim k^{-{2}/{7}}\rho _s^{{1}/{7}}S^{-{2}/{7}}$    (warm-resistive regime at large-$\varDelta '$ ).

10.4 Role of $\varDelta _{v_y}'$ in heuristic-type estimates

It is easy to verify that combining the estimates in (10.1) with the definition in (10.5) and with the scalings numerically obtained for $\varDelta _{v_y}'$, the correct scaling laws obtained in §§ 6 and 7 can be recovered in both the warm-collisionless and warm-resistive regimes.

This is discussed in detail in Appendix F, in which the logical steps of the procedure are singled out and identified, thus providing insight on the physical interpretation of the analytical results of the boundary layer calculations. In particular, the main hypotheses on which this (partial) heuristic approach relies and the corresponding results in all collisionless regimes, both ‘warm’ and ‘cold’, are summarised in tables 3–5 of Appendix F, where the logical steps of the procedure are presented as statements and formulae.

There are two main results of this analysis.

  1. (i) Introducing the scale length $\varDelta _{v_y}'$ and postulating the definition in (10.5) make it possible to obtain the scaling laws, which are, in principle, correct in all regimes and wavelength limits; however, in the large-$\varDelta '$ warm-electron regime, the procedure results to be not ‘closed’, in the sense that the estimation of the scaling law of $\varDelta _{v_y}'$ seems to be not possible by simple dimensional analysis.

  2. (ii) Quite interestingly, using the definition in (10.5), the scaling laws of all reconnection regimes can be cast in a form which is perfectly symmetric in the small- and large-$\varDelta '$ limit, with respect to the substitution $\varDelta '\leftrightarrow \varDelta _{v_y}'$.

These results, on the one hand, suggest to us that introducing the inverse scale length $\varDelta _{v_y}'$ is a promising ingredient in the attempt to extend the heuristic-type analysis. On the other hand, however, finding the further constraint that makes it possible to obtain a priori all the sought algebraic scalings without resorting to numerical analysis and by mere dimensional analysis seems an elusive task. The logical steps identified in the tables of Appendix F, as well as some further insight on the interpretation of the boundary layer results, which follows from the heuristic approach and which we are going to discuss below (§ 10.5), could imply the heuristic analysis to be intrinsically not applicable in some regimes. This problem becomes manifest, in particular, in the large-$\varDelta '$ limit when $\rho _s\neq 0$, in which the scaling of $\varDelta _{v_y}'$ is a non-trivial power law combination of the scales $\rho _s$ and $\bar {d}_e$. Nevertheless, the coherence of the results summarised in tables 35 supports the consistency of the heuristic ‘ansätze’ we have made so far.

In particular, the logical steps identified in tables 35 show that the difference between the warm and cold regimes lies in the balance condition $\psi _1\sim i k x \varphi _1 /\gamma \varphi _1$, which is valid in the cold regimes only (hypothesis [T6.H2] of table 5), and which is replaced by the balance expressed by hypothesis [T5.H2] of table 4 when $\rho _s\neq 0$. Condition [T6.H2] expresses indeed the validity of (4.1), in turn related to the inverse scale $\varDelta '$, down to the boundary layer at the frontier with the innermost region: this means that the newly introduced scale $(\varDelta _{v_y}')^{-1}$ is redundant in this regime with respect to $\delta$ and $\varDelta '^{-1}$ (as, however, it was evident already from the heuristic-type approach discussed in § 9.3). We have indeed seen from the numerical results summarised at the end of § 10.3 that the scaling of $(\varDelta _{v_y}')^{-1}$ turns out to always coincide with that of $\delta$. Such a ‘closure’ condition for $\varDelta _{v_y}$ is lost at $\rho _s\neq 0$, when hypothesis [T6.H2] is replaced by hypothesis [T5.H2], which leads to the general constraint $\gamma \sim k\rho _s(\delta /l_c)^{1/2}$: here, a value different from that of $\delta$ or $(\varDelta ')^{-1}$, with respect to which the transition from the small- to the large-$\varDelta '$ limits is measured, is in principle admitted for $l_c$.

10.5 Significance of the inverse spatial scale $\varDelta _{v_y}'$: coherence with boundary layer calculations and comparison with previous work

Some insight about the physical significance of the characteristic spatial scales associated with $\varDelta _{v_y}'$ is obtained from comparison of their scalings with identical asymptotic scalings which can be obtained from some boundary layer results available in the previous literature (Pegoraro & Schep Reference Pegoraro and Schep1986; Cowley & Hastie Reference Cowley and Hastie1988; Porcelli Reference Porcelli1991; Zocco & Schekochihin Reference Zocco and Schekochihin2011) and from related works discussing the implications of these results (Ottaviani & Porcelli Reference Ottaviani and Porcelli1995; Grasso et al. Reference Grasso, Pegoraro, Porcelli and Califano1999).

Combination of (10.5) (i.e. hypothesis T4.H1) with hypotheses T5.H3 or T6.H3 allows the identification of the small- and large-$\varDelta '$ limits, say $\varDelta '\delta \gtrless 1$, expressed in each reconnection regime in terms of $l_c$ and thus in terms of $\varDelta _{v_y}'$:

(10.11)\begin{equation} \varDelta'\delta \gtrless 1 \quad \Leftrightarrow \quad \left\{ \begin{array}{@{}l} \displaystyle{ \varDelta' d_e \gtrless \max\left\{ \dfrac{1}{\varDelta'd_e}, \dfrac{1}{\varDelta_{v_y}'d_e} \right\} }\quad (\text{collisionless}),\\ \displaystyle{ \varDelta' S^{{-}1} \gtrless \max\left\{ \dfrac{\gamma}{\varDelta'}, \dfrac{\gamma}{\varDelta_{v_y}'} \right\} }\quad (\text{resistive}). \end{array} \right. \end{equation}

In this regard, it is interesting to compare, e.g. the conditions for the warm-collisionless case with the similar conditions in Grasso et al. (Reference Grasso, Pegoraro, Porcelli and Califano1999) that have been written (cf. (17) and paragraphs below (18) therein) as

(10.12)\begin{equation} \varDelta' d_e > \min\left\{ 1, \left(\frac{d_e}{\rho_s}\right)^{1/3} \right\} ,\quad \varDelta' d_e < 1, \end{equation}

based on the results of the boundary layer calculations in the Fourier representation by Porcelli (Reference Porcelli1991) (cf. conditions on ‘$\lambda _H\equiv -{\rm \pi} /\varDelta '$’ in between (7) and (9) therein). Substitution of the scalings numerically found for $\varDelta _{v_y}'$ in § 10.3 for $\rho _s^2\gg d_e^2$ into the collisionless condition of (10.11) gives

(10.13)\begin{align} \varDelta' d_e > \max\left\{ \frac{1}{\varDelta'd_e}, \left(\frac{\rho_s}{d_e}\right)^{1/3} \right\}, \qquad \varDelta' d_e < \max\left\{ \frac{1}{\varDelta'd_e}, \varDelta' d_e \right\},\\ \text{(large-$\varDelta'$)} \qquad\qquad\qquad\qquad\qquad \text{(small-$\varDelta'$)} \quad \nonumber \end{align}

which are indeed compatible with the conditions in (10.12), once $(\varDelta 'd_e)^2< 1$ is deduced from the second part of (10.13) in the small-$\varDelta '$ limit and therefore $\varDelta 'd_e < 1$ is assumed, with $\varDelta 'd_e \sim 1$ fixing the threshold value also for the first inequality of (10.13).

In this regard, we notice that the non-trivial characteristic scale length associated with the asymptotic scaling, which we have numerically found for $\varDelta _{v_y}'$ in the large-$\varDelta '$, warm-electron limit, naturally emerges from the boundary layer analysis developed by Pegoraro & Schep (Reference Pegoraro and Schep1986); Porcelli (Reference Porcelli1991): here, the general dispersion relation encompassing both the small- and large-$\varDelta '$ limits can be written in terms of $\bar {d}_e$ as (to this purpose, we can just substitute $d_e\to \bar {d}_e$, e.g. in (25) of Ottaviani & Porcelli Reference Ottaviani and Porcelli1995)

(10.14)\begin{equation} \frac{\rm \pi}{2}\gamma^2={-}\frac{{\rm \pi}\rho_s}{\varDelta'}+\frac{\rho_s^2\bar{d}_e}{\gamma}. \end{equation}

Naming $\gamma _{LD}\sim (2/{\rm \pi} )^{1/3}\bar {d}_e^{1/3}\rho _s^{2/3}$ the solution obtained in the $\varDelta '\to \infty$ limit, which we already recovered in previous sections, one sees that the opposite, small-$\varDelta '$ limit is obtained when the condition

(10.15)\begin{equation} \frac{{\rm \pi}\rho_s}{\varDelta'}\gg \frac{\rho_s^2\bar{d}_e}{\gamma_{LD}} \quad \Leftrightarrow\quad \varDelta' \ll \frac{1}{2\rho_s^{1/3}\bar{d}_e^{2/3}} \quad \implies \quad \varDelta' \ll \varDelta_{v_y}' \end{equation}

is satisfied, where in the last passage, we have used the numerical result we previously found in this wavelength limit, $\varDelta _{v_y}'\sim \rho _s^{-1/3}\bar {d}_e^{-2/3}$. The rightmost condition of (10.15) is consistent indeed with the constraints in (9.4), previously found via the heuristic estimates discussed in § 10.2. This suggests that $\varDelta _{v_y}'$ may provide a physical interpretation of the appearance of this characteristic scale length in boundary layer calculations, performed in the framework of the two-fluid model we consider here.

This interpretation is however somewhat different from that provided by Ottaviani & Porcelli (Reference Ottaviani and Porcelli1995), and which was based on the previous boundary layer calculations (Cowley et al. Reference Cowley, Kulsrud and Hahm1986; Pegoraro & Schep Reference Pegoraro and Schep1986; Porcelli Reference Porcelli1991) performed by starting from a charge density equation in which polarisation effects were taken into account through the gyrokinetic particle response. In that modelling framework, the scale-length $\rho _s^{1/3}\bar {d}_e^{2/3}\gg \delta _1$ was noted by Ottaviani & Porcelli (Reference Ottaviani and Porcelli1995) to correspond to the distance from the neutral line at which $\gamma _{LD}$ becomes comparable to the phase-velocity $k_{\|}(x)v_{{\rm th}}^{e}=({\boldsymbol k}\boldsymbol {\cdot }{\boldsymbol B}_{0}(x)/B_0) v_{{\rm th}}^{e}$, and below which the isothermal electron closure would formally break down. As already noted in the same work, however, the appropriateness of the isothermal condition, assumed to be valid for the purpose of the boundary layer calculations, had been numerically verified to a good extent in some previous works (Berk, Mahajan & Zhang Reference Berk, Mahajan and Zhang1991; Coppi & Detragiache Reference Coppi and Detragiache1992) and, more recently, its validity has been supported by the numerical studies of Perona, Eriksson & Grasso (Reference Perona, Eriksson and Grasso2010). Relating the scale $\rho _s^{1/3}\bar {d}_e^{2/3}$ to the failure of the isothermal closure is thus consistent with the interpretation provided in further – and in part preceding– works based on a kinetic approach (Drake & Lee Reference Drake and Lee1977; Cowley et al. Reference Cowley, Kulsrud and Hahm1986; Zocco & Schekochihin Reference Zocco and Schekochihin2011). In these works, such a critical distance was shown by heuristic arguments (see Drake & Lee Reference Drake and Lee1977; Cowley et al. Reference Cowley, Kulsrud and Hahm1986, § V) to correspond to the characteristic width of the current layer, determined by the balancing of the total current generated by the parallel electron pressure gradients with the current generated by the (reconnecting) parallel electric field. Instead, in § 8 and in Betar et al. (Reference Betar, Del Sarto, Ottaviani and Ghizzo2020), we have numerically proven the current sheet to be concentrated around the neutral line in a region of width $\delta _1$. The subtle point here may be in the meaning which can be given to the notion of ‘characteristic width of the current layer’, since it is true that, while the inflection points around the peak of $J_{z,1}$ are numerically found to be located at a distance $\delta _1$ from the neutral line (cf. definition in (8.3)), the profile of the inner solution (i.e. the solution found in the ‘electron region’) extends beyond this distance. In particular, the matching with the outermost non-ideal solution (i.e. the solution in the ‘ion region’) is not to be meant as a matching in a single point, but rather as an asymptotic matching valid over an intermediate layer, whose distance from the neutral line in the warm-electron regime can be of the order of $\rho _s^{1/3}\bar {d}_e^{2/3}$. This latter point of view is in agreement with the notion of ‘inner solution width’, to which Zocco & Schekochihin (Reference Zocco and Schekochihin2011) make reference: the characteristic width of the inner solution, which is named $\delta _{\text {in}}$ in the notation of their work, differs from the width of the inner layer, which we have here named $\delta _1=\delta$, and satisfies $\delta _1 < \delta _{\text {in}}<\delta _2$.

If we look in detail at the boundary layer analysis carried out by these latter authors in the warm-electron regime, we see that the scalings they a posteriori obtained (cf. (B56) and (B101) in the ‘Collisionless two fluid limit’ and ‘Resistive two fluid limit’ of that work) for the normalisation scale $\delta _{\text {in}}$, coincide in the large-$\varDelta '$ limit with the scalings of $(\varDelta _{v_y}')^{-1}$ that we have detailed at the end of § 10.3. In particular, for the regime that we can write as $\rho _s\gtrsim \bar {d}_e$, these authors identified the small- and large-$\varDelta '$ conditions, which we have here generally expressed as $\varDelta '\delta _1 \ll 1$ and $\varDelta '\delta _1\gg 1$, respectively, via the conditions $\varDelta '\delta _{\text {in}} \ll 1$ and $\varDelta '\delta _{\text {in}}\gg 1$, instead. Using the correspondence $\delta _{\text {in}}\to (\varDelta _{v_y}')^{-1}$, these conditions would map into $\varDelta ' \ll \varDelta _{v_y}'$ and $\varDelta '\gg \varDelta _{v_y}'$, the former of which is consistent with (9.4). It should be however noted that the scale $\delta _{\text {in}}$ appears in the boundary layer analysis of Zocco & Schekochihin (Reference Zocco and Schekochihin2011) as a consequence of a normalisation choice of the non-ideal equations of their model ((B35)–(B36) therein), which differs with respect to the one we have detailed in § 7: while the scalings of their outermost and innermost non-ideal layers are a posteriori found to coincide with those of the scales $\delta _2$ and $\delta _1$ that we introduced in §§ 7.2.1 and 7.2.2, these authors chose instead $\delta _{\text {in}}\equiv (\sqrt {2}\rho _s\delta _1)^{1/2}$ as a normalisation scale of the innermost equations (our $\delta _1$ mapping into $\delta$ of (B28) of their work), and they subsequently ordered the terms of the tearing equations with respect to this spatial scale. The scalings of the relevant quantities have been thus obtained there, via some heuristic ansatz on the width of the integral at the left-hand side of the equivalent of (3.7), in the form in which they obtained it (cf. (B46), therein). To evaluate the integral, which depends on $\varphi _{1}''$ (cf. (3.10)), the terms in the corresponding auxiliary equation ((B42) therein), which are related to integrand via $\varphi _1''=\chi '$, have been so ordered in relation to the characteristic width $\delta _{\text {in}}$. This approach is at the basis of the ordering of the wavelength regime in terms of the product $\varDelta '\delta _{\text {in}}$, which differs from the ordering in terms of the product $\varDelta '\delta _1$, which we have here adopted, instead. The coherence of the results obtained in the two approaches, also emphasises the margin of arbitrariness in the choice of the normalisation scale, with respect to which it is possible to define the width of the boundary layers and to perform the integration and matching, after some appropriate approximations of the terms in the equations are assumed on heuristic basis. This makes possible the interpretation of $\delta _{\text {in}}$ as the width of the solution in the innermost equation, information which is not evident if one follows instead the normalisation procedure we have adopted in this work. At the same time, it should be noticed that the scale $(\varDelta _{v_y}')^{-1}$ seems to be not generally identifiable as the $\delta _{\text {in}}$ obtained by Zocco & Schekochihin (Reference Zocco and Schekochihin2011): although the correspondence between the scalings of $\delta _{\text {in}}$ obtained by these authors and the scaling of $(\varDelta _{v_y}')^{-1}$ that we have obtained holds in the large-$\varDelta '$ limit, it fails in the small-$\varDelta '$ limits (cf. their (B54) and (B97) with the scalings at the end of § 10.3).

In summary, while on the one hand, we note the agreement of the results obtained by solving the boundary layer equations in the different models and with integration techniques that rely on slightly different heuristic hypotheses, on the other hand, we note that the appearance of the characteristic scale, which, in previous works, has been interpreted in terms of inherently kinetic features, in the MHD model of § 2, can be instead entirely related to ‘fluid-like’ features associated with the gradients of the velocity field in the non-ideal region. It should be also noticed, in this regard, the care with which conclusions drawn by heuristic estimates based on dimensional analysis must be dealt with: although it is comforting that different models of tearing mode analysis yield the same quantitative results, their physical interpretation is a more delicate issue, which requires further insight based on the consistency of the specific hypothesis of each model and therefore may not be univocal. In particular, the symmetry between the scaling laws in the small- and large-$\varDelta '$ limits with respect to the substitution $\varDelta '\leftrightarrow \varDelta _{v_y}'$, which we have detailed in tables 35, suggests that the quantity $\varDelta _{v_y}'$, in this modelling related to inherently ‘fluid’ features, play a general, important role in the tearing mode regimes, which may deserve further investigations.

11 Summary and conclusions

We have reviewed the solution of the boundary layer problem for collisionless and resistive tearing instabilities in slab geometry, in both the small- and large-$\varDelta '$ limits. The calculations in the warm regime, in which two matching regions are required, have been solved in the coordinate space by using the integral representation of hypergeometric functions to integrate the differential equations of the boundary layer approach (§§ 6 and 7). To the best of our knowledge, this kind of analysis has not been presented before, elsewhere, and in the present work, emphasis is put on a pedagogical derivation of the results. In this way, we have recovered the results first obtained by Pegoraro & Schep (Reference Pegoraro and Schep1986), Pegoraro et al. (Reference Pegoraro, Porcelli and Schep1989) and Porcelli (Reference Porcelli1991) in a Fourier representation. While developing this analysis, we have also been able to make a direct comparison (§ 6) with calculations in the coordinate space that had been earlier performed in the cold-collisionless regime and in the purely resistive regime of tearing modes, where a single matching region is required (Furth et al. Reference Furth, Killeen and Rosenbluth1963; Coppi Reference Coppi1964c; Ara et al. Reference Ara, Basu, Coppi, Laval, Rosenbluth and Waddell1978), and with other calculations in the coordinate space that had been carried out in the warm-collisionless regime in the presence of warm ions (Zocco & Schekochihin Reference Zocco and Schekochihin2011).

