Open Access
Issue
A&A
Volume 604, August 2017
Article Number A62
Number of page(s) 9
Section The Sun
DOI https://doi.org/10.1051/0004-6361/201629863
Published online 07 August 2017

© ESO, 2017

Licence Creative CommonsOpen Access article, published by EDP Sciences, under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

1. Introduction

An axially symmetric, vertical and magnetically twisted flux tube is a convenient model for analytical studies of various magnetohydrodynamic (MHD) perturbations. For a long time this approximation was the focus of investigations of MHD wave propagation in solar and space plasmas (see e.g. Priest & Hood 1991; Bennett et al. 1999; Erdélyi & Fedun 2007; Ruderman 2007; Ladikov-Roev et al. 2013; Cheremnykh et al. 2014) and high-temperature (see e.g. Suydam 1958; Bateman 1978; Galeev & Sudan 1989; Cheremnykh 1989; Burdo et al. 1994) plasmas. This geometry is also a useful approximation in solving fundamental problems of plasma physics (see for example Trubnikov 1966; Filippov 2007; Cheremnykh et al. 1994; Zagorodny & Cheremnykh 2014), to name but a few. In spite of many previous theoretical studies of wave propagation in solar magnetic flux tubes many questions still remain open. There are at least two contradictory opinions on how the radial dependence of the equilibrium azimuthal component of the magnetic field outside of the flux tube should be modelled. Filippov (2007) and Vršnak (2008) assume that external magnetic field decreases with distance from the tube boundary inversely proportional to the radius, that is, as a function of 1 /r. This approximation was previously used by, for example, Erdélyi & Fedun (2006), Ruderman (2015), Giagkiozis et al. (2015). Recently, Giagkiozis et al. (2015) has shown that the wave solution for a background constant twist outside the tube is actually very close to the solution when the twist is proportional to 1 /r. From another point of view, by taking into account plasma conductivity, the external magnetic field does not penetrate significantly through the tube boundary and, therefore, can be neglected (see e.g. Parker 2007; Solov’ev 2011, 2012). In these papers authors applied simple electromagnetic assumptions confirming the absence of azimuthal components of the magnetic field outside the tube.

In the present work, we will examine MHD wave propagation in a magnetic flux tube with an internal twist only. To go beyond cold plasma equilibria conditions the axially aligned magnetic field inside and outside the flux tube are allowed to be different. A similar background model has been used previously by Bennett et al. (1999) and thereafter in a number of other papers, for example Erdélyi & Fedun (2007, 2010). In the framework of ideal MHD, we will assume incompressible linear perturbations and implement the thin tube approximation. Also, we will focus on the analytical solutions related to modes with only m ≥ 1. The analytical dispersion relation and expression for eigenfunctions will be obtained by assuming the small parameter ε = kza ≪ 1, where a is the radius of the magnetic flux tube and kz is the longitudinal wavenumber.

2. Derivation of linear radial component wave equation

We proceed from the linearised ideal MHD equations for the displacement vector ξ of a finite volume element. By assuming the time dependence of all perturbed physical quantities as exp(−iωt), these equations can be written as (see e.g. Priest 2014; Kadomtsev 1966; Zagorodny & Cheremnykh 2014) ρ2ξt2=ρω2ξ=F(ξ),\begin{equation} \rho\frac{\partial^2\bfxi}{\partial t^2}=-\rho\omega^2\bfxi=\vec{F}\left(\bfxi \right), \label{a1} \end{equation}(1)where

F(ξ)=δp1+(B·)δB+(δB·)B,δp1=δp+B·δB=γp·ξB2(·ξ+2κ·ξ),δB=×[ξ×B],·δB=0,κ=(τ·)τ,τ=B/B.\begin{eqnarray} &&\vec{F}\left(\vec{\bfxi}\right)=-\bfbigtriangledown\delta p_1+\left(\vec{B}\cdot\vec{\nabla}\right)\delta\vec{B}+ \left(\delta\vec{B}\cdot \vec{\nabla}\right)\vec{B}, \nonumber \\ &&\delta p_1 = \delta p +\vec{B}\cdot\delta \vec{B} = -\gamma p \vec{\nabla}\cdot \bfxi-B^2\left(\vec{\nabla}\cdot \bfxi_{\perp} +2 \bfkappa\cdot \bfxi_{\perp}\right), \nonumber \\ &&\delta p =-\bfxi \cdot \bfbigtriangledown p - \gamma p \vec{\nabla}\cdot \bfxi, \label{a2} \\ &&\delta\vec{B}=\bfbigtriangledown \times \left[\bfxi \times \vec{B}\right], \nonumber \\ &&\vec{\nabla}\cdot\delta\vec{B}=0, \nonumber \\ &&\bfkappa=\left(\bftau \cdot \bfbigtriangledown\right)\bftau, \nonumber \\ &&\bftau=\vec{B}/B. \nonumber \end{eqnarray}(2)Here, the symbol δ corresponds to the perturbed quantities, ρ is the equilibrium plasma density, p is the equilibrium plasma pressure, ξ = ξrer + ξϕeϕ + ξzez is the displacement vector (where er, eϕ and ez are unit vectors of cylindrical coordinates r, ϕ and z), γ is the adiabatic index, ω is the angular frequency, B is the equilibrium magnetic field, τ is the normalised magnetic field, δp1 is the perturbation of total plasma pressure (plasma plus magnetic), and κ is the vector of curvature of magnetic field lines. The derivation of the expression for δp1 is shown in Appendix A. ξ = ξξτ in Eq. (2) corresponds to the perpendicular component of ξ to the equilibrium magnetic field. All physical quantities depend on the radial coordinate r. For simplicity, the magnetic field is normalised as B/4πB\hbox{$\vec{B}/\sqrt{4\pi}\rightarrow \vec{B}$}. We use the cylindrical coordinate system (r,ϕ,z), hence the magnetic surfaces are nested cylinders of radius r and the unperturbed tube axis is parallel to the z-axis (see Fig. 1). Both inside and outside the tube the equilibrium magnetic field is given by B=Bϕ(r)eϕ+Bz(r)ez,\begin{equation} \vec{B}=B_{\varphi}(r)\vec{e}_\varphi+B_z(r)\vec{e}_z, \label{a3} \end{equation}(3)which must satisfy the magneto-hydrostatic equilibrium: ddr(p+B22)+Bϕ2r=0.\begin{equation} \frac{{\rm d}}{{\rm d}r}\left(p+\frac{B^2}{2}\right)+\frac{B_{\varphi}^2}{r}=0. \label{a4} \end{equation}(4)Using the relation κ=(BB·)BB=err(BϕB)2,\begin{eqnarray*} \bfkappa=\left(\frac{\vec{B}}{B}\cdot \bfbigtriangledown\right)\frac{\vec{B}}{B}=-\frac{\vec{e}_r}{r}\left(\frac{B_{\varphi}}{B}\right)^2, \end{eqnarray*}and the rules for differentiating unit vectors, er/∂ϕ = eϕ and eϕ/∂ϕ = −er, Eqs. (1) and (2) reduce to:

\begin{eqnarray} &&\label{a5}\rho\omega^2\xi_{r}-\frac{{\rm d}}{{\rm d}r}\delta p_1 -\frac{2B_{\varphi} \delta B_{\varphi}}{r}+{\rm i}\left(\frac{m}{r}B_{\varphi}+k_{z}B_{z}\right)\delta B_r=0,~~~~~~~~~~~~~~~~ \\ &&\label{a6}\rho\omega^2\xi_{\varphi}-\frac{{\rm i}m}{r}\delta p_1 +\frac{\delta B_{r}}{r}\frac{{\rm d}}{{\rm d}r}\left(r B_{\varphi}\right)+{\rm i}\left(\frac{m}{r}B_{\varphi}+k_{z}B_{z}\right)\delta B_{\varphi}=0, \\ &&\label{a7}\rho\omega^2\xi_{z}-{\rm i}k_{z}\delta p_{1} +\delta B_{r}\frac{{\rm d}}{{\rm d}r}B_{z}+{\rm i}\left(\frac{m}{r}B_{\varphi}+k_{z}B_{z}\right)\delta B_{z}=0. \end{eqnarray}To obtain Eqs. (5)(7) we assumed that all equilibrium quantities depend on r alone. Therefore, we can Fourier decompose the solution as ξ(r,t)=ξ(r)expi(ωt++kzz),\begin{eqnarray*} \bfxi\left(\vec{r}, t\right)=\bfxi(r)\exp {\rm i} \left(-\omega t+m\varphi +k_z z\right), \nonumber \end{eqnarray*}where m is the azimuthal wave number. In this study we shall not deal with the axi-symmetric case when m = 0, which describes the sausage mode. Instead, we focus on all the modes with m ≥ 1, corresponding to non-axially symmetric oscillations which are the kink m = 1 and surface m > 1 modes. A cartoon of the model geometry is shown in Fig. 1.

thumbnail Fig. 1

Magneto-hydrostatic equilibrium of the twisted magnetic flux tube.

Since the displacement vector ξ(r) is a function of the radius alone, then the problem becomes one-dimensional. For more convenient analysis of the perturbations in the Eqs. (5)(7) we changed the ϕ and z components of the displacement vector ξ and wave vector k to the components directed along the bi-normal (subscript b) and along the magnetic field lines (subscript ):

ξ=ξrer+ξbeb+ξe,ξ=ξϕBϕB+ξzBzB,e=eϕBϕB+ezBzB,kb=mrBzBkzBϕB,k=k·e=mrBϕB+kzBzB·\begin{eqnarray} && \bfxi=\xi_r \vec{e}_r+\xi_{b} \vec{e}_{b}+\xi_\parallel \vec{e}_\parallel, \nonumber \\ &&\xi_{\rm b}=\xi_{\varphi}\frac{B_z}{B}-\xi_z\frac{B_{\varphi}}{B}, \quad \vec{e}_{b}=\vec{e}_\varphi \frac{B_z}{B}-\vec{e}_z\frac{B_{\varphi}}{B}, \label{a10} \\ && \xi_{\parallel}=\xi_{\varphi}\frac{B_{\varphi}}{B}+\xi_{z}\frac{B_z}{B}, \quad \vec{e}_{\parallel}=\vec{e}_{\varphi}\frac{B_{\varphi}}{B}+\vec{e}_z\frac{B_z}{B}, \nonumber \\ && k_{b}=\frac{m}{r}\frac{B_z}{B}-k_{z}\frac{B_\varphi}{B}, \nonumber \\ && k_{\parallel}=\vec{k}\cdot\vec{e}_{\parallel}=\frac{m}{r}\frac{B_{\varphi}}{B}+k_{z}\frac{B_{z}}{B}\cdot \nonumber \end{eqnarray}(8)In these more convenient variables, the components of the perturbed magnetic field can be obtained from Eq. (2):

δBr=ikBξr,δBz=imr(ξzBϕξϕBz)1rddr(rξrBz),\begin{eqnarray} && \delta B_{r}={\rm i}k_{\parallel}B\xi_r, \nonumber \\ && \delta B_{\varphi}= {\rm i}k_{z}\left(\xi_{\varphi}B_{z}-\xi_{z}B_{\varphi}\right)-\frac{{\rm d}}{{\rm d}r}\left(\xi_r B_{\varphi}\right), \label{a8} \\ && \delta B_{z}=\frac{{\rm i}m}{r}\left(\xi_{z}B_{\varphi}-\xi_{\varphi}B_{z}\right)-\frac{1}{r}\frac{{\rm d}}{{\rm d}r}\left(r\xi_{r}B_z\right), \nonumber \end{eqnarray}(9)and Eqs. (5)(7) can be written as

(ρω2k2B2)ξrddrδp12ikBϕBzrξb\begin{eqnarray} && \left(\rho\omega^2-k_{\parallel}^2 B^2\right)\xi_{r}-\frac{{\rm d}}{{\rm d}r}\delta p_1-2{\rm i}k_{\parallel}\frac{B_{\varphi}B_z}{r}\xi_{b} \nonumber \\ && \,\,\quad\quad\quad\quad \qquad +2{\rm i}k_{b}\frac{B_{\varphi}^2}{r}\xi_{b}+2\frac{B_{\varphi}}{r} \frac{{\rm d}}{{\rm d}r}\left(\xi_{r}B_{\varphi}\right)=0, \label{a11} \\ && \left(\rho\omega^2-k_{\parallel}^2 B^2\right)\xi_{b}-{\rm i}k_{b}\delta p_1+2{\rm i}k_{\parallel}\frac{\xi_r}{r}B_{\varphi}B_{z}=0, \label{a12} \\ && \left(\rho\omega^2-k_{\parallel}^2\frac{\gamma p}{1+\beta}\right)\xi_{\parallel}-{\rm i}k_{\parallel}\frac{\beta}{1+\beta} \left[\delta p_1 -2\frac{B_{\varphi}^2\xi_{r}}{r}\right]=0, \label{a13} \end{eqnarray}where the values of the total pressure δp1 and ·ξ are also represented via radial (ξr), bi-normal (ξb) and field aligned (ξ) displacements:

δp1=(γp+B2)·ξ+iB2kξ+2Bϕ2ξrr,\begin{eqnarray} &&\delta p_1 = -\left(\gamma p+B^2\right)\vec{\nabla}\cdot\bfxi+{\rm i}B^2k_{\parallel}\xi_{\parallel}+2B_{\varphi}^2\frac{\xi_{r}}{r}, \nonumber \\ &&\vec{\nabla}\cdot\bfxi=\frac{1}{r}\frac{{\rm d}}{{\rm d}r}\left(r\xi_{r}\right)+{\rm i}k_{b}\xi_{b}+{\rm i}k_{\parallel}\xi_{\parallel}. \label{a14} \end{eqnarray}(13)Equations (10)(13) are the starting point for further analysis. Let us reduce the Eqs. (10)(12) to the one single equation for the radial component of the displacement vector (ξr). From Eqs. (11, 12) one can obtain

\begin{eqnarray} && k_{b}\xi_{b}=\frac{{\rm i}}{\rho\left(\omega^2-\omega_{\rm A}^2\right)}\left[k_{b}^2\delta p_1-2k_{\parallel}k_{b}\frac{\xi_r}{r}B_{\varphi}B_{z}\right], \label{a15} \\ && k_{\parallel}\xi_{\parallel}=\frac{{\rm i}k_{\parallel}^2}{\rho\left(\omega^2-\omega_{T}^2\right)}\frac{c_{T}^2}{c_{\rm A}^2}\left[\delta p_1-\frac{2B_{\varphi}^2}{r}\xi_{r}\right]. \label{a16} \end{eqnarray}Here, ωA2=k2cA2,ωS2=k2cS2,ωT2=k2cT2,cA2=B2ρ,cS2=γpρ,cT2=cS21+β,β=cS2cA2·\begin{eqnarray} && \omega_{\rm A}^2=k_{\parallel}^2c_{\rm A}^2, \quad \omega_{\rm S}^2=k_{\parallel}^2c_{\rm S}^2, \quad \omega_{T}^2=k_{\parallel}^2c_{T}^2, \nonumber \\ && c_{\rm A}^2=\frac{B^2}{\rho}, \quad c_{\rm S}^2=\frac{\gamma p}{\rho}, \quad c_{T}^2=\frac{c_{\rm S}^2}{1+\beta}, \quad \beta=\frac{c_{\rm S}^2}{c_{\rm A}^2}\cdot \nonumber \end{eqnarray}By substituting Eqs. (14) and (15) into Eq. (13), we can obtain