Then, by making reference to the results of the boundary layer calculations, we have been able to relate the inverse of the derivatives of the eigenfunctions evaluated on the neutral line to specific scalings with respect to the non-ideal parameters, which had not been noted before (§ 8.3). We have also shown the relation of the inverse of these derivatives with the reconnecting layer width, $\delta$, whose operational definition, which we had previously verified in different reconnection regimes (Betar et al. Reference Betar, Del Sarto, Ottaviani and Ghizzo2020), we have here discussed for the first time (§ 8). These characteristic length scales, which can be useful for numerical diagnostics, are summarised in table 5.

We have interpreted in § 9 the results of the boundary layer analysis in light of heuristic derivations for the scalings of the growth rate and for the characteristic width of the reconnection layer, by following a dimensional analysis procedure that had been already successfully used in previous works but only when $\rho _s=0$. In this way, we have highlighted in § 9.3 how the heuristic approach alone fails to provide the correct scaling when the ion-sound Larmor radius is not negligible ($\rho _s\gtrsim \bar {d}_e$).

Then, thanks to the operational definition we have given of $\delta$, and by relying on both the heuristic estimates and the numerical solutions of the eigenvalue problem, we have shown a further non-trivial relation between the first derivative of $\psi _1$, evaluated close to the neutral line, and the gradients of the velocity component parallel to it (§ 10). We have, in this way, introduced an inverse characteristic scale length which we have named $\varDelta _{v_y}'$, because of its analogy with the classical $\varDelta '$ parameter. Using this (inverse) scale length, we have therefore shown, both analytically and numerically, that, for the purpose of heuristic estimates, we can generally assume $\psi _1'|_{x=\delta }\sim \psi _1|_{x=\delta }/l_c$ with $l_c=\max \{(\varDelta ')^{-1}, (\varDelta _{v_y}')^{-1}\}\gtrsim,\delta$. Knowing the asymptotic scaling of $\varDelta _{v_y}'$, an estimate of both $\delta$ and $\gamma$ can be made in any RMHD reconnection regime (§ 10.4) by just using dimensional analysis.

It is interesting to note that, from an experimental point of view, density fluctuations $n_1$ are easier to be measured than magnetic perturbations $\psi _1$, and that the former can be related to the fluid stream function perturbation via $n_1\sim \nabla ^2\varphi$. In general, then, the estimate of the spatial gradients of $\varphi _1$ from experimentally measured profiles of the density may be more reliable than the evaluation of the spatial gradients of $\psi _1$. In this context, the characterisation of the large- and small-$\varDelta '$ limits we have provided in § 10 with (10.6), or, more generally, the relation between the value of $\varDelta _{v_y}'=(\varphi _1''/\varphi _1')|_{x=\delta }\sim \varDelta _{v_y}'/2$ (cf. definition in (10.2)) and the scaling of $\delta$ or of the other scales detailed at the end of § 10.3, may be of interest.

Finally, we note that the introduction of the inverse scale $\varDelta _{v_y}'$ makes the scaling laws in the large-$\varDelta '$ limit mirror those in the small-$\varDelta '$ limit provided the substitution $\varDelta '\leftrightarrow \varDelta _{v_y}'$.

Acknowledgements

This article was written in honour of F. Pegoraro's 70th birthday. It is our pleasure to recall his fundamental contributions to the linear and nonlinear theory of reconnecting instabilities in plasmas. The authors thank Miho Janvier (IAS CNRS – Université Paris-Saclay, France) and Chiara Marchetto (ISC-CNR Milano, Italy) for useful discussions about some bibliographical references. The authors are grateful to the ‘Maison de la simulation Lorraine’ for partial time allocation on the cluster Explor (Project No. 2019M4XXX0978).

Editor Alex Schekochihin thanks the referees for their advice in evaluating this article.

Funding

This work was carried out within the framework of the French Federation for Magnetic Fusion Studies (FR-FCM) and of the Eurofusion consortium and received funding from the Euratom research and training programme 2014–2018 and 2019–2020 under grant agreement No. 633053 (WPEDU fundings obtained through FR-FCM AAP 2017–2021 ‘Evolution of current sheets in low-collision plasmas’, in particular, are gratefully acknowledged). The views and opinions expressed herein do not necessarily reflect those of the European Commission. Neither the European Union nor the European Commission can be held responsible for them.

Declaration of interests

The authors report no conflict of interest.

Appendix A. Derivation of the model equations for tearing modes in slab RMHD

Different derivations exist in the literature of the set of tearing equations we have considered in this paper. For the inclusion of finite ion-sound Larmor radius effects in a two-field model under the strong (infinite) guide field hypothesis, we direct the reader for example to Schep et al. (Reference Schep, Pegoraro and Kuvshinov1994), Kuvshinov, Pegoraro & Schep (Reference Kuvshinov, Pegoraro and Schep1994), Bergmans (Reference Bergmans2001) and Del Sarto et al. (Reference Del Sarto, Califano and Pegoraro2006). Here below, we repropose however their complete derivation, with more details than in previous articles.

Despite the different procedures proposed over the years to obtain (2.1) and (2.2), it is agreed that their nonlinear form reads

(A1)\begin{gather} \frac{\partial}{\partial t}(\psi-d_e^2\nabla^2\psi)+[\varphi,\psi-d_e^2\nabla^2\psi]=\rho_s^2[\nabla^2\varphi,\psi]+S^{{-}1}\nabla^2\psi, \end{gather}
(A2)\begin{gather} \frac{\partial}{\partial t}(\nabla^2\varphi)+[\varphi,\nabla^2\varphi]\,=\,[\psi,\nabla^2\psi]. \end{gather}

The above equations have been written using the standard ‘Poisson's bracket’ notation, $[f(x,y,t),g(x,y,t)]\equiv \partial _xf\partial _yg-\partial _yf\partial _xg$. Each bracket term can be thus related to a convection term associated with one of the two scalar function involved, e.g.

(A3)\begin{equation} [f,g]= ({\boldsymbol\nabla}f\times{\boldsymbol\nabla}g)\boldsymbol{\cdot}{\boldsymbol e}_z= \underbrace{({\boldsymbol e}_z\times{\boldsymbol\nabla}f)}_{{\boldsymbol u}_f}\boldsymbol{\cdot}{\boldsymbol\nabla}g. \end{equation}

Also note that in (2.1), the equilibrium contribution has been removed from the $S^{-1}\nabla ^2\psi$ term for the sake of linear analysis, since, in the asymptotic limit $S^{-1}\ll 1$, the time scale of resistive dissipation of the magnetic equilibrium is a posteriori found to be much longer than that of the tearing-type instability.

The appropriateness of (A1) and (A2), and notably of the $\rho _s^2$-related contribution, is supported by comparison of the linear dispersion relation obtained in the collisionless regime ($\eta =\nu =0$) with that obtained from the full Vlasov–Maxwell system in the same geometry configuration, that is, ${\boldsymbol B}^0={\boldsymbol B}_\perp ^0+B_z^0{\boldsymbol e}_z$ and ${\boldsymbol k}={\boldsymbol k}_\perp +k_z{\boldsymbol e}_z$: the dispersion relation of shear kinetic-Alfvén waves (see e.g. Hasegawa & Uberoi Reference Hasegawa and Uberoi1982, pp. 19–21) is indeed recovered,

(A4)\begin{equation} \omega^2=c_{A,\perp}k_\perp \sqrt{\frac{1+k^2\rho_s^2}{1+k^2d_e^2}}, \end{equation}

where $c_{A,\perp }$ is the Alfvén velocity evaluated with the perpendicular magnetic component, only. Also note that (A1) and (A2) can be combined to give an equation for the energy conservation in the form

(A5)\begin{align} \frac{\partial}{\partial t}\int \int \left( \underbrace{|{\boldsymbol \nabla}\psi|^2}_{\mathcal{E}_B} +\underbrace{d_e^2|{\nabla}^2\psi|^2}_{\mathcal{E}_J} +\underbrace{|\nabla\varphi|^2}_{\mathcal{E}_{\text{kin}}} +\underbrace{\rho_s^2|\nabla^2\varphi|^2}_{\mathcal{E}_{\text{int}}} \right)\,{{\rm d} x}\,{{\rm d} y}=2S^{{-}1}\int \int |{\nabla}^2\psi|^2\, {{\rm d} x}\,{{\rm d} y}, \end{align}

where $\mathcal {E}_B$, $\mathcal {E}_J$, $\mathcal {E}_{\text {kin}}$, $\mathcal {E}_{\text {int}}$ stand for the energy contributions respectively related to the in-plane magnetic field, to the current density (or electron kinetic energy), to the ion kinetic energy and to the internal energy (in turn related to the electron thermal energy and to their parallel compressibility along magnetic lines – see Grasso et al. Reference Grasso, Pegoraro, Porcelli and Califano1999).

The difference among the ‘different’ RMHD models which lead to the same set of (A1) and (A2) is in the weight that must be attributed to the different terms of the nonlinear equations with respect to the expansion parameters that have been adopted.

Here below, we focus on the first derivation proposed by Pegoraro & Schep (Reference Pegoraro and Schep1986) and Schep et al. (Reference Schep, Pegoraro and Kuvshinov1994); Kuvshinov et al. (Reference Kuvshinov, Pegoraro and Schep1994) and that has been later re-discussed by Bergmans (Reference Bergmans2001) and Del Sarto et al. (Reference Del Sarto, Califano and Pegoraro2006). In this framework, (A1) and (A2) can be shown to follow from the $z$-component of the electron momentum equation, which gives (A1), and from the charge continuity equation, which gives (A2), under the assumption that ${\boldsymbol \nabla }=(\partial _x,\partial _y,0)$ and that the gradient of the equilibrium density, and of its fluctuations as well, are smaller than the equilibrium quantity.

We start from the fluid equations for the species $\alpha$, and we assume the ion pressure to be negligible with respect to the electron temperature (i.e. we take the cold-ion limit). The continuity and momentum equations can be written in dimensionless form, normalised to the MHD scales, as

(A6)\begin{gather} \frac{\partial n^\alpha}{\partial t} +{\boldsymbol\nabla}\boldsymbol{\cdot}\left(n^\alpha{\boldsymbol u}^\alpha\right)=0, \end{gather}
(A7)\begin{gather} d_e^2 \left( \frac{\partial {\boldsymbol u}^e}{\partial t}+{\boldsymbol u}^e \boldsymbol{\cdot}{\boldsymbol\nabla}{\boldsymbol u}^e\right)={-}d_i\left({\boldsymbol E}+{\boldsymbol u}^e\times {\boldsymbol B} - \frac{\boldsymbol J}{S}\right)-\rho_s^2\frac{{\boldsymbol\nabla}\boldsymbol{\cdot}{\boldsymbol\varPi}^e}{n^e}, \end{gather}
(A8)\begin{gather} d_i^2 \left(\frac{\partial {\boldsymbol u}^i}{\partial t}+{\boldsymbol u}^i \boldsymbol{\cdot}{\boldsymbol\nabla}{\boldsymbol u}^i\right)= d_i\left({\boldsymbol E}+{\boldsymbol u}^i\times {\boldsymbol B} - \frac{\boldsymbol J}{S}\right). \end{gather}

Above, the ion skin depth $d_i=d_e\sqrt {m_i/m_e}$ has been further introduced and the pressure tensor ${\boldsymbol \varPi }^e$ has been normalised to the electron density times an electron reference temperature. This, after normalisation to the MHD scales, makes the squared ion-sound Larmor radius appear (normalised to the equilibrium shear length).

From now on, we will refer the symbol $\perp$ to the components that are orthogonal to the guide field direction (i.e. ${\boldsymbol e}_z$), that is, that lie on the $(x,y)$-plane. Note that this differs from another notation, frequently used for example in tokamak geometry, in which the direction ‘parallel’ and ‘perpendicular’ are referred to the total magnetic field. The latter choice is generally more appropriate for the most general RMHD modelling, which we recall has been first introduced by Strauss (Reference Strauss1976, Reference Strauss1977) in toroidal geometry and then in the so-called ‘cylindrical tokamak’ approximation, by ordering the ratio between the toroidal and the poloidal gradients to be comparable to the ratio between the poloidal and toroidal magnetic components. Our choice of notation is here justified by the assumption of translation invariance for the equivalent to the ‘toroidal’ component (i.e. $\partial _z=0$), corresponding to the large guide field component too (cf. figure 5).

One of the two equations we will focus on is the $z$-component of Ohm's law, which, for reasons we will show next, is more convenient to rewrite in the form:

(A9)\begin{equation} -E_z-\frac{d_e^2}{d_i}\frac{\partial u_z^e}{\partial t} -{\boldsymbol u}_\perp^e\boldsymbol{\cdot} \left[ {\boldsymbol B}\times {\boldsymbol e}_z + \frac{d_e^2}{d_i} {\boldsymbol \nabla}u_z^e\right] \,=\, -S^{{-}1}J_z + \frac{\rho_s^2}{d_i} \frac{{\boldsymbol\nabla}\boldsymbol{\cdot}{\boldsymbol\varPi}^e}{n^e}\boldsymbol{\cdot}{\boldsymbol e}_z. \end{equation}

The other one is the charge continuity equation (i.e. the sum over species of (A6) multiplied by $q^\alpha /n^\alpha$), which, using the quasi-neutrality condition $q^en^e+q^in^i=0$, we can rewrite as

(A10)\begin{equation} q^e{\boldsymbol u}_\perp^e\boldsymbol{\cdot} {\boldsymbol \nabla}(\ln n^e)+ q^i{\boldsymbol u}_\perp^i\boldsymbol{\cdot} {\boldsymbol \nabla}(\ln n^i)+ {\boldsymbol\nabla}\boldsymbol{\cdot}(q^e{\boldsymbol u}^e_\perp{+}q^i{\boldsymbol u}^i_\perp)=0. \end{equation}

The two equations above will be then specialised by approximating the fluid equations in terms the expansion parameters:

(A11)\begin{equation} \varepsilon_B\equiv\frac{|{\boldsymbol B}^0_\perp|}{B_z^0}, \quad \varepsilon_m\equiv\frac{m_e}{m_i}. \end{equation}

The electromagnetic field components are written in terms of the scalar quantities of interest as

(A12)\begin{equation} {\boldsymbol B}= \underbrace{B_z^0{\boldsymbol\nabla}\psi\times{\boldsymbol e}_z}_{{\boldsymbol B}_\perp} +\underbrace{B_z^0(1+b_1)}_{B_z^0+B_z^1}{\boldsymbol e}_z,\quad {\boldsymbol E}={-}\underbrace{{\boldsymbol \nabla}\phi}_{{\boldsymbol E}_\perp} -\underbrace{\frac{1}{c}\frac{\partial \psi}{\partial t}{\boldsymbol e}_z}_{E_z}. \end{equation}

Since we assume ${\boldsymbol J}=({c}/{4{\rm \pi} }){\boldsymbol \nabla }\times {\boldsymbol B}$, it follows ${\boldsymbol \nabla }\boldsymbol {\cdot }{\boldsymbol B}=0$ and

(A13)\begin{equation} {\boldsymbol J}_\perp{=}\frac{c}{4{\rm \pi}}{\boldsymbol\nabla}\times b{\boldsymbol e}_z, \quad J_z={-}\frac{c B_z^0}{4{\rm \pi}}{\nabla}^2\psi. \end{equation}

After normalisation to the MHD scales, the fluid stream function $\varphi =\phi /(cB_z^0)$ appears as related to the electrostatic potential. Because of the uniform guide field hypothesis, the scalar function $b$ also coincides with the first-order perturbation of the $B_z$ component.

We furthermore assume $|{\boldsymbol B}^1_\perp |\sim |{\boldsymbol B}^0_\perp |\sim \varepsilon _B B^0_z$, whence it follows

(A14)\begin{equation} \displaystyle{ \frac{\partial}{\partial t}\sim \frac{c_{A,\perp}}{L_0}\sim \varepsilon_B \frac{B_z^0}{L_0},} \end{equation}

and we assume the charged particle dynamics along $z$ to be mostly due to the electrostatic acceleration. This means that the ion velocity component along $z$, $u_z^i$, is $\varepsilon _m$ times smaller than the electron velocity component, $u_z^e$, which is instead comparable in amplitude to both ${\boldsymbol u}_\perp ^e$ and ${\boldsymbol u}_\perp ^i$. This also means that $J_z\,\equiv q_in_iu_z^i+q_en_eu_z^e\,=\,q_en_eu_z^e+\mathcal {O}(\varepsilon _m)$. In normalised units, ${\boldsymbol J}=n^e({\boldsymbol u}^i-{\boldsymbol u^e})$, and therefore

(A15)\begin{equation} u_z^e=d_i\frac{\nabla^2\psi}{n_e}+\mathcal{O}(\varepsilon_m). \end{equation}

The idea is then to use these hypotheses combined with a strong guide field limit so to consider the drift-ordering expansion for the ${\boldsymbol u}_\perp ^\alpha$ fluid velocities. Using the standard procedure, we obtain from (A7) and (A8)) and (A12) the normalised velocity components:

(A16)\begin{gather} {\boldsymbol u}_\perp^e= \underbrace{{\boldsymbol e}_z\times{\boldsymbol\nabla}\varphi +({\boldsymbol e}_z\times{\boldsymbol\nabla}\psi)u_z^e -\frac{\rho_s^2}{d_i} {\boldsymbol e}_z\times \frac{{\boldsymbol \nabla}\boldsymbol{\cdot}{\boldsymbol\varPi}^e}{n_e}- S^{{-}1}{\boldsymbol e}_z\times{\boldsymbol J}_\perp }_{1{{\rm st}}\quad\text{order}} \ +\ \underbrace{\mathcal{O} (\varepsilon_m^{{1}/{2}}\varepsilon_B^2)}_{2{{\rm nd}}\quad\text{order}}, \end{gather}
(A17)\begin{gather} {\boldsymbol u}_\perp^i= \underbrace{{\boldsymbol e}_z\times{\boldsymbol\nabla}\varphi -S^{{-}1}{\boldsymbol e}_z\times{\boldsymbol J}_\perp{+}\mathcal{O}(\varepsilon_m\varepsilon_B)}_{1{{\rm st}}\quad\text{order}} -\underbrace{d_i \left(\frac{\partial}{\partial t}{\boldsymbol\nabla}\varphi +[({\boldsymbol e}_z\times{\boldsymbol\nabla}\varphi)\boldsymbol{\cdot}{\boldsymbol\nabla}] {\boldsymbol\nabla}\varphi\right) }_{2{{\rm nd}}\quad\text{order}}\ +\ \mathcal{O}(\varepsilon_B^3). \end{gather}

The $\mathcal {O}(\varepsilon _m\varepsilon _B)$ neglected terms in (A17) are due to the $({\boldsymbol e}_z\times {\boldsymbol \nabla }\psi )u_z^i$ contribution of the first-order drift-expansion. We note that the terms neglected in both equations are comparable if we order $\varepsilon _m\sim \varepsilon _B^2$.

Equations (A1) and (A2)) follow then from substitution of (A16) and (A17)) into (A9) and (A10) after a few further specific hypotheses. They are:

  1. (i) the ordering $B_z^1/B_z^0=b\sim \varepsilon _B^2$, which also implies $|{\boldsymbol J}_\perp |\sim \varepsilon _B^2$; and

  2. (ii) the assumption that the anisotropic electron pressure components are given by the first-order FLR corrections to a double adiabatic-type pressure tensor.