δp1=1kb2+χ2[ρ(ω2ωA2)1rddr(rξr)\begin{eqnarray} && \delta p_1 =\frac{1}{k_{b}^2+\chi^2}\left[\rho\left(\omega^2-\omega_{\rm A}^2\right)\frac{1}{r}\frac{{\rm d}}{{\rm d}r}\left(r\xi_r\right) \right. \nonumber \\ && \left.\qquad\,\, +~2k_{\parallel}k_{b}\frac{\xi_r}{r}B_{\varphi}B_z+\frac{2B_{\varphi}^2}{r}\xi_r\chi^2\right], \label{a17} \end{eqnarray}(16)where χ2=(ω2ωA2)(ωS2ω2)(cS2+cA2)(ω2ωT2)·\begin{eqnarray*} \chi^2=\frac{\left(\omega^2-\omega_{\rm A}^2\right)\left(\omega_{\rm S}^2-\omega^2\right)}{\left(c_{\rm S}^2+c_{\rm A}^2\right)\left(\omega^2-\omega_{T}^2\right)}\cdot \end{eqnarray*}From Eqs. (4), (10), (14) and (16) we obtain the governing wave equation for the linear radial component (ξr): ddr[ρ(ω2ωA2)kb2+χ21rddr(rξr)]+2rξrddr[Bϕ2r2χ2kb2+χ2+BϕBzr2kkbkb2+χ2]=ρ(ω2ωA2)ξr+2ξrBϕddr(Bϕr)\begin{eqnarray} &&\hspace*{-2.5mm} \frac{{\rm d}}{{\rm d}r}\left[\frac{\rho\left(\omega^2-\omega_{\rm A}^2\right)}{k_{b}^2+\chi^2}\frac{1}{r}\frac{{\rm d}}{{\rm d}r}\left(r\xi_r\right)\right]+ 2r\xi_r\frac{{\rm d}}{{\rm d}r}\left[\frac{B_{\varphi}^2}{r^2}\frac{\chi^2}{k_{b}^2+\chi^2} \right. \nonumber \\ &&\qquad + \left.\frac{B_{\varphi}B_z}{r^2}\frac{k_{\parallel}k_{b}}{k_{b}^2+\chi^2}\right]=\rho\left(\omega^2-\omega_{\rm A}^2\right)\xi_r+2\xi_rB_{\varphi}\frac{{\rm d}}{{\rm d}r}\left(\frac{B_{\varphi}}{r}\right) \nonumber \\ &&\qquad \qquad\qquad-~ 4\xi_r\frac{B_{\varphi}^2}{r^2\rho}\frac{\chi^2}{k_{b}^2+\chi^2}\frac{\left(k_{\parallel}B_z-k_{\rm b}B_{\varphi}\right)^2}{\left(\omega^2-\omega_{\rm A}^2\right)}\cdot \label{a18} \end{eqnarray}(17)In Cheremnykh (2015) it was shown that Eq. (17) is equivalent to the well known Hain-Lüst equation (see e.g. Hain & Lüst 1958) and also it was shown that from this equation we can obtain Suydam’s criterion (see e.g. Suydam 1958) and stability criterion for ballooning modes.

3. Incompressible perturbations in the long wavelength approximation

Let us assume that plasma perturbation is incompressible, that is the velocity perturbation δv = ξ/∂t is very small relative to the sound speed, cS2\hbox{$c_{\rm S}^2\rightarrow\infty$}(γ → ∞). It follows from the second equation of (13) and Eqs. (14), (15) that ·ξ=1rddr(rξr)δp1ρ(ω2ωA2)(kz2+m2r2)+2kmr2BϕBξrρ(ω2ωA2)·\begin{equation} \vec{\nabla}\cdot{\bfxi}=\frac{1}{r}\frac{{\rm d}}{{\rm d}r}\left(r\xi_{r}\right)-\frac{\delta p_1}{\rho \left(\omega^2-\omega_{\rm A}^2\right)}\left(k_{z}^2+\frac{m^2}{r^2}\right)+2k_{\parallel}\frac{m}{r^2}\frac{B_{\varphi}B\xi_r}{\rho\left(\omega^2-\omega_{\rm A}^2\right)}\cdot \label{a19} \end{equation}(18)For an incompressible perturbation the expression (16) for δp1 becomes δp1=1kz2+(m2/r2)[ρ(ω2ωA2)1rddr(rξr)+2kmrξrrBϕB].\begin{equation} \delta p_1=\frac{1}{k_{z}^2+\left(m^2/r^2\right)}\left[\rho\left(\omega^2-\omega_{\rm A}^2\right)\frac{1}{r}\frac{{\rm d}}{{\rm d}r}\left(r\xi_r\right)+2k_{\parallel}\frac{m}{r}\frac{\xi_r}{r}B_{\varphi}B\right]. \label{a20} \end{equation}(19)By substituting Eq. (19) into Eq. (18), we obtain the condition for the divergence of an incompressible flow, that is, ·ξ=0.\begin{equation} \vec{\nabla}\cdot{\bfxi}=0. \label{a21} \end{equation}(20)Equation (20) can also be obtained in another way. By substituting δp1 from Eq. (13) into Eq. (12), we obtain iω2ξ=kcS2·ξ\hbox{${\rm i}\omega^2\xi_{\parallel}= k_{\parallel}c_{\rm S}^2 \vec{\nabla} \cdot \bfxi$}. By assuming cS2\hbox{$c_{\rm S}^2\rightarrow \infty$}, we also arrive at the same Eq. (20).

For incompressible perturbations Eq. (17) reduces to the equation ddr[ρ(ω2ωA2)kz2+m2/r21rddr(rξr)]+2rξrddr[BBϕr2k(m/r)kz2+m2/r2]ξr[ρ(ω2ωA2)+2Bϕddr(Bϕr)\begin{eqnarray} && \hspace*{-4mm}\frac{{\rm d}}{{\rm d}r}\left[ \frac{\rho\left(\omega^2-\omega_{\rm A}^2\right)}{k_{z}^2+m^2/r^2} \frac{1}{r} \frac{{\rm d}}{{\rm d}r} \left(r\xi_r\right)\right]+2r\xi_r \frac{{\rm d}}{{\rm d}r}\left[\frac{BB_{\varphi}}{r^2}\frac{k_{\parallel}\left(m/r\right)}{k_{z}^2+m^2/r^2}\right] \nonumber \\ &&\qquad -~~ \xi_r\left[\rho\left(\omega^2-\omega_{\rm A}^2\right)+2B_{\varphi}\frac{{\rm d}}{{\rm d}r}\left(\frac{B_{\varphi}}{r}\right) \right. \nonumber \\ && \left. \qquad\qquad\qquad -\frac{4\left(B_{\varphi}^2/r^2\right)}{k_{z}^2+m^2/r^2}\frac{k_z^2\omega_{\rm A}^2}{\left(\omega^2-\omega_{\rm A}^2\right)}\right]=0 \label{a22} \end{eqnarray}(21)and coincides with the Eq. (14.36) in Miyamoto (2005). Equation (21) has only one singular point at Alfvén frequency, when ω2=ωA2\hbox{$\omega^2=\omega_{\rm A}^2$}. To obtain this equation we assumed that kcS → ∞ and, therefore, the longitudinal component wave vector k cannot vanish.