The original derivation had been obtained by Schep et al. (Reference Schep, Pegoraro and Kuvshinov1994) in the strictly collisionless limit, in which, in the geometry of interest to us, the pressure tensor components including FLR ‘gyrofluid’ corrections can be written in dimensional units as

(A18)\begin{equation} {\boldsymbol\varPi}^e=\left( \begin{array}{@{}ccc@{}} \displaystyle{ P_\perp^e +\dfrac{P_\perp^e}{\varOmega_e}\left(\dfrac{\partial u_x^e}{\partial y} + \dfrac{\partial u_y^e}{\partial x}\right)} & \displaystyle{ \dfrac{P_\perp^e}{\varOmega_e}\left(\dfrac{\partial u_y^e}{\partial y} - \dfrac{\partial u_x^e}{\partial x}\right)} & \displaystyle{ \dfrac{P_\perp^e}{\varOmega_e}\left(\dfrac{\partial u_z^e}{\partial y} \right)}\\ \displaystyle{ \dfrac{P_\perp^e}{\varOmega_e}\left(\dfrac{\partial u_y^e}{\partial y} - \dfrac{\partial u_x^e}{\partial x}\right)} & \displaystyle{ P_\perp^e -\dfrac{P_\perp^e}{\varOmega_e}\left(\dfrac{\partial u_x^e}{\partial y} + \dfrac{\partial u_y^e}{\partial x}\right)} & \displaystyle{ -\dfrac{P_\perp^e}{\varOmega_e}\left(\dfrac{\partial u_z^e}{\partial x} \right)}\\ \displaystyle{ \dfrac{P_\perp^e}{\varOmega_e}\left(\dfrac{\partial u_z^e}{\partial y} \right)} & \displaystyle{ -\dfrac{P_\perp^e}{\varOmega_e}\left(\dfrac{\partial u_z^e}{\partial x} \right)} & P_{\|}^e \end{array} \right). \end{equation}

The components of matrix (A18) can be obtained by following a standard procedure (see e.g. Thompson Reference Thompson1961; Roberts & Taylor Reference Roberts and Taylor1962; MacMahon Reference MacMahon1965 for the ion case and Cerri et al. (Reference Cerri, Henri, Califano, Del Sarto, Faganello and Pegoraro2013) for a recent re-derivation for both species in the strictly collisionless limit). We make the further hypothesis that

(A19)\begin{equation} P_{\|}^e\simeq P_{{\perp}}^e=n^eT^0 \end{equation}

with $T^0$ reference electron temperature (uniform in space and constant in time). Since normalisation to the MHD reference quantities means that ${\boldsymbol u}^e$ is normalised to $c_{A,\perp }=d_i\varOmega _i$ and spatial derivatives are normalised to $1/a$, using $\varOmega _i/\varOmega _e\simeq \varepsilon _m$, the combination of (A3), (A15), (A18) and (A19) gives:

(A20)$$\begin{gather} \frac{\rho_s^2}{d_i} \frac{({\boldsymbol\nabla}\boldsymbol{\cdot}{\boldsymbol\varPi}^e)}{n^e} \boldsymbol{\cdot}{\boldsymbol e}_z\,=\, \varepsilon_m d_i\, \frac{\rho_s^2}{n^e} \left( \frac{\partial n^e}{\partial x}\frac{\partial }{\partial y} \left(\frac{\nabla^2\psi}{n^e}\right)- \frac{\partial n^e}{\partial y}\frac{\partial }{\partial x} \left(\frac{\nabla^2\psi}{n^e}\right) \right)\nonumber\\ = \varepsilon_m d_i {\rho_s^2} \left[\ln n^e,\frac{\nabla^2\psi}{n^e}\right], \end{gather}$$
(A21)\begin{gather} \frac{\rho_s^2}{d_i}{\boldsymbol e}_z\times \frac{{\boldsymbol\nabla}\boldsymbol{\cdot}{\boldsymbol\varPi}^e}{n^e}= \frac{\rho_s^2}{d_i}{\boldsymbol e}_z\times{\boldsymbol\nabla}(\ln n^e) + \mathcal{O}\left(\varepsilon_m\varepsilon_B \left|\frac{{\boldsymbol\nabla}n^e}{n^e}\right|\right). \end{gather}

Note that, in (A21), FLR corrections contribute with terms $\sim \varepsilon _m\varepsilon _B$ smaller than the isotropic diamagnetic drift term: comparison with the neglected terms of (A16) and (A17) shows that they can be disregarded.

We can now rewrite (A9) without collisions as

(A22)\begin{equation} \frac{\partial}{\partial t} \left( \psi-d_e^2\frac{\nabla^2\psi}{n^e} \right) +{\boldsymbol u}_\perp^e\boldsymbol{\cdot}{\boldsymbol\nabla} \left(\psi-d_e^2 \frac{\nabla^2\psi}{n^e} \right) \,=\, \varepsilon_m d_i {\rho_s^2} \left[\ln n^e,\frac{\nabla^2\psi}{n^e}\right]. \end{equation}

From substitution of (A21) into (A16), we get, using $d_e^2=\varepsilon _m d_i^2$,

(A23)$$\begin{gather} {\boldsymbol u}_\perp^e\boldsymbol{\cdot}{\boldsymbol\nabla} \left(\psi-d_e^2 \frac{\nabla^2\psi}{n^e} \right) =\left[\varphi,\psi-d_e^2 \frac{\nabla^2\psi}{n^e}\right] -\frac{\rho_s^2}{d_i}[\ln n^e, \psi] \nonumber\\ +\, \varepsilon_m d_i {\rho_s^2} \left[\ln n^e,\frac{\nabla^2\psi}{n^e}\right] +\mathcal{O}(\varepsilon_B^3) + \mathcal{O}(\varepsilon_m^{1/2}\varepsilon_B^2). \end{gather}$$

Note the presence of a $[\ln n^e, \nabla ^2\psi /n^e]$ term, identical to that of (A21), which is here due the diamagnetic drift contribution in the convection term ${\boldsymbol u}_\perp ^e\boldsymbol {\cdot }{\boldsymbol \nabla u}_z^e$ of the electron momentum equation. We can then write

(A24)\begin{equation} \frac{\partial}{\partial t} \left( \psi-d_e^2\frac{\nabla^2\psi}{n^e} \right) + \left[\varphi,\psi-d_e^2 \frac{\nabla^2\psi}{n^e}\right] = \frac{\rho_s^2}{d_i}[\ln n^e, \psi]+ \mathcal{O}(\varepsilon_B^3) + \mathcal{O}(\varepsilon_m^{1/2}\varepsilon_B^2). \end{equation}

The exact cancellation of the bracket terms proportional to $\varepsilon _m d_i\rho _s^2$ is known as ‘gyroviscous (or, better ‘gyrofluid’) cancellation’ (see Roberts & Taylor Reference Roberts and Taylor1962). The terms involved are however at least $\varepsilon _m^{1/2}$ smaller than the other ones, even when $d_e$ and $\rho _s$ are left unordered with respect to $\varepsilon _B$ or $\varepsilon _m$. Equation (A1) in the collisionless limit is finally recovered once we use continuity equations for ions and quasi-neutrality. From (A6), the continuity equation for ions can be rewritten as

(A25)\begin{equation} \frac{\partial}{\partial t}\ln n^i + [\varphi,\ln n^i] = d_i \frac{\partial}{\partial t}\nabla^2\varphi + d_i [\varphi,\nabla^2\varphi] +\mathcal{O}(\varepsilon_B^3). \end{equation}

Only the first-order ${\boldsymbol E}\times {\boldsymbol B}$ term of (A17) survives as a contribution to the convection term ${\boldsymbol u}^i_\perp \boldsymbol {\cdot }{\boldsymbol \nabla }\ln n^i$, since we can heuristically order $|{\boldsymbol \nabla }\ln n^i|\sim \varepsilon _B$. Equation (A25) leads us to solve ${\rm d}(\ln n^i)/{\rm d}t=d_i\,{\rm d}(\nabla ^2\varphi )/{\rm d}t$, of which we can take the solution $\ln n^i=d_i \nabla ^2\varphi$ that verifies the heuristic hypothesis about the ordering of the spatial gradient of $\ln n^i$. From quasi neutrality, we finally write

(A26)\begin{equation} \ln n^e = d_i \nabla^2\varphi, \end{equation}

which brings (A24) to the form (A1) once we approximate $n^e\simeq n^0$ (i.e. $n^e\simeq 1$ in normalised units) in the left-hand side gradients of $\nabla ^2\psi /n^e$.

The equation for $\nabla ^2\varphi$ is finally obtained from (A10) using the same approximation of the density $n^e$. Such equation follows from combining (A25) with the electron continuity equation, which, using (A16), reads

(A27)\begin{equation} \frac{\partial}{\partial t}\ln n^e + [\varphi,\ln n^e] + d_i \left[ \psi,\frac{\nabla^2\psi}{n^e} \right] + \mathcal{O}(\varepsilon_B^3) = \mathcal{O}(\varepsilon_m^{1/2}\varepsilon_B^2). \end{equation}

Note that the $[\psi,\nabla ^2\psi /n^e]$ contribution comes from the only non-null divergence term of (A16), that is, from ${\boldsymbol \nabla }\boldsymbol {\cdot }(({\boldsymbol e}_z\times {\boldsymbol \nabla }\psi )u_z^e)$.

When collisions are included, the $+S^{-1}\nabla ^2\psi$ contribution is recovered at the right-hand side of (A22) and (A23). In principle, the pressure tensor (A18) should be modified so as to include ‘gyroviscous’ corrections. These are provided, for example, by the model of Braginskii (Reference Braginskii1958). However, as long as the cold ion limit is formally assumed, so as to get rid of the ion temperature contribution, the corrections to the components of ${\boldsymbol \varPi }^e$ due to electron-ion viscosity are $\varepsilon _m$ smaller than $S^{-1}$, and thus are also negligible in (A18). The last, further contribution we should care about is an additional $+S^{-1}\nabla ^2b \sim \mathcal {O}(S^{-1}\varepsilon _B^2)$ term at the right-hand side of (A25), which comes from the divergence of the $-S^{-1}{\boldsymbol e}_z\times {\boldsymbol J}_\perp$ term of ion drift velocity, (A17). This term is negligible under the further assumption $S^{-1}\ll d_i$. It must be noted that the latter choice is consistent with the fact that this two-field set of equations correspond to the extended MHD model in which the Hall term is neglected.

The quantitative comparison between the derivation here presented and that suggested by Kleva et al. (Reference Kleva, Drake and Waelbroeck1995), Wang & Bhattcharjee (Reference Wang and Bhattcharjee1995) and Bian & Tsiklauri (Reference Bian and Tsiklauri2009), in which the $\rho _s^2$-related Poisson bracket is interpreted as associated with the Hall-term contribution in Ohms’ law – see also Del Sarto et al. (Reference Del Sarto, Califano and Pegoraro2006) and appendices of Del Sarto et al. (Reference Del Sarto, Pucci, Tenerani and Velli2016) – will be discussed elsewhere.

We finally recall that although in this modelling we assumed, since the beginning, strict translational invariance along $z$, i.e. $\partial /\partial z=0$, the RMHD modelling in a strong guide field, large aspect ratio tokamak with toroidal coordinates $(r,\theta,\varphi )$ allows for the inclusion of the derivatives along $\varphi$ that are ordered $\partial /\partial \varphi \sim \varepsilon _B \partial /\partial \theta$, i.e. $k_\varphi \sim \varepsilon _B k_\theta$ (Strauss Reference Strauss1976, Reference Strauss1977). This maps into a Cartesian ‘extended-slab’ RMHD modelling that includes the derivatives along $z$ as corrections of order $\epsilon _B$ with respect to the perpendicular derivatives, i.e. $\partial /\partial z \sim \varepsilon _B \partial /\partial y$ or $k_z\sim \varepsilon _B k_y$ (see, e.g. Matthaeus & Lamkin Reference Matthaeus and Lamkin1986 or Biskamp Reference Biskamp2000, p. 17). In the simplest, cold-electron limit, these corrections lead to further terms $+\partial \varphi _1 /\partial z$ and $+\partial \nabla ^2\psi _1 /\partial z$ on the right-hand side of (A1) and (A2), respectively. The role of these terms, as well as of further $\rho _s^2$-related contributions linked to the parallel electron compressibility in the nonlinear evolution of tearing modes with different helicities, has been first investigated in an extended version of model equations (A1) and (A2) by Grasso et al. (Reference Grasso, Borgogno, Califano, Farina, Pegoraro and Porcelli2004) and Borgogno et al. (Reference Borgogno, Grasso, Califano, Pegoraro and Farina2005).

Appendix B. About alternative definitions of the reconnection rate

In the literature, alternative quantitative definitions exist, which, for historical reasons, are typically associated with the term ‘reconnection rate’, although they have been formulated under the hypothesis of steady reconnection. In the notation and geometry choice we make here, this alternative estimate of the reconnection rate can be written as

(B1)\begin{equation} \mathcal{R}_{\text{steady}} = \frac{1}{L_0}U_{x}|_{\tiny \text{near }X\text{-point}}^{\tiny \text{at some } x\neq 0}. \end{equation}

This definition originates from the Sweet–Parker model (Parker Reference Parker1957), in which the rate at which the magnetic flux is reconnected, normalised with respect to the reference length $L_0$, is estimated as the component of the velocity ${\boldsymbol U}$ perpendicular to the neutral line such that

(B2)\begin{equation} ({U_{x}} B_{y})|_{\tiny \text{near}\ X\text{-point}}^{\tiny \text{at some } x\neq 0} = \mathcal{F}_{z}|_{\tiny \text{near}\ X\text{-point}}^{\tiny \text{at some } x\neq 0}. \end{equation}

This relation holds in the neighbourhood of the neutral line under steadiness assumptions, for which the out-of-plane component of the electric field ($E_z$), which in planar reconnection is completely inductive (i.e. $E_z=-\partial \psi /\partial t$), is zero – cf. (2.7). In this case, the velocity ${U_{x}}$ gives the rate per unit length $L_0$ at which the magnetic field lines are pushed and merged at the $X$-point.

Quantitative estimates of $\mathcal {R}_{\text {steady}}$ can be obtained by arguments relying on the continuity of the flow, which relate the velocity $U_x$, at which the magnetic field is dragged to the $X$ point, to the velocity $U_x$ at which, after reconnection has occurred, the field is transported out of it, along the neutral line ($U_{x}L\simeq U_{y}a$ – cf. figure 2a). These arguments also require a force balance condition implying that the upstream velocity $U_{y}$, evaluated close to the neutral line and sufficiently far from the $X$-point, is of the order of the Alfvén velocity evaluated with respect to $B_{y}$: these arguments allow one to relate $U_{x}|_{\text{near }X\text{-point}}$ to the dissipation mechanism given by $\mathcal {F}_{z}$ (Parker Reference Parker1957; Petschek Reference Petschek1964; Parker Reference Parker1973; Park et al. Reference Park, Monticello and White1984; Wesson Reference Wesson1990).

Note that in the tearing mode scenario the reference length $L_0$ is a macroscopic quantity associated with the shear length $a$ of the magnetic equilibrium, i.e. $L_0=a$. In the first astrophysical applications of the steady reconnection scenario, instead, in which a distinction between equilibrium and perturbed quantities is not required, the macroscopic reference length was usually assumed to be of the order of the current sheet length $L$, i.e. $L_0=L\gg a$, since the shear length $a$ of the magnetic field $B_y(x)$ associated with the reconnecting current sheet was, in those contexts, a microscopic quantity. Thus, an alternative definition of ‘reconnection rate per unit length $L$’ is often met in the literature. This was defined by Petschek (Reference Petschek1964) analogously to the Alfvénic Mach-number, from which it takes the symbol, as

(B3)\begin{equation} M_{\text{rec}} = \frac{ U_{y}|_{\tiny \text{near }X\text{-point}}} {\quad U_{x}|_{\tiny \text{far from }X\text{-point}}}\simeq \frac{U_{y}|_{\tiny \text{near }X\text{-point}}}{c_A(L)}, \end{equation}

where $c_A(L)$ is the Alfvén velocity evaluated in terms of $B_y$, measured sufficiently far from the neutral line along the neutral line $x=0$, in a region where ideal MHD is valid. In the steady reconnection scenario, this definition is practically equivalent to (B1).

Finally, a further alternative definition of the reconnection rate is also often adopted,

(B4)\begin{equation} \tilde{\mathcal{R}}_{\text{rec}} = \left.\frac{ 1}{E_{\|}}\frac{{\rm d} E_{\|}}{{\rm d}t} \right|_{X\text{-point}},\quad E_{\|}\equiv \frac{{\boldsymbol E}\boldsymbol{\cdot}{\boldsymbol B}}{B}. \end{equation}

We conclude by noting that while no ambiguity exists for the evaluation of the reconnection rate in the tearing-type instabilities with fixed wavenumber on which we are interested in this work, the situation becomes of course more complex in experimental contexts or in nonlinear numerical simulations of reconnection processes, especially if several reconnection sites ($X$-points) are simultaneously present and dynamically evolve in time. In those cases, estimates based on local measures like those provided by (2.13), (B3) and (B4) are of course more appealing for practical reasons. Their use for getting numerical values to be compared with scalings predicted by specific reconnection models should be however handled with care, since the way precision issues are determined in the measurement of numerical or experimental values is often unlikely to be sufficient to discriminate between a reconnection scenario or the other, e.g. in assessing the regime of reconnection events secondary to primary ones. Conversely, theoretical arguments supporting the existence or not of some specific regime are likely to provide a more accurate guide for the operational definition of reconnection rate that should be adopted in each case (cf. reference quoted just before (2.10)).

Appendix C. Integration of the boundary layer equation via Fourier transformation

Tackling the integration of tearing modes in the Fourier space with respect to the variable $x$ is particularly useful since the overall order of the differential equations results is lowered (note that $\psi _0\sim x \to {\rm d}/{\rm d}k_x$, whereas $\psi _1^{(N)}\to \sim k_x^N \hat {\psi }_1$). This approach is possible since the eigenfunctions we seek are bounded at each time with respect to the $x$ variable. Also note that Fourier representation is quite natural in the mathematical treatment of the gyrokinetic operator in the Vlasov–Poisson equation which, at some level of the analysis, must be treated when FLR effects of electrons and/or of ions are included. However, while the integration in the Fourier space may be more efficient, from an analytical point of view, for obtaining the dispersion relation and the asymptotic scalings of $\gamma$ and $\delta \sim \delta _1$, it has the drawback of requiring further analysis – not always trivial – for the computation of the eigenfunction profile in closed form in the coordinate space (i.e. the inverse Fourier transforms of the eigenfunctions integrated in the $k_x$-space must be evaluated).