For perturbations with a small azimuthal wave number m, Eq. (21) can be simplified further by implementing the long wavelength approximation, that is, for the case when ε = kza ≪ 1. In this limit, for a homogeneous longitudinal magnetic field, Bz = const., Eq. (21) is given by

1rddr[(ρω2F2)rdφdr]+dF2drφr(ρω2F2)m2φr2\begin{eqnarray} && \frac{1}{r}\frac{{\rm d}}{{\rm d}r}\left[\left(\rho\omega^2-F^2\right)r\frac{{\rm d}\phi}{{\rm d}r}\right]+\frac{{\rm d}F^2}{{\rm d}r}\frac{\phi}{r}- \left(\rho\omega^2-F^2\right)\frac{m^2\phi}{r^2} \nonumber \\ &&\qquad\qquad +~ 4\frac{B_{\varphi}^2}{r^2}\frac{k_{z}^2F^2}{\left(\rho\omega^2-F^2\right)}\phi=0, \label{a23} \end{eqnarray}(22)where φ=rξr,F(r)=mrBϕ(r)+kzBz.\begin{eqnarray*} \phi=r\xi_{r}, \, F\left(r\right)=\frac{m}{r}B_{\varphi}\left(r\right)+k_{z}B_{z}. \nonumber \end{eqnarray*}To obtain Eq. (22) we also used the relevant long wavelength approximation that m2+kz2r2=m2+(ra)2ε2m2.\begin{eqnarray*} m^2+k_{z}^2r^2=m^2+\left(\frac{r}{a}\right)^2\varepsilon^2\approx m^2. \nonumber \end{eqnarray*}Further analysis of the Eq. (22) requires an equilibrium model for the flux tube and the boundary conditions for the perturbed quantities on its surface.

4. Boundary conditions

To obtain the first boundary condition we used the equation of incompressibility given by Eq. (20) and relation Eq. (13). By following the methodology presented in Jackson (1998) and Priest & Hood (1991), let us integrate Eq. (20) over a small interval 2δ near flux tube surface (i.e, within the layer aδ and a + δ, a is the tube radius, δa) results in

aδa+δr·ξdr=aδa+δ[ddr(rξr)+ir(kbξb+kξ)]dr=rξr|a+δaδ+iaδa+δr(kbξb+kξ)dr=0.\begin{eqnarray} \int_{a-\delta}^{a+\delta}r\vec{\nabla}\cdot\bfxi {\rm d}r \,\,&=&\, \int_{a-\delta}^{a+\delta}\left[\frac{{\rm d}}{{\rm d}r}\left(r\xi_r\right)+{\rm i}r\left(k_{b}\xi_{b}+k_{\parallel}\xi_{\parallel}\right)\right]{\rm d}r \nonumber \\ \,& =&\, r\xi_r\left|_{a-\delta}^{a+\delta}\right. +{\rm i}\int_{a-\delta}^{a+\delta}r\left(k_{b}\xi_{b}+k_{\parallel}\xi_{\parallel}\right){\rm d}r \nonumber\\ \,&=&\,0. \nonumber \end{eqnarray}By taking into account the continuity of the integrated function, this equation can be represented as φ|a+δaδ+2iδ[r(kbξb+kξ)]|r=a=0.\begin{eqnarray*} \phi\left|_{a-\delta}^{a+\delta}\right.+2{\rm i}\delta\left[r\left(k_{b} \xi_{b}+k_{\parallel} \xi_{\parallel}\right)\right]\left|_{r=a}\right.=0. \nonumber \end{eqnarray*}Assuming that δ → 0, the first boundary condition can be represented in the form: φ=φ(a+0)φ(a0)=0.\begin{equation} \left<\phi\right>=\phi\left(a+0\right)-\phi\left(a-0\right)=0. \label{a25} \end{equation}(23)This condition requires that the radial plasma displacement is continuous at r = a. Now let us obtain the second boundary condition. Equation (22) can be rewritten as ddr[(ρω2F2)rdr]+φddr[2mrBϕFm2(Bϕr)2]\begin{eqnarray} && \hspace*{-4mm}\frac{{\rm d}}{{\rm d}r}\left[\left(\rho\omega^2-F^2\right)r\frac{d\phi}{{\rm d}r}\right]+\phi\frac{{\rm d}}{{\rm d}r}\left[\frac{2m}{r}B_{\varphi}F-m^2\left(\frac{B_{\varphi}}{r}\right)^2\right] \nonumber \\ &&\qquad\qquad - \frac{\phi}{r}\left[m^2\left(\rho \omega^2-F^2\right)-\frac{4B_{\varphi}^2k_{z}^2F^2}{\left(\rho\omega^2-F^2\right)}\right]=0. \label{a26} \end{eqnarray}(24)By taking into account that the magnetic field and plasma equilibrium parameters are different inside and outside of the magnetic flux tube, Eq. (24) has different solutions for r>a and r<a. The inner and outer solutions should agree for values of r in an intermediate region (aδ,a+δ)\hbox{$\left(a-\delta, \, a+\delta\right)$} if δ → 0. By assuming that radial plasma displacement ξr is continuous, the matching condition at the boundary (see e.g. Soloviev 1975; Priest & Hood 1991) can be obtained by integrating Eq. (24) between aδ and a + δ: (ρω2F2)rdφdr+2mrBϕm2(Bϕr)2φ=0.\begin{equation} \left<\left(\rho\omega^2-F^2\right)r\frac{{\rm d}\phi}{{\rm d}r}+\frac{2m}{r}B_{\varphi}F\phi- m^2\left(\frac{B_{\varphi}}{r}\right)^2\phi\right>=0. \label{a27} \end{equation}(25)For the function φ inside and outside the flux tube, Eq. (25) represents dispersion the relation for MHD oscillations. According to Eq. (19) the total pressure perturbation is given by δp1=1m2[2mrBϕ+(ρω2F2)rdφdr]·\begin{eqnarray*} \delta p_{1}=\frac{1}{m^2}\left[\frac{2m}{r}B_{\varphi}F\phi+\left(\rho\omega^2-F^2\right)r\frac{{\rm d}\phi}{{\rm d}r}\right]\cdot \nonumber \end{eqnarray*}This means that Eq. (25) is equivalent to δp1Bϕ2r2φ=0.\begin{equation} \left<\delta p_{1}-\frac{B_{\varphi}^2}{r^2}\phi\right>=0. \label{a27a} \end{equation}(26)Equation (26) is the dynamic boundary condition that is often applied to study MHD perturbations of magnetic flux tubes (Erdélyi & Fedun 2006; Bennett et al. 1999; Erdélyi & Fedun 2007, to name but a few).