C.1 A brief historical review on the boundary layer approach to tearing-type equations in the Fourier space

Fourier analysis was probably first applied to integrate interchange-type eigenmodes of MHD-type equations by Coppi (Reference Coppi1964b), and was used in the tearing mode problem by Ara et al. (Reference Ara, Basu, Coppi, Laval, Rosenbluth and Waddell1978) to find approximate solutions of the inner equations from which to estimate the weight of ion–ion viscosity in the dispersion relation via a variational approach. A more detailed Fourier approach to tearing-type equations has been developed by Pegoraro & Schep (Reference Pegoraro and Schep1981) to model low-frequency modes in toroidal tokamak geometry. This approach, which has been later used in several other specific reconnection models, consists in performing the Fourier transformation with respect to the stretched variable $\zeta$ of (5.9) and (5.10), and then combining the inner-layer equations for $\psi _1$ and $\varphi _1$ into a single equation for $\psi _1$. The integration of the tearing-type equations including FLR effect of both ions (for which a continuity equation was derived from a gyrokinetic one) and electrons (of which the isothermal limit was taken) has been first detailed by Pegoraro & Schep (Reference Pegoraro and Schep1986), using the ballooning representation, and then, starting from a different kinetic model and using a somewhat different approach, by Cowley et al. (Reference Cowley, Kulsrud and Hahm1986). The technique detailed by Pegoraro & Schep (Reference Pegoraro and Schep1986) has been further used for the internal $m = 1$ kink mode in high-$\beta$ regimes (Pegoraro et al. Reference Pegoraro, Porcelli and Schep1989). Relying on the same type of analysis, Porcelli (Reference Porcelli1991) derived the scaling laws for the collisionless regime in which magnetic reconnection occurs due to finite electron inertia and in combination with FLR effects. The Fourier analysis was also applied to investigate the scaling laws of the collisionless reconnection instabilities occurring in high frequency fluid regimes where ions form a neutralising background (Bulanov et al. Reference Bulanov, Pegoraro and Sakharov1992; Attico, Califano & Pegoraro Reference Attico, Califano and Pegoraro2000, Reference Attico, Califano and Pegoraro2001). The integration of tearing-type modes in Fourier space was then extended to other regimes in which electron temperature gradients at equilibrium are admitted and finite electron $\beta$ effects are included, while a full gyro-kinetic model is taken for describing the ion response: motivated by the operational regime of large-size tokamaks, which is expected to fall in the semi-collisional regimes where the width of the semi-collisional region is much smaller that the ion Larmor radius and finite $\beta$-effects are expected to play an important role, Connor et al. (Reference Connor, Hastie and Zocco2012b) investigated tearing-type instabilities in low$-\beta$ and high-$\beta$ regimes and the transition between them by relying on integration in the Fourier-space. In doing so, these authors also made a comparison of the solutions they obtained with those previously found in the coordinate space by Cowley et al. (Reference Cowley, Kulsrud and Hahm1986). Here, solutions of tearing equations in the semi-collisional limit were obtained by starting from a full kinetic model and by performing an expansion of the eigenfunctions in powers of the ratio between the integration layer width over the ion Larmor radius. The technique used by Connor et al. (Reference Connor, Hastie and Zocco2012b) is analogous to the boundary layer approach discussed by Pegoraro & Schep (Reference Pegoraro and Schep1981, Reference Pegoraro and Schep1986): using the fact that the dynamics of ions decouples from that of electrons at a distance of the order of the ion gyro-radius, the system of equations in each region was first transformed into a single equation for the current density in the Fourier space. Then, for the low$-\beta$ regime, the current density equation for ions was solved by Connor et al. (Reference Connor, Hastie and Zocco2012b) in Fourier space, while that for electrons was solved in real space. These solutions were made to match each other at some intermediate layer, while the solution in the ion region was also required to match that in the ideal MHD region, located far from the resonance surface. The technique to find the solution in the ion and electron regions consisted in expanding and matching the current density in powers of $\beta$. This method leads to a general dispersion relation that extends that of previous models by including four branches: the ion drift mode, the drift tearing mode, and other two branches corresponding to the kinetic Alfvén waves which couple to drift tearing modes for finite values of $\beta$.

C.2 Comparison of the analysis of §§ 46 with the Fourier approach

In the analysis of §§ 46, we have considered the low$-\beta$ regime with isothermal electrons, no equilibrium density fluctuations and cold ions, which rules out the coupling with drift modes while the coupling with kinetic Alfvén waves is restricted to the branch described by (A4). It is easy to show the essential points of the Fourier approach of Pegoraro & Schep (Reference Pegoraro and Schep1986) discussed above, by making a direct comparison with the equations discussed in §§ 46 by Fourier transforming (5.9) and (5.10), and to combine them together so as to write a single equation for the Fourier components of the current density. To this purpose, let us first define Fourier transform for a function $f(\zeta )$ as

(C1)\begin{equation} \hat{f} = \int_{-\infty}^{+\infty} f(\zeta) {\rm e}^{{\rm i} q \zeta} \,{\rm d}\zeta, \quad \zeta = \frac{x}{\sigma}, \quad q = k_x \sigma, \end{equation}

where we have re-called the definition of the stretched variable, and therefore the relation of the ‘stretched’ Fourier coordinate $q$ with $k_x$, the standard Fourier-conjugate wavenumber of $x$. For the scale $\sigma$, we have used the same notation of Porcelli (Reference Porcelli1991). In the notation adopted throughout the present manuscript, it would read indeed $\delta _1$ or $\delta _2$. Multiplying (5.9) and (5.10) by ${\rm e}^{{\rm i}q\zeta }$ and integrating from $-\infty$ to $+\infty$ while using the definition (C1), one obtains respectively

(C2)\begin{equation} \frac{q^2}{\hat{\mathcal{G}}^2} \frac{{\rm d}^2\hat{\psi}_1}{{\rm d}q^2}- \left(1 + \frac{q^2}{\hat{\mathcal{G}}^2}\right) \hat{\psi}_1 ={-} {\rm i} d_e^2 \frac{{\rm d}\hat{\tilde{\varphi}}}{{\rm d}q}, \quad \hat{\tilde{\varphi}} ={-} {\rm i} \frac{{\rm d}\hat{\psi}_1}{{\rm d}q}. \end{equation}

Taking the derivative of the second equation in (C2) with respect to $q$ and substituting the results in the former, one finds

(C3)\begin{equation} \left( d_e^2 + \frac{q^2}{\hat{\mathcal{G}}^2} \right) \frac{{\rm d}^2\hat{\psi}_1}{{\rm d}q^2} - \left( 1 + \frac{q^2}{\hat{\mathcal{G}}^2} \right) \hat{\psi}_1 = 0. \end{equation}

This equation represents an eigenvalue problem for the eigenfunction $\hat {\psi }_1$. For analytical convenience, it is however useful to transform it in an eigenvalue problem for the current density $\hat {J}_1 = q^2 \hat {\psi }_1$. After this substitution, one obtains

(C4)\begin{equation} \left( d_e^2 \hat{\mathcal{G}}^2 + q^2 \right) \frac{{\rm d}^2}{{\rm d}q^2} \left( \frac{\hat{J}_1}{q^2} \right) - (q^2 \hat{\mathcal{G}}^2 + 1) \hat{J}_1 = 0. \end{equation}

This is the cold-ion limit of (3) of Porcelli (Reference Porcelli1991). The equivalence is evident once the following correspondence between variables and parameters of the aforementioned reference and those of §§ 46 are established: $\mathcal {\varGamma } = \gamma$, $\varDelta ^2 = d_e^2 (1 + S^{-1}/{\gamma }) \rightarrow d_e^2$, and $\rho _\tau = (\rho _s^2 + \rho _i^2)^{{1}/{2}} \rightarrow \rho _s$.

Here, we do not discuss further the solution of (C4), since it represents a specific case of the equation already solved by Pegoraro & Schep (Reference Pegoraro and Schep1981, Reference Pegoraro and Schep1986) and further used by Pegoraro et al. (Reference Pegoraro, Porcelli and Schep1989), Porcelli (Reference Porcelli1991), Attico et al. (Reference Attico, Califano and Pegoraro2000, Reference Attico, Califano and Pegoraro2001) and Connor et al. (Reference Connor, Hastie and Zocco2012b).

Appendix D. An example of renormalisation of a differential equation

Consider (6.12) that we rewrite here for convenience as

(D1)\begin{equation} \frac{\tilde{\varphi}_1''}{c_0} - \frac{A}{B C} {\zeta^2} \frac{\tilde{\varphi}_1}{c_0} = \frac{A}{B}\, \zeta, \end{equation}

where

(D2)\begin{equation} A\equiv \ell^2,\quad B\equiv{\bar{d}_e^2},\quad C\equiv {\mathcal{G}^2}. \end{equation}

As an example, let us consider the case in which also the scale $\ell$ is non-trivial (i.e. $\ell \neq 1$). From the point of view of physical dimensions, (6.12) and (D1) are already written in dimensionless form but we want to check if it is possible to bring them to the ‘cleaner’ form of (6.13), after further renormalisation of quantities. Suppose then to multiply (D1) by $A^{a}B^{b}C^{c}$, where $a,b,c$ are here numerical coefficients to be determined a posteriori. Estimating, just for the purpose of dimensional analysis, $\tilde {\varphi }_1''\sim \tilde {\varphi }_1/\zeta ^2$, we obtain the following system:

(D3)\begin{equation} A^{a}B^{b}C^{c} \left(\frac{\tilde{\varphi}_1}{c_0}\right) \zeta^{{-}2}\sim 1, \quad A^{a+1}B^{b-1}C^{c-1} \left(\frac{\tilde{\varphi}_1}{c_0}\right) \zeta^{2}\sim 1, \quad A^{a+1}B^{b-1}C^{c} \zeta\sim 1. \end{equation}

The choice of fixing all right-hand side coefficients equal to a number is made possible by the fact that the number of independent conditions is equal to or larger than the number of unknown coefficients $a,b,c$: the parameters $A, B$ and $C$ can be thus completely absorbed in the redefinition of $\zeta$ and $\tilde {\varphi }_1/c_0$. This is the case also when the scale $\ell$ is absent, i.e. when $\ell =A=1$. When the number of parameters is larger than the number of constraints, some further parameter(s) corresponding to a dimensional combination(s) of powers of $A, B,\ldots$ will appear in the renormalised equations. The system above can be therefore solved to find that the values of the sought coefficients are $a=-3/4$, $b=3/4$, $c=-1/4$. This means that we can define

(D4)\begin{equation} \left. \begin{gathered} z\equiv A^{a+1}B^{b-1}C^{c}\zeta = A^{1/4}B^{{-}1/4}C^{{-}1/4} {\zeta},\\ \varPhi \equiv A^{{-}a-1}B^{{-}b+1}C^{{-}c-1} \left(\frac{\tilde{\varphi}_1}{c_0}\right) = A^{{-}1/4}B^{1/4} C^{{-}3/4} \left(\frac{\tilde{\varphi}_1}{c_0}\right). \end{gathered} \right\} \end{equation}

For $\ell =1$, this corresponds to (6.13).

Appendix E. Convergence and independence of the solutions of the hypergeometric form of the inner equation

We here discuss some technical details about the solutions ((7.25) and (7.27)) of the innermost equation, which we have obtained by evaluation of the residues of its hypergeometric form.

E.1 Convergence of $\psi _\alpha$ and $\psi _\beta$

Following the ratio test of sequences for the coefficients of $(1 + \zeta ^2)^{s}$, one finds $\lim _{m\to \infty } {g_{m+1}}/{g_m} = 1$, meaning that the power series (7.25) and (7.27) converge when $\| 1 + \zeta ^2 \| > 1$. This region covers all the complex plane, except when $\zeta =0$. In the complex plane, the boundary corresponding to the convergence radius $\| 1 + \zeta ^2 \| =1$ is given by

(E1)\begin{equation} (1 + \zeta_R^2 - \zeta_I^2)^2 + 4 \zeta_R^2 \zeta_I^2 = 1, \end{equation}

which is a curve forming two lobes joining at $\zeta = 0$ and intersecting the $\zeta _R=0$ axis at $\zeta _I = \sqrt {2}$, while the maximum width of the lobes occur at $\zeta _R = \pm {1}/{2}$ and $\zeta _I = \pm \sqrt {3}/{2}$.

The region inside the lobes of figure 17 represents the divergence region, which includes the singular points at $\zeta _I = \pm i$.

Figure 17. The region where the solutions $\psi _\alpha$ and $\psi _\beta$ diverge is here displayed in a grey colour inside two lobes that join each other at $\zeta =0$. Outside these two lobes, both series converge.

E.2 Linear independence of $\psi _\alpha$ and $\psi _\beta$

It is not difficult to prove that the two solutions given by (7.25) and (7.27) are linearly independent in the domain of convergence, say $\mathcal {D}_{\alpha,\beta }$, where both $\psi _\alpha$ and $\psi _\beta$ are defined. One way to prove this is to show that their Wronskian is non-zero everywhere in $\mathcal {D}_{\alpha,\beta }$. An easier way to proceed is to investigate the independence of the leading terms in both series. Taking $m = 0$, (7.25) and (7.27) become

(E2)\begin{equation} \psi_{\alpha,0} = \frac{\varGamma(-\nu)}{\varGamma\left(\dfrac{5}{4}- \dfrac{\nu}{2}\right) \varGamma\left(\dfrac{1}{4}- \dfrac{\nu}{2}\right)} (1 + \zeta^2)^{{\alpha}/{2}},\quad \psi_{\beta,0} = \frac{\varGamma(\nu)}{\varGamma\left(\dfrac{5}{4} + \dfrac{\nu}{2}\right) \varGamma\left(\dfrac{1}{4} + \dfrac{\nu}{2}\right)} (1 + \zeta^2)^{{\beta}/{2}}. \end{equation}

The coefficients in the two previous relations are non-zero. Therefore, if $\alpha \neq \beta$, then $\psi _{\alpha,0}$ and $\psi _{\beta,0}$ are linearly independent, and so are $\psi _\alpha$ and $\psi _\beta$. However, if $\alpha = \beta$, then $\nu = 0$. Using the definition of the $\varGamma$ function for positive and negative arguments, this means that the growth rate $\gamma$ is purely imaginary. Therefore, in this case, the configuration is stable to tearing-type modes, which is not a case of interest here.

Another interesting remark related to (7.25) and (7.27) is that they are even functions with respect to $\zeta$. Due to this symmetry, we expect that there is also an odd solution with respect to $\zeta$ which will be obviously linearly independent of the even solutions. The existence of this additional solution would apparently raise a paradox, because we would now have three independent solutions: the two independent even solutions $\psi _\alpha$ and $\psi _\beta$, and one odd solution, whereas a second-order ODE cannot have more than two linearly independent solutions. This apparent paradox can be solved by observing that going from $\zeta = 0^+$ to $\zeta = 0^{-}$ in the complex plane by passing through $\zeta =0$ is not possible for any of the two series, because both $\psi _\alpha$ and $\psi _\beta$ diverge there. Therefore, one should follow one of the lobes drawing the boundary of the divergence region in the complex plane (see figure 17), that is, one of the lobes encircling the singularities at $\zeta =\pm i$. Thus, to construct this solution, one can use a linear combination of the two series. This combination can be obtained using the integral representation and by choosing a path $C_\alpha$ (or $C_\beta$, alternatively) that goes around the poles associated with the exponent $\alpha$ (or $\beta$) by thus avoiding those of $\beta$ (or $\alpha$).

Appendix F. Heuristic estimation of the scaling laws by making use of the inverse gradient scale $l_c$ and of $\varDelta _{v_y}'$

Here we discuss the logical steps of a unified heuristic procedure which would make it possible to deduce the correct scalings in any regime and wavelength limit among those considered in this work, were the scaling of $\varDelta _{v_y}'$ be always deductible: in any regime, we obtain a scaling law, which is symmetric with respect to that of the small-$\varDelta '$ limit prior to the substitution of $\varDelta '\leftrightarrow \varDelta _{v_y}'$. The proposed procedure, however, is ‘self-consistent’ in all regimes with the exception of the large-$\varDelta '$, warm-electron limit, where the scaling of $\varDelta _{v_y}'$ appears to be a priori non-deducible via dimensional analysis. To facilitate the identification and presentation of the logical steps of the proposed heuristic procedure, they are summarised as statements and formulae in tables 35.

In particular, in table 3, we recall the approximated linear equations, definitions and general constraints we rely upon, as well as the conclusions that can be drawn for them regardless of the reconnection regime considered.

Finally, in table 5, we consider the cold-tearing regimes ($\rho _s= 0$), both collisionless and resistive, by showing how the heuristic approach that takes $\varDelta _{v_y}'$ into account works and still provides the good scalings. Although having introduced the new scale $\varDelta _{v_y}'$ makes the heuristic procedure slightly longer than the one discussed in § 9.3, it allows one to appreciate the ‘symmetry’ between $\varDelta '$ in the small-$\varDelta '$ limit and $\varDelta _{v_y}'$ in the large-$\varDelta '$ limit also in these cold-electron regimes. More importantly, and differently from the warm-reconnection regimes, it becomes manifest that for $\rho _s=0$, it is possible to close the system of heuristic conditions also in the large-$\varDelta '$ limit, so as to determine a priori the corresponding scaling of $\varDelta _{v_y}'$, which here always results to be the same of $\delta$.

Table 3. Starting equations, definitions and assumptions based on general constraints that we take for the heuristic-type approach in all reconnection regimes. Equations [T4.E1–T4.E2] are the eigenmode equations ((2.1) and (2.2)) in which we have assumed $k$ to be negligible with respect to the inverse scale of the spatial gradients of the eigenfunctions. Equations [T4.D1–T4.D3] are the operational definitions we have given in §§ 8, 9 and 10 of $\delta$, $l_c$ and $\varDelta _{v_y}'$, respectively. Equations [T4.C1,T4.C2] and [T4.H1] represent further conditions on the parameters that must be fulfilled regardless of the reconnection regime: [T4.C1] is a constraint that follows from [T4.E2] and definitions [T4.D2–T4.D3]; conditions [T4.C2a] and [T4.C2] correspond to (10.9) that we have discussed in the text; condition [T4.H1] corresponds to the ansatz (10.5) discussed in the text. Equations [T4.R1a] and [T4.R1b] are two immediate conclusions than can be drawn and allow us to specify hypothesis [T4.H1] in the small- and large-$\varDelta '$ limits, respectively.

In table 4, we specialise the results to the warm-tearing regimes ($\rho _s\neq 0$), both collisionless and resistive: in each regime, we report the further ansatz required to get the sought scalings in both the small- and large-$\varDelta '$ limits. The ‘symmetry’ of the scalings written in the small-$\varDelta '$ limit in terms of $\varDelta '$ with respect to the scaling written in the large-$\varDelta '$ limit in terms of $\varDelta _{v_y}'$ can be in this way appreciated, although the system of equations is not closed in the large-$\varDelta '$ limit and we must rely on additional information (e.g. numerical calculations performed by following the procedure detailed in previous § 10.3) to explicitly evaluate $\varDelta _{v_y}'$.

Table 4. Combining the hypotheses and results reported in table 2 with hypotheses [T5.H2–T5.H3] that are specific of the warm regimes, it is possible to deduce in a self-consistent way the set of equations [T5.R2–T5.R5]. It must be emphasised, however, that the system of condition is not closed in the large-$\varDelta '$ regime and additional information about $\varDelta _{v_y}'$ is required to obtain the corresponding scaling laws.

Table 5. Combining the hypotheses and results reported in table 2 with hypotheses [T6.H2] and [T6.H3] that are specific of the cold regimes, it is possible to deduce in a self-consistent way the set of equations [T6.C3]–[T6.R4]. Note that, in this case, the set of conditions is closed also in the large-$\varDelta '$ regime thanks to [T6.H2], which leads to the constraint [T6.C4] that sets the scaling of $\varDelta _{v_y}'$ to be the same of $\delta ^{-1}$. All scalings can be this way a priori determined without resorting to numerical calculations.

Footnotes

1 After acceptance of this manuscript the authors have become aware of another recent work (Granier et al. Reference Granier, Borgogno, Grasso and Tassi2022), in which an integration of the boundary layer equations in the coordinate space has been performed in the warm-electron, small-$\varDelta'$ regime of tearing modes.

2 In (Granier et al. Reference Granier, Borgogno, Grasso and Tassi2022) it has been pointed out that in the warm small-$\varDelta'$ regime the equation for $\varphi_1$ changes of form while moving from the layer of width $\delta_2$ to the innermost layer of width $\delta_1$: while this would formally require to carry out the full boundary layer calculation for its integration, doing so only yields negligible additive corrections of order $O(\delta_1/\rho_s)\ll 1$ to the growth rate scalings.