5. General dispersion relation for m 1 modes

In this section, Eq. (22) together with the boundary conditions Eqs. (23) and (26) will be used to obtain the dispersion equation governing oscillations of the magnetic tube. In the long wavelength approximation in Eq (22) the last term is proportional to ε2 and hence a good approximation can be neglected, resulting in 1rddr[(ρω2F2)rdφdr]+dF2drφrm2r2(ρω2F2)φ=0.\begin{eqnarray} \frac{1}{r}\frac{{\rm d}}{{\rm d}r}\left[\left(\rho\omega^2-F^2\right)r\frac{{\rm d}\phi}{{\rm d}r}\right]+\frac{{\rm d}F^2}{{\rm d}r}\frac{\phi}{r}- \frac{m^2}{r^2}\left(\rho\omega^2-F^2\right)\phi=0. \label{a28} \end{eqnarray}(27)This equation has been obtained previously by Wesson (1978) in the study of stability of high-temperature plasma. Changing back to the physical variable ξr, it is easy to show that Eq. (27) is equivalent to 1rddr[(ρω2F2)r3dξrdr]+(1m2)(ρω2F2)ξr=0.\begin{eqnarray} \frac{1}{r}\frac{{\rm d}}{{\rm d}r}\left[\left(\rho\omega^2-F^2\right)r^3\frac{{\rm d}\xi_r}{{\rm d}r}\right]+ \left(1-m^2\right) \left(\rho\omega^2-F^2\right)\xi_r=0. \label{CC1} \end{eqnarray}(28)For the specific kink mode value of m = 1, from Eq. (28) we obtain: 1rddr[(ρω2F2)r3dξrdr]=0.\begin{eqnarray} \frac{1}{r}\frac{{\rm d}}{{\rm d}r}\left[\left(\rho\omega^2-F^2\right)r^3\frac{{\rm d}\xi_r}{{\rm d}r}\right]=0. \label{CC2} \end{eqnarray}(29)Assuming zero background magnetic twist outside the tube but an arbitrary twist inside: B={(0,Bϕ(r),Bzi),ra(0,0,Bze)r>a.\begin{equation} \vec{B} = \left\{ \begin{array}{lr} \left(0, B_{\varphi}\left(r\right), B_{z\rm i}\right), & r\le a\\ \left(0, 0, B_{z\rm e}\right) & r > a. \end{array} \right. \label{CC3} \end{equation}(30)The physical solution of Eq. (29), for a trapped mode with background magnetic field Eq. (30) bounded at r = 0 and tending to ξr = 0 as r → ∞ is given as ξr(r)={ξa=const.,raξa(ar)2,r>a.\begin{equation} \xi_r(r) = \left\{ \begin{array}{lr} \xi_a={\rm const.}, & r\le a\\ \xi_a\left(\frac{a}{r}\right)^{2}, & r > a. \end{array} \right. \label{CC4} \end{equation}(31)Physically, the radial displacement inside the tube is restricted to be constant with the radius for the m = 1 mode (see Eq. (31)) or else Eq. (29) would give a singularity at r = 0. Mathematically, this singularity can be eliminated only in the case where Bϕ is not finite at the tube axis which is unphysical. The fact that Eq. (31) is the same as for an untwisted tube is a new and interesting analytical result since previously Ruderman (2007), for example, only demonstrated this for the particular internal background magnetic twist of Bϕr. Here we have shown that the radial profile of ξr the kink mode, in the thin tube approximation, is completely independent of any prescribed background internal twist. This is also in agreement with the purely numerical study of Terradas & Goossens (2012) who solved the ideal linearised MHD equations in the zero-β regime using the PDE2D code (Sewell 2005). Terradas & Goossens found that the kink mode frequency in the long wavelength approximation was not affected by their particular choice of a quadratic radial profile of Bϕ shown in Eq. (8) of their paper, which is consistent with our more general analytical result.

By substituting Eq. (31) into the boundary condition Eq. (25) we obtain: (ρi+ρe)ω2=(Fi(r)2+Fe2)|r=a2Bϕ(r)Fi(r)r|r=a+Bϕ2(r)r|r=a·\begin{equation} \left.\left(\rho_{\rm i}+\rho_{\rm e}\right)\omega^2=\left(F_{\rm i}\left(r\right)^2+F_{\rm e}^2\right) \right|_{r=a}- \left.2\frac{B_{\varphi}(r)F_{\rm i}\left(r\right)} {r} \right|_{r=a}+ \left.\frac{B_{\varphi}^2\left(r\right)}{r}\right|_{r=a}\cdot \label{CC5} \end{equation}(32)Here, subscripts i and e correspond to the internal and external parts of the magnetic flux tube correspondingly. By taking into account that

Fi(r)=Bϕ(r)/r+kzBzi,\begin{eqnarray} && F_{\rm i}\left(r\right)=B_{\varphi}\left(r\right)/r+k_{z}B_{z\rm i}, \nonumber \\ && F_{\rm e}=k_{z}B_{z\rm e}, \label{CC6} \end{eqnarray}(33)from Eq. (32) we obtain the dispersion relation ω2=kz2(ρi+ρe)(Bzi2+Bze2),\begin{equation} \omega^2=\frac{k_{z}^2}{\left(\rho_{\rm i}+\rho_{\rm e}\right)}\left(B_{z\rm i}^2+B_{z\rm e}^2\right), \label{CC7} \end{equation}(34)which again, is the same as the kink mode in the thin tube approximation without twist. Therefore, by Eqs. (34) and (31), both the frequency and ξr eigenfunctions are unaffected by the choice of internal twist for the kink mode.

By inspection of Eq. (28) it can be seen that the case is entirely different for modes with m ≥ 2. In such cases the dispersion relation and ξr eigenfunction will depend substantially on the radial profile of magnetic twist. However, unlike the m = 1 kink mode, Eq. (28) is less tractable for analysis when the radial profile of Bϕ is arbitrary for the modes with m ≥ 2. Hence, in the next section we shall choose a specific radial twist profile which will enable us to do this.

6. Internal background magnetic twist with Bϕr

To make analysis of Eq. (28) more straightforward for fluting modes with m ≥ 2 we will assume that inside the tube the magnetic twist varies linearly and outside it is zero, for example Bennett et al. (1999), Parker (2007), Solov’ev (2011, 2012): B={(0,Bϕ(a)r/a,Bzi),ra(0,0,Bze)r>a\begin{equation} \vec{B} = \left\{ \begin{array}{lr} \left(0, B_{\varphi}\left(a\right)r/a, B_{z\rm i}\right), & r\le a\\ \left(0, 0, B_{z\rm e}\right) & r > a \end{array} \right. \label{a29} \end{equation}(35)where Bzi and Bze are constant internal and external magnetic fields. We note that after introduction of this specific magnetic geometry (i.e. Eq. (35)), the quantity Fr)(\hbox{$F\left(r\right)$} loses its dependence on r. From Eqs. (27) and (35) we obtained that inside and outside the flux, tube Eq. (28) can be represented as (ρω2F2)[1r(r3dξrdr)+(1m2)ξr]=0.\begin{equation} (\rho \omega^2-F^2)\left[\frac{1}{r}\left(r^3\frac{{\rm d} \xi_r}{{\rm d} r}\right)+\left(1-m^2\right) \xi_r \right]=0. \label{a30} \end{equation}(36)By applying the boundary conditions given by Eq. (23) the solution of Eq. (36), which describes perturbation of the plasma cylinder border for m ≥ 1, is ξr={ξa(ra)m1,raξa(ar)m+1,r>a.\begin{equation} \xi_r = \left\{ \begin{array}{lr} \xi_a\left(\frac{r}{a}\right)^{m-1}, & r\le a\\ \xi_a\left(\frac{a}{r}\right)^{m+1}, & r > a. \end{array} \right. \label{a31} \end{equation}(37)Here ξa = ξr(r = a). By substituting Eq. (37) into Eq. (25) we obtain the dispersion relation in the form:

ω2=1(ρi+ρe)[kz2(Bzi2+Bze2)+m(m1)Bϕ2(a)a2\begin{eqnarray} && \omega^2=\frac{1}{\left(\rho_{\rm i}+\rho_{\rm e}\right)}\left[k_{z}^2\left(B_{z\rm i}^2+B_{z\rm e}^2\right)+m\left(m-1\right)\frac{B_{\varphi}^2(a)}{a^2}\right. \nonumber \\ && \left. \qquad +\frac{2k_z}{a}\left(m-1\right)B_{\varphi}(a)B_{z\rm i}\right]. \label{a32} \end{eqnarray}(38)This dispersion relation is valid for positive azimuthal m and longitudinal kz wave numbers. It is important to note that this relation is invariant under the substitution (m,kz) → (−m,kz), resulting in:

ω2=1(ρi+ρe)[kz2(Bzi2+Bze2)+m(msign(m))Bϕ2(a)a2\begin{eqnarray} && \omega^2=\frac{1}{\left(\rho_{\rm i}+\rho_{\rm e}\right)}\left[k_{z}^2\left(B_{z\rm i}^2+B_{z\rm e}^2\right)+m\left(m-{\rm sign}(m)\right)\frac{B_{\varphi}^2(a)}{a^2}\right. \nonumber \\ && \left.\qquad + \frac{2k_z}{a}\left(m-{\rm sign}(m)\right)B_{\varphi}(a)B_{z\rm i}\right]. \label{a32.1} \end{eqnarray}(39)Hence, Eq. (39) describes wave propagation in both directions along the twisted magnetic tube. If current is absent, meaning that Bϕi = 0, the two last terms vanish and Eq. (39) describes the kink mode (see e.g. Ryutov & Ryutova 1976; Spruit 1982; Edwin & Roberts 1983). Although m in Eq (39) can either be positive or negative, in the following sections we analyse only the case when m> 0. ξr(r)={ξa=const.,raξa(ar)2,r>a,\begin{equation} \xi_r(r) = \left\{ \begin{array}{lr} \xi_a={\rm const.}, & r\le a\\ \xi_a\left(\frac{a}{r}\right)^{2}, & r > a, \end{array} \right. \label{a34} \end{equation}(40)where ξa = ξr(r = a) = φa/r. This function is constant up to the boundary of the flux tube and then decreasing to the infinitely small values as r → ∞ (see Fig. 2).

thumbnail Fig. 2

Radially dependent eigenfunctions ξr(r) for axially symmetric magnetic flux tube with a free boundary are shown for azimuthal wave numbers m = 1, 2, 3, and 4.