References

REFERENCES

Abramowitz, M. & Stegun, I. 1964 Handbook of Mathematical Functions: with Formulas, Graphs, and Mathematical Tables. Dover Ed.Google Scholar
Agullo, O., Muraglia, M., Benkadda, S., Poyé, A., Dubuit, N., Garbet, X. & Sen, A. 2017 a Nonlinear dynamics of turbulence driven magnetic islands. I. Theoretical aspects. Phys. Plasmas 24, 042308.CrossRefGoogle Scholar
Agullo, O., Muraglia, M., Benkadda, S., Poyé, A., Dubuit, N., Garbet, X. & Sen, A. 2017 b Nonlinear dynamics of turbulence driven magnetic islands. II. Numerical simulations. Phys. Plasmas 24, 042309.CrossRefGoogle Scholar
Alfvén, H. 1942 Existence of electromagnetic-hydrodynamic waves. Nature 150, 405.CrossRefGoogle Scholar
Ali, A., Li, J. & Kishimoto, Y. 2014 On the abrupt growth dynamics of nonlinear resistive tearing mode and the viscosity effect. Phys. Plasmas 21, 052312.CrossRefGoogle Scholar
Ara, G., Basu, B., Coppi, B., Laval, G., Rosenbluth, M.N. & Waddell, B.V. 1978 Magnetic reconnection and $m = 1$ oscillations in current carrying plasmas. Ann. Phys. 112, 443476.CrossRefGoogle Scholar
Aschwanden, M.J. 2020 Global energetics of solar flares. IV. Coronal mass ejection energetics. Astrophys. J. 903, 23.CrossRefGoogle Scholar
Aschwanden, M.J., Caspi, A., Cohen, C.M.S., Holman, G., Jing, J., Kretzschmar, M., Kontar, E.P., McTiernan, J.M., Mewaldt, R.A., O'Flannagain, A., et al. 2019 Global energetics of solar flares and coronal mass ejections. J. Phys.: Conf. Ser. 1332, 012002.Google Scholar
Aschwanden, M.J., Holman, G., O'Flannagain, A., Caspi, A., McTiernan, J.M. & Kontar, E.P. 2016 Global energetics of solar flares. III. Nonthermal energies. Astrophys. J. 832, 27.CrossRefGoogle Scholar
Aschwanden, M.J., Xu, Y. & Jing, J. 2014 Global energetics of solar flares. I. Magnetic energies. Astrophys. J. 797, 50.CrossRefGoogle Scholar
Attico, N., Califano, F. & Pegoraro, F. 2000 Fast collisionless reconnection in the whistler frequency range. Phys. Plasmas 7, 23812387.CrossRefGoogle Scholar
Attico, N., Califano, F. & Pegoraro, F. 2001 Charge separation effects in electron- magnetohydrodynamic reconnection. Phys. Plasmas 8, 16.CrossRefGoogle Scholar
Axford, W.I. 1984 Magnetic field reconnection. Geophysical Monograph Series, vol. 30, p. 1. American Geophysical Union.Google Scholar
Aydemir, A.Y. 1992 Nonlinear studies of $m= 1$ modes in high-temperature plasmas. Phys. Fluids B 4, 34693472.CrossRefGoogle Scholar
Basu, B. & Coppi, B. 1976 $\ldots$ Tech. Rep. PRR-76/38. M.I.T., Cambridge, MA.Google Scholar
Basu, B. & Coppi, B. 1981 Theory of $m=1$ modes in collisionless plasmas. Phys. Fluids 24, 465471.Google Scholar
Batchelor, G.K. 1950 On the spontaneous magnetic field in a conducting luquid in turbulent motion. Proc. R. Soc. Lond. A 201, 405.Google Scholar
Baum, P.J., Bratenahl, A. & White, R.S. 1971 Experimental study of the reconnection process. Radio Sci. 8, 917.CrossRefGoogle Scholar
Bender, C.M. & Orszag, S.A. 1978 Advanced Mathematical Methods for Scientists and Engineers 1: Asymptotic Methods and Perturbation Theory. Springer.Google Scholar
Bergmans, J. 2001 Vortex interactions in magnetized plasmas. PhD thesis, Technische Universiteit Eindhoven, Eindhoven (The Netherlands).Google Scholar
Bergmans, J. & Schep, T.J. 2001 Merging of plasma currents. Phys. Rev. Lett. 87, 195002.CrossRefGoogle ScholarPubMed
Berk, H.L., Mahajan, S.M. & Zhang, Y.Z. 1991 $m=1$ kink mode for layer widths comparable to the ion Larmor radius. Phys. Fluids B 3, 351.CrossRefGoogle Scholar
Bertin, G. 1982 Effects of local current gradients on magnetic reconnection. Phys. Rev. A 25, 1786.CrossRefGoogle Scholar
Betar, H., Del Sarto, D., Ottaviani, M. & Ghizzo, A. 2020 Multiparametric study of tearing modes in thin current sheets. Phys. Plasmas 27 (10), 102106.CrossRefGoogle Scholar
Bettarini, L., Landi, S., Rappazzo, F.A., Velli, M. & Opher, M. 2006 Tearing and Kelvin–Helmholtz instabilities in the heliospheric plasma. Astron. Astrophys. 452, 321.CrossRefGoogle Scholar
Bhattacharjee, A., Brunel, F. & Tajima, T. 1983 Magnetic reconnection driven by the coalescence instability. Phys. Fluids 26, 3332.CrossRefGoogle Scholar
Bhattacharjee, A., Germaschewski, K. & Ng, C.S. 2005 Current singularities: drivers of impulsive reconnection. Phys. Plasmas 12, 042305.CrossRefGoogle Scholar
Bhattacharjee, A., Huang, Y.M., Yang, H. & Rogers, B. 2009 Fast reconnection in high-Lundquist-number plasmas due to the plasmoid instability. Phys. Plasmas 16, 112102.CrossRefGoogle Scholar
Bian, N.H. & Tsiklauri, D. 2009 Compressible magnetohydrodynamics in a strong magnetic field. Phys. Plasmas 16, 064503.CrossRefGoogle Scholar
Bickley, W.G. 1937 LXXIII. The plane jet. Lond. Edinb. Dublin Philos. Mag. J. Sci. 23, 727.CrossRefGoogle Scholar
Birn, J. & Hones, E.W. Jr. 1981 Three-dimensional computer modeling of dynamic reconnection in the geomagnetic tail. J. Geophys. Res. 86, 6802.CrossRefGoogle Scholar
Biskamp, D. 1982 Dynamics of a resistive sheet pinch. Z. Naturforsch. 37a, 840.CrossRefGoogle Scholar
Biskamp, D. 1986 Magnetic reconnection via current sheets. Phys. Fluids 29, 1520.CrossRefGoogle Scholar
Biskamp, D. 1996 Magnetic reconnection in plasmas. Astrophys. Space Sci. 242, 165.CrossRefGoogle Scholar
Biskamp, D. 2000 Magnetic Reconnection in Plasmas. Cambridge University Press.CrossRefGoogle Scholar
Biskamp, D. & Bremer, U. 1994 Dynamics and statistics of inverse cascade processes in 2D magnetohydrodynamic turbulence. Phys. Rev. Lett. 72, 3819.CrossRefGoogle ScholarPubMed
Biskamp, D. & Sato, T. 1997 Partial reconnection in the nonlinear internal kink mode. Phys. Plasmas 4, 1326.CrossRefGoogle Scholar
Biskamp, D., Schwarz, E. & Drake, J.F. 1995 Ion-controlled collisionless magnetic reconnection. Phys. Rev. Lett. 75, 3850.CrossRefGoogle ScholarPubMed
Biskamp, D., Schwarz, E. & Drake, J.F. 1997 Two-fluid theory of collisionless magnetic reconnection. Phys. Plasmas 4, 1002.CrossRefGoogle Scholar
Biskamp, D., Schwarz, E. & Zeiler, A. 1998 Instability of a magnetized plasma jet. Phys. Plasmas 5, 24852488.CrossRefGoogle Scholar
Biskamp, D. & Welter, H. 1980 Coalescence of magnetic islands. Phys. Rev. Lett. 44, 1069.CrossRefGoogle Scholar
Biskamp, D. & Welter, H. 1989 Dynamics of decaying two-dimensional magnetohydrodynamic turbulence. Phys. Fluids B 1, 1989.CrossRefGoogle Scholar
Boozer, A.H. 2012 Theory of tokamak disruptions. Phys. Plasmas 19, 058101.CrossRefGoogle Scholar
Borgogno, D., Grasso, D., Califano, F., Pegoraro, F. & Farina, D. 2005 Aspects of three-dimensional magnetic reconnection. Phys. Plasmas 12, 032309.CrossRefGoogle Scholar
Boyd, T.J.M. & Sanderson, J.J. 2003 The Physics of Plasmas. Cambridge University Press.CrossRefGoogle Scholar
Braginskii, S.I. 1958 Transport phenomena in a completely ionized two-temperature plasma. J. Expl Theor. Phys. 6, 358.Google Scholar
Bratenahl, A. & Yeates, C.M. 1970 Experimental study of magnetic flux transfer at the hyperbolic neutral point. Phys. Fluids 13, 26962696.CrossRefGoogle Scholar
Brennan, D.P., Strait, E.J., Turnbull, A.D., Chu, M.S., La Haye, R.J., Luce, T.C., Taylor, T.S., Kruger, S. & Pletzer, A. 2002 Tearing mode stability studies near ideal stability boundaries in DIII-D. Phys. Plasmas 9 (7), 2998.CrossRefGoogle Scholar
Bulanov, S.V. 2017 Magnetic recoonection: from MHD to QED. Plasma Phys. Control. Fusion 59, 014029.CrossRefGoogle Scholar
Bulanov, S.V., Pegoraro, F. & Sakharov, A.S. 1992 Magnetic reconnection in electron magnetohydrodynamics. Phys. Fluids 8, 24992508.CrossRefGoogle Scholar
Bulanov, S.V., Sakai, J. & Syrovatskii, S.I. 1979 Tearing-mode instability in approximatively steady MHD configurations. Sov. J. Plasma Phys. 5, 157.Google Scholar
Cafaro, E., Grasso, D., Pegoraro, F., Porcelli, F. & Saluzzi, A. 1998 Invariants and geometric structures in nonlinear Hamiltonian magnetic reconnection. Phys. Rev. Lett. 80, 4430.CrossRefGoogle Scholar
Carreras, B.A., Rosenbluth, M.N. & Hicks, H.R. 1981 Nonlinear destabilisation of tearing modes. Phys. Rev. Lett. 46, 1131.CrossRefGoogle Scholar
Cassak, P. & Shay, M.A. 2012 Magnetic reconnection for coronal conditions: reconnection rates, secondary islands and onset. Space Sci. Rev. 172, 283.CrossRefGoogle Scholar
Cerri, S.S., Henri, P., Califano, F., Del Sarto, D., Faganello, M. & Pegoraro, F. 2013 Extended fluid models: pressure tensor effects and equilibria. Phys. Plasmas 20, 112112.CrossRefGoogle Scholar
Chen, X.L. & Morrison, P.J. 1990 Resistive tearing instability with equilibrium shear flow. Phys. Fluids B 2, 495.CrossRefGoogle Scholar
Chen, Q., Otto, A. & Lee, L.C. 1997 Tearing instability, Kelvin–Helmholtz instability, and magnetic reconnection. J. Geophys. Res. 102, 151.CrossRefGoogle Scholar
Choi, M.J. 2021 Interaction between a magnetic island and turbulence. Rev. Mod. Plasma Phys. 9, 1375.Google Scholar
Comisso, L. & Bhattacharjee, A. 2016 On the value of the reconnection rate. Phys. Plasmas 82, 595820601.CrossRefGoogle Scholar
Comisso, L., Grasso, D., Waelbroeck, F.L. & Borgogno, D. 2013 Gyro-induced acceleration of magnetic reconnection. Phys. Plasmas 20, 092118.CrossRefGoogle Scholar
Comisso, L., Huang, Y.-M., Lingam, M., Hirvijoki, E. & Bhattacharjee, A. 2018 Magnetohydrodynamic turbulence in the plasmoid-mediated regime. Astrophys. J. 854, 103.CrossRefGoogle Scholar
Connor, J.W., Cowley, S.C. & Hastie, R.J. 1990 Micro-tearing stability in tokamaks. Plasma Phys. Control. Fusion 32, 799.CrossRefGoogle Scholar
Connor, J.W., Hastie, R.J., Marchetto, C. & Roach, C.M. 2012 a A one-dimensional tearing mode equation for pedestal stability studies in tokamaks. J. Plasma Phys. 84, 725840301.CrossRefGoogle Scholar
Connor, J.W., Hastie, R.J. & Zocco, A. 2012 b Unified theory of the semi-collisional tearing mode and internal kink mode in a hot tokamak: implications for sawtooth modelling. Plasma Phys. Control. Fusion 54, 035003.CrossRefGoogle Scholar
Coppi, B. 1964 a Addendum on inertial interchange modes. Phys. Lett. 12, 213214.CrossRefGoogle Scholar
Coppi, B. 1964 b Gravitational instabilities in a compressible collisional plasma layer. Ann. Phys. 30, 178.CrossRefGoogle Scholar
Coppi, B. 1964 c Inertial instabilities in plasmas. Phys. Lett. 11, 226228.CrossRefGoogle Scholar
Coppi, B. 1965 Current-driven instabilities in configurations with sheared magnetic fields. Phys. Fluids 8, 2237.CrossRefGoogle Scholar
Coppi, B. & Detragiache, P. 1992 Magnetic topology transitions in collisionless plasmas. Phys. Lett. A 168, 59.CrossRefGoogle Scholar
Coppi, B., Galvao, R., Pellat, R., Rosenbluth, M.N. & Rutherford, P. 1976 Resistive internal kink modes. Fyz. Plazmy 2, 961.Google Scholar
Coppi, B., Laval, G. & Pellat, R. 1966 Dynamics of the geomagnetic tail. Phys. Rev. Lett. 16, 12071210.CrossRefGoogle Scholar
Coppi, B., Mark, J.W.-K., Sugiyama, L. & Bertin, G. 1979 Reconnecting modes in collisionless plasmas. Phys. Rev. Lett. 42, 1058.CrossRefGoogle Scholar
Cowley, S.C. & Hastie, R.J. 1988 Electron diamagnetism and toroidal coupling of tearing modes. Phys. Fluids 31, 426.CrossRefGoogle Scholar
Cowley, S.C., Kulsrud, R.M. & Hahm, T.S. 1986 Linear stability of tearing modes. Phys. Fluids 29 (10), 32303244.CrossRefGoogle Scholar
Cowling, T.G. 1933 The magnetic field of sunspots. Mon. Not. R. Astron. Soc. 94, 39.CrossRefGoogle Scholar
Cowling, T.G. 1967 Commission de la Magneto-Hydrodynamique et de la Physique des Gaz Ionises. In Reports on Astronomy/Proceedings of the Thirteenth General Assembly Prague 1967 (ed. L. Perek), p. 965. Springer.CrossRefGoogle Scholar
Cross, M.A. & van Hoven, G. 1971 Magnetic and gravitational energy release by resistive instabilities. Phys. Rev. A 4, 2347.CrossRefGoogle Scholar
Daughton, W. 1999 The unstable eigenmodes of a neutral sheet. Phys. Plasmas 6, 13291343.CrossRefGoogle Scholar
Daughton, W. & Karimabadi, H. 2005 Kinetic theory of collisionless tearing at the magnetopause. J. Geophys. Res. 110, A03217.Google Scholar
Daughton, W. & Scudder, J. 2006 Fully kinetic simulations of undriven magnetic reconnection with open boundary conditions. Phys. Plasmas 13, 072101.CrossRefGoogle Scholar
Del Sarto, D., Califano, F. & Pegoraro, F. 2003 Secondary instabilities and vortex formation in collisionless-fluid magnetic reconnection. Phys. Rev. Lett. 91, 235001.CrossRefGoogle ScholarPubMed
Del Sarto, D., Califano, F. & Pegoraro, F. 2005 Current layer cascade in collisionless electron-magnetohydrodynamic reconnection and electron compressibility effects. Phys. Plasmas 12, 012317.CrossRefGoogle Scholar
Del Sarto, D., Califano, F. & Pegoraro, F. 2006 Electron parallel compressibility in the nonlinear development of two-dimensional magnetohydrodynamic reconnection. Mod. Phys. Lett. B 8, 931961.CrossRefGoogle Scholar
Del Sarto, D., Marchetto, C., Pegoraro, F. & Califano, F. 2011 Finite Larmor Radius effects in the nonlinear dynamics of collisionless magnetic reconnection . Plasma Phys. Control. Fusion 53, 035008.CrossRefGoogle Scholar
Del Sarto, D. & Ottaviani, M. 2017 Secondary fast reconnecting instability in the sawtooth crash. Phys. Plasmas 24, 012102.CrossRefGoogle Scholar
Del Sarto, D., Ottaviani, M., Pucci, F., Tenerani, A. & Velli, M. 2018 Spontaneous magnetic reconnection of thin current sheets. In Comptes-rendues de la $21^e$ Rencontre du Non Linéaire, Paris (ed. É. Falcon, M. Lefranc, F. Pétrélis, & C.-T. Pham), vol. 21, p. 13. Université Paris Diderot, Non-Linéaires Publications.Google Scholar
Del Sarto, D. & Pegoraro, F. 2017 Shear-induced pressure anisotropization and correlation with fluid vorticity in a low collisionality plasma. Mon. Not. R. Astron. Soc. 475, 181.CrossRefGoogle Scholar
Del Sarto, D., Pucci, F., Tenerani, A. & Velli, M. 2016 “Ideal” tearing and the transition to fast reconnection in the weakly collisional MHD and EMHD regimes . J. Geophys. Res. 121, A021975.CrossRefGoogle Scholar
Deng, X.H. & Matsumoto, H. 2001 Rapid magnetic reconnection in the earth's magnetosphere mediated by whistler waves. Nature 401, 557.CrossRefGoogle Scholar
Doerk, H., Jenko, F., Pueschel, M.J. & Hatch, D.R. 2011 Gyrokinetic microtearing turbulence. Phys. Rev. Lett. 106, 155003.CrossRefGoogle ScholarPubMed
Dorland, W. & Hammett, G.W. 1993 Gyrofluid turbulence models with kinetic effects. Phys. Fluids B 5, 812835.CrossRefGoogle Scholar
Drake, J.F. & Kleva, R.G. 1991 Collisionless reconnection and the sawtooth crash. Phys. Rev. Lett. 66, 1458.CrossRefGoogle ScholarPubMed
Drake, J.F. & Lee, Y.C. 1977 Kinetic theory of tearing instabilities. Phys. Fluids 20, 035008.CrossRefGoogle Scholar
Drake, J.F., Swisdak, M., Schoeffler, K.M., Rogers, B.N. & Kobayashi, S. 2006 Formation of secondary islands during magnetic reconnection. Geophys. Res. Lett. L13105, 035008.Google Scholar
Dubrovin, B.A., Nivikov, S.P. & Fomenko, A.T. 1991 Modern Geometry. Methods and Applications: Part I. The Geometry of Surfaces, Transformation Groups, and Fields. Springer.Google Scholar
Dungey, J.W. 1950 Some Researches in Cosmic Magnetism. PhD thesis, Cambridge University, Cambridge (UK).Google Scholar
Dungey, J.W. 1953 Conditions for the occurrence of electrical discharges in astrophysical systems. Lond. Edinb. Dubl. Phil. Mag. 29 (7), 752–738.Google Scholar
Eastwood, J.P., Hapgood, M.A., Biffis, E., Benedetti, D., Bisi, M.M., Green, L., Bentley, R.D. & Burnett, C. 2018 Quantifying the economic value of space weather forecasting for power grids: an exploratory study. Space Weath. 16, 2052.CrossRefGoogle Scholar
Einaudi, G. & Rubini, F. 1986 Resistive instabilities in a flowing plasma: I. Inviscid case. Phys. Fluids 29, 2563.CrossRefGoogle Scholar
Elsasser, W.M. 1950 a Induction effects in terrestrial magnetism. Part III. Electric modes. Phys. Rev. 72, 827.Google Scholar
Elsasser, W.M. 1950 b The Earth's interior and geomagnetism. Rev. Mod. Phys. 22, 183.CrossRefGoogle Scholar
Faganello, M., Pegoraro, F., Califano, F. & Marradi, L. 2010 Collisionless magnetic reconnection in the presence of a sheared velocity field. Phys. Plasmas 17, 062102.CrossRefGoogle Scholar
Fedele, B., Negulescu, C. & Ottaviani, M. 2021 Analysis of the kolmogorov model with an symptotic preserving method. Phys. Lett. A 127522, 410.Google Scholar
Finn, J.M. & Kaw, P.K. 1977 Coalescence instability of magnetic islands. Phys. Fluids 20, 72.CrossRefGoogle Scholar
Firpo, M.-C. 2015 Microtearing turbulence: magnetic braiding and disruption limit. Phys. Fluids 22, 122511.Google Scholar
Fitzpatrick, R. 2008 Scaling of the peak magnetic reconnection rate in the inviscid Taylor problem. Nucl. Fusion 15, 024503.Google Scholar
Fitzpatrick, R., Hastie, R.J., Martin, T.J. & Roach, C.M. 1993 Stability of coupled tearing modes in tokamaks. Nucl. Fusion 33 (10), 1533.CrossRefGoogle Scholar
Franci, L., Cerri, S.S., Califano, F., Landi, S., Papini, E., Verdini, A., Matteini, L., Jenko, F. & Hellinger, P. 2017 Magnetic reconnection as a driver for a sub-ion-scale cascade in plasma turbulence. Astrophys. J. Lett. 850, L16.CrossRefGoogle Scholar
Franci, L., Landi, S., Matteini, L., Verdini, A. & Hellinger, P. 2016 Plasma beta dependence of the ion-scale spectral break of solar windturbulence: high-resolution 2D hybrid simulations. Astrophys. J. 883, 91.CrossRefGoogle Scholar
Franci, L., Papini, E., Del Sarto, D., Hellinger, P., Burgess, D., Matteini, L., Landi, S. & Montagud-Camps, V. 2022 Plasma turbulence in the near-Sun and near-Earth solar wind: a comparison via observation-driven 2D hybrid simulations. Universe 8, 453.CrossRefGoogle Scholar
Fruchtman, A. 1991 Penetration and expulsion of magnetic fields in plasmas due to the hall field. Phys. Fluids B 3, 1908.CrossRefGoogle Scholar
Furth, H. 1963 Hydromagnetic instabilities due to finite resistivity. In Propagation and Instabilities in Plasmas – Seventh Lockheed symposium on magnetohydrodynamics (December 1962) (ed. W.I. Futtermann), pp. 87–102. Stanford University Press.Google Scholar
Furth, H. 1964 Instabilities due to finite resistivity or finite current-carrier mass. In Proceedings of the “Enrico Fermi” International School of Physics of the Italian Physical Society, Course XXV. Advanced plasma Theory (ed. M.N. Rosenbluth), p. 159. Academic Press.Google Scholar
Furth, H.P., Killeen, J. & Rosenbluth, M.N. 1963 Finite-resistivity instabilities of a sheet pinch. Phys. Fluids 6, 459484.CrossRefGoogle Scholar
Furth, H.P., Rutherford, P.H. & Selberg, H. 1973 Tearing mode in cylindrical tokamak. Phys. Fluids 16, 1054.CrossRefGoogle Scholar
Giovanelli, R.G. 1946 A theory of chromospheric flares. Nature 158, 81.CrossRefGoogle Scholar
Gladd, N.T., Drake, J.F., Chang, C.L. & Liu, C.S. 1980 Electron temperature gradient driven microtearing mode. Phys. Fluids 23, 1182.CrossRefGoogle Scholar
Gopalswamy, N. 2016 History and development of coronal mass ejections as a key player in solar terrestrial relationship. Geosci. Lett. 3 (1), 1.CrossRefGoogle Scholar
Gosling, J.T. & McComas, D.J. 1987 Field line draping about fast coronal mass ejecta: a source of strong out-of-the-ecliptic interplanetary magnetic fields. Geophys. Res. Lett. 14 (4), 355.CrossRefGoogle Scholar
Gradshteyn, I.S. & Ryzhik, I.M. 2015 Table of Integrals, Series, and Products (VIII Edition). Elsevier.Google Scholar
Granier, C., Borgogno, D., Grasso, D. & Tassi, E. 2022 Gyrofluid analysis of electron $\beta_{e}$ effects on collisionless reconnection. J. Plasma Phys. 88 (1), 905880111.CrossRefGoogle Scholar
Grasso, D., Borgogno, D., Califano, F., Farina, D., Pegoraro, F. & Porcelli, F. 2004 Numerical solution of a reduced model of collisionless magnetic reconnection in two and three dimensions. Phys. Plasmas 164, 23.Google Scholar
Grasso, D., Borgogno, D. & Pegoraro, F. 2007 Secondary instabilities in two- and three-dimensional magnetic reconnection in fusion relevant plasmas. Phys. Plasmas 14, 055703.CrossRefGoogle Scholar
Grasso, D., Borgogno, D., Pegoraro, F. & Tassi, E. 2009 Coupling between reconnection and Kelvin–Helmholtz instabilities in collisionless plasmas. Nonlinear Process. Geophys. 16, 241.CrossRefGoogle Scholar
Grasso, D., Borgogno, D., Tassi, E. & Perona, A. 2020 Asymmetry effects driving secondary instabilities in two-dimensional collisionless magnetic reconnection. Phys. Plasmas 2, 012302.CrossRefGoogle Scholar
Grasso, D., Califano, F., Pegoraro, F. & Porcelli, F. 2001 Phase mixing and island saturation in Hamiltonian reconnection. Phys. Rev. Lett. 86, 5051.CrossRefGoogle ScholarPubMed
Grasso, D., Ottaviani, M. & Porcelli, F. 2001 Linear stability and mode structure of drift tearing modes. Phys. Plasmas 8, 4306.CrossRefGoogle Scholar
Grasso, D., Pegoraro, F., Porcelli, F. & Califano, F. 1999 Hamiltonian magnetic reconnection. Plasma Phys. Control. Fusion 41, 1497.CrossRefGoogle Scholar
Greene, J.M. 1993 Reconnection of vorticity lines and magnetic lines. Phys. Fluids B 5, 2355.CrossRefGoogle Scholar
Gu, Y.J., Pegoraro, F., Sasarov, P.V., Golovin, D., Yogo, A., Korn, G. & Bulanov, S.V. 2019 Electromagnetic burst generation during annihilation of magnetic field in relativistic laser-plasma interaction. Sci. Rep. 9, 19462.CrossRefGoogle ScholarPubMed
Hahm, T.S. 1988 Neoclassical tearing modes in a tokamak. Phys. Fluids 31, 37093712.CrossRefGoogle Scholar
Hahm, T.S. & Kulsrud, R.M. 1985 Forced magnetic reconnection. Phys. Fluids 28, 2412.CrossRefGoogle Scholar
Harris, E.G. 1962 On a plasma sheath separating regions of oppositely directed magnetic field. Il Nuovo Cimento 23, 115.CrossRefGoogle Scholar
Hasegawa, A. & Uberoi, C. 1982 The Alfvén Wave. Technical Information Centre, U.S. Departement of Energy.Google Scholar
Hautz, R. & Scholer, M. 1987 Numerical simulations on the structure of plasmoids in the deep tail. Geophys. Res. Lett. 14, 969.CrossRefGoogle Scholar
Hazeltine, R.D. & Ross, D.W. 1975 Kinetic theory of tearing instability. Phys. Fluids 18, 1778.CrossRefGoogle Scholar
Hesse, M., Forbes, T.G. & Birn, J. 1999 The diffusion region in collisionless magnetic reconnection. Phys. Plasmas 6, 1781.CrossRefGoogle Scholar
Hesse, M., Forbes, T.G. & Birn, J. 2005 On the relation between reconnected magnetic flux and parallel electric fields in the solar corona. Astrophys. J. 1227, 631.Google Scholar
Hicks, H.R., Carreras, B.A. & Holmes, J.A. 1984 The effect of diamagnetic rotation on the nonlinear coupling of tearing modes. Phys. Fluids 27, 909.CrossRefGoogle Scholar
Hofmann, I. 1974 Resistive tearing modes in a sheet pinch with shear flow. Plasma Phys. 17, 143.CrossRefGoogle Scholar
Holoborodko, P. 2012 Multiprecision Computing Toolbox for MATLAB.Google Scholar
Hones, E.W. 1979 Transient phenomena in the magnetotail and their relation to substorms. Space Sci. Rev. 23 (3), 393.CrossRefGoogle Scholar
Hoyle, F. 1949 Some recent Researches in Solar Physics. Cambridge University Press.Google Scholar
Huang, Y.-M. & Bhattacharjee, A. 2010 Scaling laws of resistive magnetohydrodynamic reconnection in the high-Lundquist-number, plasmoid-unstable regime. Phys. Plasmas 17, 062104.CrossRefGoogle Scholar
Janvier, M. 2017 Three-dimensional magnetic reconnection and its application to solar flares. J. Plasma Phys. 83, 535830101.CrossRefGoogle Scholar
Jardin, S.C., Krebs, I. & Ferraro, N. 2020 A new explanation of the sawtooth phenomena in tokamaks. J. Plasma Phys. 27, 032509.CrossRefGoogle Scholar
Kleva, R.G., Drake, J.F. & Waelbroeck, F.L. 1995 Fast reconnection in high temperature plasmas. Phys. Plasmas 2, 23.CrossRefGoogle Scholar
Knipp, D.J., Fraser, B.J., Shea, M.A. & Smart, D.F. 2018 On the little-known consequences of the 4 August 1972 ultra-fast coronal mass ejecta: facts, commentary, and call to action. Space Weath. 16, 1635.CrossRefGoogle Scholar
Kowal, G., Falceta-Gonçalves, D.A., Lazarian, A. & Vishniac, E.T. 2020 Kelvin–Helmholtz versus tearing instability: what drives turbulence in stochastic reconnection? Astrophys. J. 892, 50.CrossRefGoogle Scholar
Kumar, P., Karpen, J.T., Antiochos, S.K., Wyper, P.F. & DeVore, C.R. 2019 First detection of plasmoids from breakout reconnection on the Sun. Astrophys. J. Lett. 885, L15.CrossRefGoogle Scholar
Kuvshinov, B.N., Pegoraro, F. & Schep, T.J. 1994 Hamiltonian formulation of low-frequency nonlinear plasma dynamics. Phys. Lett. A 191, 296300.CrossRefGoogle Scholar
Kuvshinov, B.N. & Schep, T.J. 1997 Geometrical approach to fluid models. Phys. Plasmas 4, 537.CrossRefGoogle Scholar
Lütjens, H. & Luciani, J.-F. 2002 Linear and nonlinear thresholds of neoclassical tearing modes in tokamaks. Phys. Plasmas 9, 4837.CrossRefGoogle Scholar
Lakhina, G.S. & Tsurutani, B.T. 2016 Geomagnetic storms: historical perspective to modern view. Geosci. Lett. 3, 111.CrossRefGoogle Scholar
Landi, S., Del Zanna, L., Papini, E., Pucci, F. & Velli, M. 2015 Resistive magnetohydrodynamics simulations of the ideal tearing mode. Astrophys. J. 806, 131.CrossRefGoogle Scholar
Laval, G., Pellat, R. & Vuillemin, M. 1966 Instabiliteés electromagnetiques des plasmas sans collisions, vol. 2, p. 259. International Atomic Energy Agency.Google Scholar
Le, A., Egedal, J., Ohia, O., Daughton, W., Karimabadi, H. & Lukin, V.S. 2013 Regimes of the electron diffusion region in magnetic reconnection. Phys. Rev. Lett. 110, 135004.CrossRefGoogle ScholarPubMed
Leamon, R.J., Matthaeus, W.H., Smith, C.W., Zank, G.P., Mullan, D.J. & Oughton, S. 2000 MHD-driven kinetic dissipation in the solar wind and corona. Atsrophys. J. 537 (2), 1054.CrossRefGoogle Scholar
Li, J.H. & Ma, Z.W. 2012 Roles of super-Alfvenic shear flows on Kelvin–Helmholtz and tearing instability in compressible plasma. Phys. Scr. 86, 045503.CrossRefGoogle Scholar
Lin, J., Cranmer, S.R. & Farrugia, C.J. 2008 Plasmoids in reconnecting curent sheets: solar and terrestrial contexts compared. J. Geophys. Res. 113, A11107.Google Scholar
Longcope, D.W. & Strauss, H.R. 1993 The coalescence instability and the development of current sheets in two- dimensional magnetohydrodynamics. Phys. Scr. 5, 2858.Google Scholar
Loureiro, N.F., Cowley, S.C., Dorland, W.D., Haines, M.G. & Schekochihin, A. 2005 X-point collapse and saturation in the nonilnear tearing mode reconnection. Phys. Rev. Lett. 95, 235003.CrossRefGoogle ScholarPubMed
Loureiro, N.F. & Hammett, G.W. 2008 An iterative semi-implicit scheme with robust damping. J. Comput. Phys. 227, 45184542.CrossRefGoogle Scholar
Loureiro, N.F., Schekochihin, A. & Cowley, S.C. 2007 Instability of current sheets and formation of plasmoid chains. Phys. Plasmas 14, 0703.CrossRefGoogle Scholar
Loureiro, N.F. & Uzdensky, D.A. 2016 Magnetic reconnection: from the Sweet–Parker model to stochastic plasmoid chains. Mon. Not. Lett. 58, 014021.Google Scholar
Loureiro, N.F., Uzdensky, D.A., Schekochihin, A., Cowley, S.C. & Yousef, T.A. 2009 Turbulent magnetic reconnection in two dimensions. Mon. Not. Lett. 399, L146.Google Scholar
Lovelock, D. & Rund, H. 1989 Tensors, Differential Forms and Variational Principles. Dover Publ.Google Scholar
Lundquist, S. 1951 On the stability of magneto-hydrostatic fields. Phys. Rev. 83, 307.CrossRefGoogle Scholar
Lyutikov, M. 2003 Explosive reconnection in magnetars. Mon. Not. R. Astron. Soc. 346, 540.CrossRefGoogle Scholar
MacMahon, A. 1965 Finite gyro-radius corrections to the hydromagnetic equations for a Vlasov plasma. Phys. Fluids 8, 1804.CrossRefGoogle Scholar
Mahajan, S.M., Hazeltine, R.D., Strauss, H.R. & Ross, D.W. 1978 Collisionless “current-channel” tearing modes. Phys. Rev. Lett. 41, 1375.CrossRefGoogle Scholar
Mahajan, S.M., Hazeltine, R.D., Strauss, H.R. & Ross, D.W. 1979 Unified theory of tearing modes. Phys. Fluids 22, 2147.CrossRefGoogle Scholar
Mandt, M.E., Denton, R.E. & Drake, J.F. 1994 Transition to whistler mediated magnetic reconnection. Geophys. Res. Lett. 21 (1), 73.CrossRefGoogle Scholar
Matthaeus, W.H. & Lamkin, S.L. 1986 Turbulent magnetic reconnection. Phys. Fluids 29, 2513.CrossRefGoogle Scholar
Matthaeus, W.H. & Velli, M. 2011 Who needs turbulence? A review of turbulence effects in the Heliosphere and on the fundamental process of reconnection. Space Sci. Rev. 160, 145.CrossRefGoogle Scholar
Migliuolo, S., Pegoraro, F. & Porcelli, F. 1991 Stabilization of collisional drift-tearing modes at the breakdown of the constant-$\psi$ approximation. Phys. Fluids B 3, 1338.CrossRefGoogle Scholar
Militello, F., Borgogno, D., Grasso, D., Marchetto, C. & Ottaviani, M. 2011 Asymmetric tearing mode in the presence of viscosity. Phys. Plasmas 18, 112108.CrossRefGoogle Scholar
Militello, F., Huysmans, G., Ottaviani, M. & Porcelli, F. 2004 Effects of local features of the equilibrium current density profile on linear tearing modes. Phys. Plasmas 11, 125.CrossRefGoogle Scholar
Milosevich, G., Morrison, P.J. & Tassi, E. 2018 Direction of cascades in a magnetofluid model with electron skin depth and ion sound Larmor radius scales. Phys. Plasmas 25, 072303.CrossRefGoogle Scholar
Muraglia, M., Agullo, O., Benkadda, S., Garbet, X., Beyer, P. & Sen, A. 2009 Nonlinear dynamics of magnetic islands imbedded in small-scale turbulence. Phys. Rev. Lett. 103, 145001.CrossRefGoogle ScholarPubMed
Muraglia, M., Agullo, O., Benkadda, S., Yagi, M., Garbet, X. & Sen, A. 2011 Generation and amplification of magnetic islands by drift interchange turbulence. Phys. Rev. Lett. 107, 095003.CrossRefGoogle ScholarPubMed
National Research Council 2008 Severe Space Weather Events: Understanding Societal and Economic Impacts: A Workshop Report. The National Academies Press.Google Scholar
Nelson, G.A. 2016 Space radiation and human exposures, a primer. Radiat. Res. 185, 349.CrossRefGoogle ScholarPubMed
Newcomb, W.A. 1958 Motions of magnetic lines of force. Ann. Phys. 3, 347.CrossRefGoogle Scholar
Nishikawa, K.-I. 1982 Stabilizing effect of a normal magnetic field on the collisional tearing mode. Phys. Fluids 25, 1384.CrossRefGoogle Scholar
Nishizuka, N., Karlicky, M., Janvier, M. & Bárta, M. 2015 Particle acceleration in plasmoid ejections derived from radio drifting pulsating structures. Astrophys. J. 126, 799.Google Scholar
Ofman, L., Chen, X.L., Morrison, P.J. & Steinolfson, R.S. 1991 Resistive tearing mode instability with shear flow and viscosity. Phys. Fluids 3, 1364.CrossRefGoogle Scholar
Ottaviani, M., Arcis, N., Escande, D.F., Grasso, D., Maget, P., Militello, F., Porcelli, F. & Zwingmann, W. 2004 Progress in the theory of magnetic reconnection phenomena. Plasma Phys. Control. Fusion 46, B201.CrossRefGoogle Scholar
Ottaviani, M. & Porcelli, F. 1993 Nonlinear collisionless magnetic reconnection. Phys. Rev. Lett. 71, 3802.CrossRefGoogle ScholarPubMed
Ottaviani, M. & Porcelli, F. 1995 Fast nonlinear magnetic reconnection. Phys. Plasmas 2, 41044117.CrossRefGoogle Scholar
Papini, E., Franci, L., Landi, S., Hellinger, P., Verdini, A. & Matteini, L. 2019 a Statistics of magnetic reconnection and turbulence in Hall-MHD and hybrid-PIC simulations. Nuovo Cimento C 42, 23.Google Scholar
Papini, E., Landi, S. & Del Zanna, L. 2019 b Fast magnetic reconnection: secondary tearing instability and role of the hall term. Astrophys. J. 885, 56.CrossRefGoogle Scholar
Park, W., Monticello, D.A. & White, R.B. 1984 Reconnection rates of magnetic fields including the effects of viscosity. J. Geophys. Res. 27, 137.Google Scholar
Parker, E.N. 1957 Sweet's mechanism for merging magnetic fields in conducting fluids. J. Geophys. Res. 62, 509.CrossRefGoogle Scholar
Parker, E.N. 1973 The reconnection rate of magnetic fields. Astrophys. J. 180, 247.CrossRefGoogle Scholar
Pegoraro, F., Borgogno, D., Califano, F., Del Sarto, D., Echkina, E., Grasso, D., Liseikina, T. & Porcelli, F. 2004 Developments in the theory of collisionless reconnection in magnetic configurations with a strong guide field. Nonlinear Process. Geophys. 11, 567.CrossRefGoogle Scholar
Pegoraro, F., Kuvshinov, B.N., Romanelli, M. & Schep, T.J. 1995 Scale-invariant plasma motions near X-points. In Small-Scale Structures in Three-Dimensional Hydrodynamic and Magnetohydrodynamic Turbulence (ed. M. Meneguzzi, A. Pouquet & P.-L. Sulem), Lecture Notes in Physics, vol. 462, p. 295. Springer.CrossRefGoogle Scholar
Pegoraro, F., Porcelli, F. & Schep, T.J. 1989 Internal kink modes in the ion-kinetic regime. Phys. Fluids B 1, 364374.CrossRefGoogle Scholar
Pegoraro, F. & Schep, T.J. 1981 Low-frequency modes with high toroidal mode numbers: a general formulation. Phys. Fluids 24, 478.CrossRefGoogle Scholar
Pegoraro, F. & Schep, T.J. 1986 Theory of resistive modes in the ballooning representation. Plasma Phys. Control. Fusion 28, 647.CrossRefGoogle Scholar
Perona, A., Eriksson, L.-G. & Grasso, D. 2010 Electron response to collisionless magnetic reconnection. Phys. Plasmas 17, 042104.CrossRefGoogle Scholar
Petschek, H.E. 1964 Magnetic field annihilation. In AAS-NASA Symposium on the Physics of Solar Flares: Proceedings of a Symposium Held at the Goddard Space Flight Center, Greenbelt, Maryland, October 28–30, 1963, vol. 50, p. 425. Scientific and Technical Information Division, National Aeronautics and Space Administration.Google Scholar
Pneuman, G.W. 1983 Ejection of magnetic fields from the sun: acceleration of a solar wind containing diamagnetic plasmoids. Astrophys. J. 265, 468.CrossRefGoogle Scholar
Pogutse, I.O., Smolyakov, A.I. & Hirose, A. 1998 Magnetohydrodynamic equations for plasmas with finite-larmor-radius effects. J. Plasma Phys. 60, 133149.CrossRefGoogle Scholar
Porcelli, F. 1987 Viscous resistive magnetic reconnection. Phys. Fluids 30, 1734.CrossRefGoogle Scholar
Porcelli, F. 1991 Collisionless $m = 1$ tearing mode. Phys. Rev. Lett. 66, 425.CrossRefGoogle ScholarPubMed
Porcelli, F., Borgogno, D., Califano, F., Grasso, D., Ottaviani, M. & Pegoraro, F. 2002 Recent advances in collisionless magnetic reconnection. Plasma Phys. Control. Fusion 44, B389.CrossRefGoogle Scholar
Porcelli, F., Boucher, D. & Rosenbluth, M.N. 1996 Model for the sawtooth period and amplitude. Plasma Phys. Control. Fusion 38, 2163.CrossRefGoogle Scholar
Priest, E.R. 1976 Current sheet models of solar flares. Sol. Phys. 47, 41.CrossRefGoogle Scholar
Priest, E.R. & Sonnerup, B.U.Ö. 1975 Theories of magnetic field annihilation. Geophys. J. Intl 41, 405.CrossRefGoogle Scholar
Pritchett, P.L., Lee, Y.C. & Drake, J.F. 1979 Coalescence of magnetic islands. Phys. Fluids 22, 2140.CrossRefGoogle Scholar
Pritchett, P.L., Lee, Y.C. & Drake, J.F. 1980 Linear analysis of double tearing mode. Phys. Fluids 23, 1368.CrossRefGoogle Scholar
Pucci, F. & Velli, M. 2014 Reconnection of quasi-singular current sheets: the “ideal” tearing mode. Astrophys. J. Lett. 780, L19.CrossRefGoogle Scholar
Pucci, F., Velli, M., Shi, C., Singh, K.A.P., Tenerani, A., Alladio, F., Ambrosino, F., Buratti, P., Fox, W., Jara-Almonte, J., et al. 2020 Onset of fast magnetic reconnection and particle energization in laboratory and space plasmas. J. Plasma Phys. 86, 535860601.CrossRefGoogle Scholar
Pucci, F., Velli, M., Tenerani, A. & Del Sarto, D. 2018 Onset of fast “ideal” tearing in thin current sheets: dependence on the equilibrium current profile. Phys. Plasmas 25, 032113.CrossRefGoogle Scholar
Roberts, K.V. & Taylor, J.B. 1962 Magnetohydrodynamic equations for finite larmor radius. Phys. Rev. Lett. 8, 197.CrossRefGoogle Scholar
Rogers, B.N., Kobayashi, S., Ricci, P., Dorland, W., Drake, J. & Tatsuno, T. 2007 Gyrokinetic simulations of collisionless magnetic reconnection. Phys. Plasmas 14, 092110.CrossRefGoogle Scholar
Schekochihin, A.A. 2020 MHD turbulence: a biased review. arXiv:2010.00699.Google Scholar
Schep, T.J., Pegoraro, F. & Kuvshinov, B.N. 1994 Generalized two fluid theory of nonlinear magnetic structures. Phys. Plasmas 1, 28432852.CrossRefGoogle Scholar
Schnack, D.D. Jr. 1978 Non-linear numerical studies of the tearing mode. PhD thesis, Lawrence Livermore Laboratory, University of California, Livermore, CA.CrossRefGoogle Scholar
Schnack, D.D. Jr. 2009 Lectures in Magnetohydrodynamics. With An Appendix on Extended MHD. Springer.CrossRefGoogle Scholar
Scholer, M. 1987 Earthward plasma flow during near-earth magnetotail reconnection: numerical simulations. J. Geophys. Res. 92 (A11), 12425.CrossRefGoogle Scholar
Schouten, J.A. 1989 Tensors Analysis for Physicists. Dover Publ.Google Scholar
Servidio, S., Dmitruk, P., Greco, A., Wan, M., Donato, S., Cassak, P.A., Shay, M.A., Carbone, V. & Matthaeus, W.H. 2011 Magnetic reconnection as an element of turbulence. Nonlinear Process. Geophys. 18, 675.CrossRefGoogle Scholar
Servidio, S., Matthaeus, W.H., Shay, M.A., Cassak, P.A. & Dmitruk, P. 2009 Magnetic reconnection in two-dimensional magnetohydrodynamic turbulence. Phys. Rev. Lett. 102, 115003.CrossRefGoogle ScholarPubMed
Shay, M.A, Drake, J.F., Denton, R.E. & Biskamp, D. 1998 Structure of the dissipation region during collisionless magnetic reconnection. J. Geophys. Res. 103, 9165.CrossRefGoogle Scholar
Shibata, K. 1998 Evidence of magnetic reconnection in solar flares and a unified model of flares. Earth Planet. Space 264, 129.Google Scholar
Shibata, K. & Tanuma, S. 2001 Plasmoid induced-reconnection and fractal reconnection. Earth Planet. Space 53, 473.CrossRefGoogle Scholar
Singh, A., Pucci, F., Tenerani, A., Shibata, K., Hillier, A. & Velli, M. 2019 Dynamic evolution of current sheets, ideal tearing, plasmoid formation and generalized fractal reconnection scaling relations. Phys. Plasmas 881, 52.Google Scholar
Smolyakov, A.I., Pogutse, I.O. & Hirose, A. 1995 Fluid model of collisionless plasma with finite Larmor radius effects. Phys. Plasmas 2, 44514454.CrossRefGoogle Scholar
Somov, B.V. & Verneta, A.I. 1988 Magnetic reconnection in a high-temperature plasma of solar flares. III. Stabilizing effect of the transverse magnetic field in a non-neutral current sheet. Solar Phys. 117, 89.CrossRefGoogle Scholar
Somov, B.V. & Verneta, A.I. 1989 Magnetic reconnection in a high-temperature plasma of solar flares. IV. Resistive tearing mode in non-neutral current sheets. Solar Phys. 120, 93.CrossRefGoogle Scholar
Sonnerup, B.U.Ö. 1973 Magnetic field reconnection and particle acceleration. In High Energy Phenomena on the Sun Symposium Proceedings NASA-GSFC Doc. X-693-73-193, p. 357.Google Scholar
Strauss, H.R. 1976 Nonlinear three-dimensional magnetohydrodynamics of non-circular tokamaks. Phys. Rev. 19, 134.Google Scholar
Strauss, H.R. 1977 Dynamics of high-$\beta$ tokamaks. Phys. Rev. 20, 1354.Google Scholar
Strauss, H.R. 1988 Turbulent reconnection. Astrophys. J. 326, 412.CrossRefGoogle Scholar
Sweet, P.A. 1958 The neutral point theory of solar flares. In Electromagnetic Phenomena in Cosmical Physics. International Astronomical Union Symposium No. 6 (ed. B. Lehnert), p. 123. Cambridge University Press.CrossRefGoogle Scholar
Syrovatskii, S.I. 1966 a Dynamic dissipation of a magnetic field in particle acceleration. Sov. Astron. 10, 270.Google Scholar
Syrovatskii, S.I. 1966 b Dynamic dissipation of magnetic energy in the vicinity of a neutral magnetic field line. Sov. Phys. JETP 23, 754.Google Scholar
Syrovatskii, S.I. 1981 Pinch sheets and reconnection in astrophysics. Rev. Astron. Atrophys. 19, 163.CrossRefGoogle Scholar
Tajima, T. & Shibata, K. 1997 Plasma Astrophysics. Addison-Wesley.Google Scholar
Tanuma, S., Yokoyama, T., Kudoh, T. & Shibata, K. 2001 Two-dimensional magnetohydrodynamic numerical simulations of magnetic reconnection triggered by a supernova shock in the interstellar medium: generation of X-ray gas in the galaxy. Astrophys. J. 551, 312.CrossRefGoogle Scholar
Tassi, E., Sulem, P.-L. & Passot, T. 2016 Reduced models accounting for parallel magnetic perturbations: gyrofluid and finite Larmor radius–Landau fluid approaches. J. Plasma Phys. 82, 705820601.CrossRefGoogle Scholar
Tavani, M., Bulgarelli, A., Vittorini, V., Pellizzoni, A., Striani, E., Caraveo, P., Weisskopf, M.C., Tennant, A., Pucella, G., Trois, A., et al. 2011 Discovery of powerful gamma-ray flares from the Crab Nebula. Science 331 (6018), 736739.CrossRefGoogle ScholarPubMed
Tenerani, A. & Velli, M. 2020 Spectral signatures of recursive magnetic field reconnection. Mon. Not. R. Astron. Soc. 491, 4627.CrossRefGoogle Scholar
Tenerani, A., Velli, M., Pucci, F., Landi, S. & Rappazzo, A.F. 2016 ‘Ideally’ unstable current sheets and the triggering of fast magnetic reconnection. J. Plasma Phys. 82, 535820501.CrossRefGoogle Scholar
Tenerani, A., Velli, M., Rappazzo, A.F. & Pucci, F. 2015 The tearing mode instability of thin current sheets: the transition to fast reconnection in the presence of viscosity. Astrophys. J. 801, 145.CrossRefGoogle Scholar
Thompson, W.B. 1961 Magnetohydrodynamic equations for finite Larmor radius. Rep. Prog. Phys. 24, 363.CrossRefGoogle Scholar
Truesdell, C. 1950 The effect of the compressibility of the Earth in its magnetic field. Phys. Rev. 78, 823.CrossRefGoogle Scholar
Truesdell, C. 1954 The Kinematics of Vorticity. Indiana University Press.Google Scholar
Tur, A.V. & Yanovsky, V.V. 1993 Invariants in dissipationless hydrodynamic media. J. Fluid Mech. 248, 67.CrossRefGoogle Scholar
Uzdensky, D.A. 2011 Magnetic reconnection in extreme astrophysical environments. Space Sci. Rev. 160, 45.CrossRefGoogle Scholar
Uzdensky, D.A., Loureiro, N.F. & Schekochihin, A.A. 2010 Fast magnetic reconnection in the plasmoid-dominated regime. Geophys. Res. Lett. 105, 235002.Google ScholarPubMed
Vaivads, A., Khotyaintsev, Y., André, M., Retino, A., Buchert, S.C., Rogers, B.N., Décréau, P., Paschmann, G. & Phan, T.D. 2004 Structure of the magnetic reconnection diffusion region from four-spacecraft observations. Phys. Rev. Lett. 93, 105001.CrossRefGoogle ScholarPubMed
Van Hoven, G. & Cross, M.A. 1971 Resistive instability of a sheet pinch. Phys. Fluids 14, 1141.CrossRefGoogle Scholar
Van Hoven, G. & Cross, M.A. 1973 a Energy release by magnetic tearing: the nonlinear limit. Phys. Rev. A 7, 1347.CrossRefGoogle Scholar
Van Hoven, G. & Cross, M.A. 1973 b Inertial magnetic field reconnection and magnetospheric substorm. Phys. Rev. Lett. 30, 642.CrossRefGoogle Scholar
Vasyliunas, V.M. 1975 Theoretical models of magnetic field line merging. Rev. Geophys. 13, 303.CrossRefGoogle Scholar
Velli, M. & Hood, A.W. 1989 Resistive tearing in line-tied magnetic fields: slab geometry. Solar Phys. 119, 107.CrossRefGoogle Scholar
Waelbroeck, F.L., Hazeltine, R.D. & Morrison, P.J. 2009 A Hamiltonian electromagnetic gyrofluid model. Phys. Plasmas 16, 032109.CrossRefGoogle Scholar
Waelbroeck, F.L. & Tassi, E. 2012 A compressible Hamiltonian electromagnetic gyrofluid model. Commun. Nonlinear Sci. Numer. Simul. 17, 21712178.CrossRefGoogle Scholar
Wan, M., Matthaeus, W.H., Servidio, S. & Oughton, S. 2013 Generation of X-points and secondary islands in magnetohydrodynamic turbulence. Phys. Plasmas 20, 042307.CrossRefGoogle Scholar
Wang, X. & Bhattacharjee, A. 1997 Forced reconnection and mode locking in rotating cylindrical plasmas. Phys. Plasmas 4, 748.CrossRefGoogle Scholar
Wang, X. & Bhattcharjee, A. 1995 Nonlinear dynamics of the $m=1$ kink-tearing instability in a modified magnetohydrodynamic model. Phys. Plasmas 2, 171.CrossRefGoogle Scholar
Wang, S., Lee, L.C. & Wei, C.Q. 1988 Streaming tearing instability in the current sheet with a super-Alfvénic flow. Phys. Fluids 31, 1544.CrossRefGoogle Scholar
Wang, Z.-X., Wei, L., Wang, X., Zheng, S. & Liu, Y. 2011 Fast linear growth of collisionless double tearing modes in a cylindrical plasma. Nucl. Fusion 51, 033003.CrossRefGoogle Scholar
Watson, P.G. & Craig, I.J.D. 1998 Magnetic annihilation and reconnection in two dimensions. Astrophys. J. 505, 363.CrossRefGoogle Scholar
Wei, F., Hu, Q., Feng, X. & Schwen, R. 2000 Simulation of turbulent magnetic reconnection in the small-scale solar wind. Sci. China A 43, 629.CrossRefGoogle Scholar
Wesson, J.A. 1986 Sawtooth oscillations. Plasma Phys. Control. Fusion 28, 243.CrossRefGoogle Scholar
Wesson, J.A. 1990 Sawtooth reconnection. Nucl. Fusion 30, 2545.CrossRefGoogle Scholar
Wesson, J. 2004 Tokamaks. Oxford University Press.Google Scholar
White, R.B. 1983 Resistive instabilities and field line reconnection. In Handbook of Plasma Physics, Vol. I (ed. A.A. Galeev & R.N. Sudan), p. 611. North Holland Publisher.Google Scholar
White, R.B. 1986 Resistive reconnection. Rev. Mod. Phys. 58, 183.CrossRefGoogle Scholar
Wilson, H.R. 2012 Neoclassical tearing modes. Fusion Sci. Technol. 61, 113121.CrossRefGoogle Scholar
Woltjer, L. 1958 A theorem on force-free magnetic fields. Proc. Natl Acad. Sci. 44, 489.CrossRefGoogle ScholarPubMed
Yamada, M., Ji, H., Hsu, S., Carter, T., Kulsrud, R., Ono, Y. & Perkins, F. 1997 Identification of Y-shaped and O-shaped diffusion regions during magnetic reconnection in a laboratory plasma. Phys. Rev. Lett. 78, 3117.CrossRefGoogle Scholar
Yamada, M., Kulsrud, R. & Ji, H. 2010 Magnetic reconnection. Rev. Mod. Phys. 82, 603.CrossRefGoogle Scholar
Yan, X., Xue, Z., Jiang, C., Priest, E.R., Kliem, B., Yang, L., Wang, J., Kong, D., Song, Y., Feng, X. et al. 2022 Fast plasmoid-mediated reconnection in a solar flare. Nat. Commun. 13, 640.CrossRefGoogle Scholar
Yu, Q. 2010 Linear and nonlinear stability of drift-tearing mode. Nucl. Fusion 50, 025041.CrossRefGoogle Scholar
Zakharov, L. & Rogers, B. 1992 Two-fluid magnetohydrodynamic description of the internal kink mode in tokamaks. Phys. Fluids B 4 (10), 32853301.CrossRefGoogle Scholar
Zank, G.P. & Matthaeus, W.H. 1992 The equations of reduced magneto- hydrodynamics. J. Plasma Phys. 48, 85.CrossRefGoogle Scholar
Zhou, M., Wu, D.H., Loureiro, N.F. & Uzdensky, D.A. 2021 Statistical description of coalescing magnetic islands via magnetic reconnection. J. Plasma Phys. 87, 905870620.CrossRefGoogle Scholar
Zocco, A., Helander, P. & Weitzner, H. 2020 Magnetic reconnection in 3d fusion devices: non-linear reduced equations and linear current-driven instabilities. Plasma Phys. Control. Fusion 63, 025001.CrossRefGoogle Scholar
Zocco, A. & Schekochihin, A.A. 2011 Reduced fluid-kinetic equations for low-frequency dynamics, magnetic reconnection, and electron heating in low-beta plasmas. Phys. Plasmas 18 (10), 102309.CrossRefGoogle Scholar
Figure 0

Figure 1. Cartoon schematising three steps of a simplified 2-D magnetic reconnection process that emphasises the notion of change of magnetic topology. (a) Two magnetic lines connecting different pairs of fluid elements, A $\leftrightarrow$ B and C $\leftrightarrow$ D, are twisted by the ideal fluid motion so as to locally generate strong spatial gradients, i.e. current densities. (b) Magnitude of the spatial gradients locally compensates the smallness of the non-ideal parameters in the generalised Ohm's law, by thus allowing the relaxation of the topological constraints that in the ideal limit, forbid the intersection of initially distinct magnetic lines; an $X$-point is formed. (c) Magnetic reconnection process undergone at the $X$-point, by which magnetic energy is dissipated and/or converted into other forms of plasma energy, which makes other magnetic configurations accessible to the plasma; the magnetic topology has globally changed as the magnetic lines now connect the pairs of fluid elements A $\leftrightarrow$ D and B $\leftrightarrow$ C.

Figure 1

Figure 2. Comparison of four reconnection scenarios. (a) Sketch of the steady Sweet–Parker reconnection scenario. No particular hypotheses are done on the boundary conditions of the reconnecting current sheet, whereas steadiness and a specific scaling of the aspect ratio with non-ideal parameters like resistivity (i.e. $L/a\sim S^{-1}$, Parker 1957; Sweet 1958) or electron inertia (Wesson 1990) are required. The configuration does not represent an instability: no distinction is made between equilibrium and perturbed quantities. (b) Typical ‘magnetic island’ configuration in the tearing-type reconnection. Periodic boundary conditions along the neutral line are required and hyperbolic patterns of the velocity field near the $X$- and $O$-points develop together with the island formation. The wavenumber of oscillations of the mode fixes the number of magnetic islands which are generated. In the linear stage, the island width is $w\simeq \delta \ll a$. (c,d) Merger configuration (c) of two undisplaced line currents (Syrovatskii 1966a; Biskamp & Welter 1980; Kleva et al.1995), which resembles (d) a non-periodic version of the ‘coalescence instability’ scenario. The latter concerns a periodic structure constituted by magnetic islands in a chain (Finn & Kaw 1977; Pritchett, Lee & Drake 1979; Bhattacharjee, Brunel & Tajima 1983). In the displaced current merger case, the attraction of two parallel current filaments squeezes the magnetic field that piles up and typically dissipates at the $X$-point: in this configuration, one may rather speak of ‘magnetic annihilation’ (cf. e.g. Priest & Sonnerup 1975; Watson & Craig 1998; Gu et al.2019). In the coalescence instability scenario, a somewhat similar process occurs (an out-of-plane current density being associated with the $O$-point of each island) once a non-rigid displacement along the neutral line perturbs the periodic configuration by making the islands get closer, pairwise. Similarly, a periodic 2-D ‘paving’ of magnetic cell-like structures of different shape (not shown here) can be destabilised the same way by planar, non-rigid displacements (Longcope & Strauss 1993).

Figure 2

Figure 3. Growth rates of collisionless tearing modes numerically computed from (2.1)–(2.2) with the solver of Betar et al. (2020), for values of the non-ideal parameters at the limits of (or beyond the) applicability of boundary layer theory, and for a wavelength $ka\sim O(1)$ (likely value for modes on large aspect ratio current sheets having a microscopic thickness – a condition typically met, e.g. in MHD turbulence (see, e.g. Franci et al.2016; Del Sarto & Pegoraro 2017; Franci et al.2022). Note that once $\gamma$ approaches unity, the effect of the non-ideal parameters on the magnetic equilibrium $\psi _0$ is not negligible anymore, and (2.1) and (2.2) in the resistive limit (not shown here) should be modified accordingly (cf. Appendix A). (a) Case of cold electrons. (b) Case of warm electrons, for different values of $d_e/a$ and for values of $\rho _s/a$ approaching one.

Figure 3

Figure 4. (a) Sketch of the boundary layer decomposition for the calculation of the tearing eigenmodes with asymptotic matching techniques, when a single boundary layer is sufficient (e.g. case of the resistive tearing first considered by Furth et al.1963). The width $\delta$ of the inner boundary layer is here the natural candidate for the width of the ‘reconnecting layer’, although the latter is not formally defined, i.e. we can assume $\delta =\delta _1$ (an assumption we will support in § 8 with further arguments). (b) Sketch of the boundary layer decomposition when two boundary layers are required (e.g. case of the warm-collisionless RMHD). Note that the reconnecting layer width, which we call $\delta$, is, again, a priori not defined but, as we will see (§ 8), further arguments can be found to identify it with the innermost layer width, i.e. $\delta =\delta _1$.

Figure 4

Figure 5. Scheme of the correspondence between linear reconnecting instabilities in the RMHD slab geometry limit and in the RMHD ‘cylindrical tokamak’ approximation, which is compatible with the strong guide field assumption (Strauss 1976). At the bottom of the figure, we highlight the fact that the slab geometry easily allows modelling both of tearing modes of different wavenumber (centre frame) and of the double tearing mode, i.e. the tearing-like instability simultaneously occurring on two sufficiently close resonant surfaces (Furth et al.1973), of which we have shown here only the symmetric case (rightmost frame). Note however that we have considered here the ‘cylindrical tokamak’, or ‘large aspect ratio tokamak’ approximation, in the strict limit of $\partial / \partial \varphi \to 0$, which maps into the $\partial / \partial z \to 0$ assumption that we consider in this article. In reality, the RMHD modelling also allows for inclusion of the neglected derivatives with respect to the axis-symmetric coordinate ordered as $\partial / \partial \varphi \sim \partial / \partial z \sim \varepsilon _B$, in terms of the ratio $\varepsilon _B$ between the in-plane and the guide magnetic field components at equilibrium (cf. (A11) and discussion in the last paragraph of Appendix A).

Figure 5

Figure 6. Dispersion relation of the purely resistive tearing as a function of $ka$ (dots correspond to numerically computed values), evaluated for the equilibrium $\psi _0=\cos (x/a)$. The smaller box is a ‘zoom’ of the plot for low values of $\gamma$, which correspond to the range of smaller values of $S^{-1}$ indicated in the caption. The dashed line corresponds to the fastest mode scaling, which for $\psi _0=\cos (x/a)$ is $\gamma _M\sim S^{-1/2}$. This can be obtained by using in (24) of Del Sarto et al. (2016) the value $p=2$, which can be deduced from the $ka\ll 1$ limit of the corresponding $\varDelta '(ka)$ formula (see Ottaviani & Porcelli 1993 or (8) of Ottaviani & Porcelli 1995). An 8th-order accurate scheme has been used here for integration on a uniform grid by using the solver discussed by Betar et al. (2020).

Figure 6

Table 1. Known scalings available in the previous literature as obtained from a boundary layer analysis in the collisionless and resistive regimes. The general dispersion relation of Porcelli (1991) and Ottaviani & Porcelli (1995) for $d_e^2\gg \rho _s^2$ and that of Porcelli (1991) and Comisso et al. (2013) for $\rho _s>d_e\neq 0$ are reported. In both cases, we have adopted the notation used in Del Sarto & Ottaviani (2017, appendix C), from which the small and large wavelength limits can be obtained once $\delta _L\to \delta$ is assumed and some estimation is made for $\delta =\delta _{LD}$ or $\delta =\delta _{SD}$. The general dispersion relation valid in all wavelength limits of the cold-electron resistive regimes (Ara et al.1978) and of the warm-electron resistive regimes (Pegoraro & Schep 1986) is not reported explicitly but it can be obtained from the collisionless cases previous substitution $d_e\to (S\gamma )^{-1/2}$. Note that, although available since the early solutions of the boundary layer calculations, the scalings of $\delta _{LD}$ and of $\delta _{SD}$ we have reported here in the different regimes have not been always pointed out in the related, reference works. These scalings correspond indeed to the characteristic width of the innermost boundary layer (i.e. $\delta _1$) in each reconnection regime, but the identification of the latter as the reconnecting layer width requires indeed further hypotheses which are discussed in § 8.

Figure 7

Figure 7. Profile of the eigenfunction $\psi _1$ in the purely resistive regime, computed via the numerical solver of Betar et al. (2020) for $S^{-1}=10^{-5}$ and $k=10$ (green curve), $k=1.7$ (orange curve) and $k=0.1$ (blue curve). These three values are respectively exemplificative of: the stable regime ($\varDelta '=-19.19<0$, green curve); of the small-$\varDelta '$ wavelength limit ($\varDelta '=3.2766$, red curve); and of the large-$\varDelta '$ wavelength limit ($\varDelta '=48.9$, blue curve). In the stable case ($\varDelta '<0$), the eigenfunction does not change concavity as $x$ approaches the neutral line, whereas for the unstable tearing-type solutions obtained for $\varDelta '>0$, the value $c_0\equiv \lim _{x\to 0}\psi _1(x)$ decreases approaching zero as the value of $\varDelta '(k)$ increases approaching $+\infty$.

Figure 8

Figure 8. Profiles of $\psi _1$ (blue), $\chi _{\text {in}}^{\text {appr}}$ of (7.46) (black-dashed) and $x \psi _1'$ (orange) in the reconnection layer around the neutral line: these profiles are obtained by numerical integration of the complete eigenvalue problem and should be compared to the approximated analytical solutions ((7.57) and (7.58)). The width of reconnection layer, evaluated as it will be specified in § 7, is $\delta = \delta _1 \simeq 10^{-6}$. The physical parameters of the numerical integration performed with the solver of Betar et al. (2020) in the large-$\varDelta '$ limit are shown in the figure.

Figure 9

Figure 9. Spatial profile of $\psi _1$, numerically computed with the eigenvalue solver of Betar et al. (2020) in the collisionless regime for a choice of the parameters $d_e=2\times 10^{-2}$, $\rho _s=5\times 10^{-3}$, and for a value of $k=1.7$. These values correspond to $\varDelta '(k)=3.3$ and to $\delta =\delta _1\simeq 2.4\times 10^{-6}$, evaluated according to the definition in (8.3) introduced next. The smaller box inserted at the top right-hand corner of the figure shows a zoom of the solution inside the innermost layer of width $\sim \delta =\delta _1$, in which the appropriateness of the constant-$\psi _1$ condition is evidenced.

Figure 10

Figure 10. Comparison of the spatial profiles of the terms $\rho _s^2/(\bar {d}_e^2\tilde {\mathcal {G}}^2)\times \tilde {\zeta }^2\tilde {\varphi }_1''$ (blue curve), of $\rho _s^2/(\bar {d}_e^2\tilde {\mathcal {G}}^2)\times \tilde {\zeta }^2\tilde {\varphi }_1$ (orange curve) and of $\tilde {\varphi }_1$ (green curve) of (7.60), in a region of the domain with width $|x|<\delta _2=\rho _s$. The parameters of the numerical integration performed with the eigen-solver of Betar et al. (2020) in the small-$\varDelta '$ limit are shown at the top of the figure. The value of $\delta =\delta _1$ in the bottom right corner has been evaluated according to the definition in (8.3).

Figure 11

Figure 11. Sketch of the hyperbolic pattern of the velocity field associated with linear perturbations around the X-point. The yellow stripe represents the reconnecting layer in the vicinity of the neutral line, where a current sheet $J_{z,1}$ of width $\sim \delta$ is concentrated (the width of the magnetic island, instead, may well trespass the width of this region, during its nonlinear evolution – cf. caption of figure 2). Note that the estimate $v_{x,1}\sim \gamma \delta$, with $v_{x,1}=-{\rm i}k\varphi _1$, holds inside of this region, whereas the $v_{y,1}=\varphi _1'$ component exhibits a change of sign on the neutral line and therefore displays a strong gradient across the $\varDelta x\simeq \delta$ interval.

Figure 12

Figure 12. Examples of evaluation of the half-width of the current layer as the distance of the local maxima (minima) of $\psi _1'''=-J_{z,1}'$ (a) and of $\varphi _1''$ (b) from the neutral line (the parameters of the numerical calculation are indicated at the top of the panels).

Figure 13

Figure 13. Scalings of $\delta _\varphi$ (a,c) and $\delta _\psi$ (b,d) calculated using the definitions given by (8.1). The scalings, shown on the different panels, prove that these two quantities follow the same scaling laws.

Figure 14

Figure 14. (a) Example of numerical evaluation of the coefficients $c_0$ and $c_1$ of (10.10), once the profile of $\psi _1(x)$ and of its derivatives $\psi _1'(x)$ have been computed for $|x|\geq \delta$ in a neighbourhood of $x=\delta$. (b) Analogous example of evaluation of $q_0$ and $q_1$ of (10.10) from $\varphi _1'$ and $\varphi _1''$, calculated once the eigenfunction $\varphi _1$ has been computed; in this example, $\varDelta '_{v_y}\approx 73.82$. (c) A zoomed version of panel (b) in which the inner region, here of width $\delta \approx 0.00184$ (computed according to the method outlined in § 8), is shaded in light-blue colour. All the values required to estimate $\varDelta '_{v_y}$ using (10.2) are here shown; it is this way that $\varDelta '_{v_y}\approx 74.48$ is obtained, in very good agreement with the result computed with the alternative method sketched in panel (b).

Figure 15

Table 2. Scaling of the spatial scales associated with the second- and fourth-order derivative of $\psi _1(x)$, evaluated on the neutral line $x=0$ in different reconnection regimes and wavelength limits. In each regime, $\delta _{SD}$ and $\delta _{LD}$ are those of table 1 and correspond to the different values of $\delta _1$ evaluated in §§ 6 and 7.

Figure 16

Figure 15. Examples of scaling laws of $D'$, $\varDelta '_{v_y}$ (cf. (10.10)) and $\delta$ with respect to $d_e$, for $\rho _s= 0$. The large-$\varDelta '$ limit is in panel (a), the small-$\varDelta '$ in panel (b). The values of $D'$ and $\varDelta '_{v_y}$ are computed numerically according to the procedure sketched in figure 14. In the small-$\varDelta '$ limit, the values of $\varDelta '$, obtained from (4.13), coincide with the orange line in panel (b).

Figure 17

Figure 16. Examples of scaling laws of $D'$, $\varDelta '_{v_y}$ (cf. (10.10)) and $\delta$ for $\rho _s\neq 0$. Panels (a,b) correspond to the large-$\varDelta '$ limit, whereas panels (c,d) correspond to the small-$\varDelta '$ limit. The values of $D'$ and $\varDelta '_{v_y}$ are computed numerically according to the procedure sketched in figure 14. Notice that $D'$ and $\delta$ are independent of $\rho _s$ in the small-$\varDelta '$ limit (cf. panel d). In the small-$\varDelta '$ limit, the values of $\varDelta '$ obtained from (4.13) coincide with the blue lines in panels(c,d).

Figure 18

Figure 17. The region where the solutions $\psi _\alpha$ and $\psi _\beta$ diverge is here displayed in a grey colour inside two lobes that join each other at $\zeta =0$. Outside these two lobes, both series converge.

Figure 19

Table 3. Starting equations, definitions and assumptions based on general constraints that we take for the heuristic-type approach in all reconnection regimes. Equations [T4.E1–T4.E2] are the eigenmode equations ((2.1) and (2.2)) in which we have assumed $k$ to be negligible with respect to the inverse scale of the spatial gradients of the eigenfunctions. Equations [T4.D1–T4.D3] are the operational definitions we have given in §§ 8, 9 and 10 of $\delta$, $l_c$ and $\varDelta _{v_y}'$, respectively. Equations [T4.C1,T4.C2] and [T4.H1] represent further conditions on the parameters that must be fulfilled regardless of the reconnection regime: [T4.C1] is a constraint that follows from [T4.E2] and definitions [T4.D2–T4.D3]; conditions [T4.C2a] and [T4.C2] correspond to (10.9) that we have discussed in the text; condition [T4.H1] corresponds to the ansatz (10.5) discussed in the text. Equations [T4.R1a] and [T4.R1b] are two immediate conclusions than can be drawn and allow us to specify hypothesis [T4.H1] in the small- and large-$\varDelta '$ limits, respectively.

Figure 20

Table 4. Combining the hypotheses and results reported in table 2 with hypotheses [T5.H2–T5.H3] that are specific of the warm regimes, it is possible to deduce in a self-consistent way the set of equations [T5.R2–T5.R5]. It must be emphasised, however, that the system of condition is not closed in the large-$\varDelta '$ regime and additional information about $\varDelta _{v_y}'$ is required to obtain the corresponding scaling laws.

Figure 21

Table 5. Combining the hypotheses and results reported in table 2 with hypotheses [T6.H2] and [T6.H3] that are specific of the cold regimes, it is possible to deduce in a self-consistent way the set of equations [T6.C3]–[T6.R4]. Note that, in this case, the set of conditions is closed also in the large-$\varDelta '$ regime thanks to [T6.H2], which leads to the constraint [T6.C4] that sets the scaling of $\varDelta _{v_y}'$ to be the same of $\delta ^{-1}$. All scalings can be this way a priori determined without resorting to numerical calculations.