The square of the frequency of fluting modes (m ≥ 2) given by Eq. (38), in contrast to kink modes (m = 1), has a minimum value when kz=(m1)aBϕ(a)BziBzi2+Bze2\begin{eqnarray*} k_z=-\frac{\left(m-1\right)}{a}\frac{B_{\varphi}(a)B_{z\rm i}}{B_{z\rm i}^2+B_{z\rm e}^2} \nonumber \end{eqnarray*}and is equal to ωmin2=Bϕ2(a)a2(m1)(ρi+ρe)[1+(m1)Bze2Bzi2+Bze2]·\begin{equation} \omega_{\rm min}^2=\frac{B_{\varphi}^2(a)}{a^2}\frac{\left(m-1\right)}{\left(\rho_{\rm i}+\rho_{\rm e}\right)}\left[1+\frac{\left(m-1\right)B_{z\rm e}^2}{B_{z\rm i}^2+B_{z\rm e}^2}\right]\cdot \label{a35} \end{equation}(41)In contrast to the kink mode, Eq. (41) shows that a low frequency cut-off is introduced for fluting modes due to background twist. For all m ≥ 2 modes the eigenfunction ξr has a form of power function Eq. (37) and describes perturbations which are localised at the surface of the twisted magnetic flux tube (see Fig. 2). It is also interesting to note that the sign of the ξr eigenfunctions ξr do not vary as r increases.

7. Comparison with previous results

Equation (35) describes the same background magnetic field configuration as Bennett et al. (1999), to allow for direct comparison. In the paper by Bennett et al. the authors obtain the specific dispersion relation which describes sausage (m = 0) waves in an incompressible magnetic flux tube with uniform twist embedded within an untwisted magnetic environment, that is:

(ω2ωA2)xIm(x)dIm(x)dx|x=|m0|a2mBϕ(a)ωAiaρi(ω2ωAi2)24ωAi2Bϕ2(a)a2ρi=\begin{eqnarray} && \frac{\left. \left(\omega^2-\omega_{\rm A}^2\right)\frac{x}{I_{m}(x)}\frac{{\rm d}I_{m}(x)}{{\rm d}x}\right|_{x=\left|m_0\right|a}-\frac{2mB_{\varphi}(a)\omega_{\rm Ai}}{a\sqrt{\rho_{\rm i}}}}{{\left(\omega^2-\omega_{\rm Ai}^2\right)^2-4\omega_{\rm Ai}^2}\frac{B_{\varphi}^2(a)}{a^2\rho_{\rm i}}}= \nonumber \\ &&\qquad \frac{\left.\frac{y}{K_{m}(y)}\frac{{\rm d}K_{m}(y)}{{\rm d}y}\right|_{y=\left|k_z\right|a}}{\frac{\rho_{\rm e}}{\rho_{\rm i}}\left(\omega^2-\omega_{\rm Ae}^2\right)+\left. \frac{B_{\varphi}^2(a)}{a^2\rho_{\rm i}}\frac{y}{K_{m}(y)}\frac{{\rm d}K_{m}(y)}{{\rm d}y}\right|_{y=\left|k_z\right|a}}\cdot \label{a38} \end{eqnarray}(42)Here Im(x) and Km(x) are modified Bessel functions of the first and second kind with an imaginary argument, m02a2=kz2a2(14Bϕ2r2F2(ρω2F2)2),\begin{equation} m_{0}^2a^2=k_{z}^2a^2\left(1-\frac{4B_{\varphi}^2}{r^2}\frac{F^2}{\left(\rho\omega^2-F^2\right)^2}\right) \label{a39} , \end{equation}(43)and ω2=1ρα(maBϕ(a)+kzB)2.\begin{equation} \omega_{A\alpha}^2=\frac{1}{\rho_\alpha}\left(\frac{m}{a}B_{\varphi}(a)+k_zB_{z\alpha}\right)^2. \end{equation}(44)The subscript α refers to i and e, which corresponds to the quantities either inside and outside of the tube, respectively. We note that Bennett et al. (1999) have shown numerically (see. Fig. 5 of their paper) that for the kink mode (m = 1), the phase and group velocities are approximately equal, such that: ωkz∂ωkz(Bzi2+Bze2ρi+ρe)12,\begin{equation} \frac{\omega}{k_{z}}\approx \frac{\partial \omega}{\partial k_{z}}\approx \left(\frac{B_{z\rm i}^2+B_{z\rm e}^2}{\rho_{\rm i}+\rho_{\rm e}}\right)^{\frac{1}{2}}, \label{aa1} \end{equation}(45)which is in a great agreement with Eq. (34). We now want to check that the dispersion relations given by Eqs. (38) and (42) are equivalent to each other in the long wavelength approximation when m = 0 and furthermore, that the eigenfunctions are also the same. As mentioned in Bennett et al. (1999), it is difficult to analyse analytically the modes with m ≥ 1 and hence this was not attempted in their work. This is partly due to the fact that modes with m ≥ 1 can exist with both m02>0\hbox{$m_{0}^2>0$} and m02<0\hbox{$m_{0}^2<0$}. Equation (42) was obtained for the particular case of m02>0\hbox{$m_{0}^2>0$}. Let us show that these conditions do not actually affect the final form of the dispersion equation in the long wavelength approximation. In the case in which m02<0\hbox{$m_{0}^2<0$} only one change in the left hand side of Eq. (42) will appear due to the replacement (see Appendix B, Eq. (B.8)) xImdImdx|x=m0axJmdJmdx|x=|m0|a·\begin{equation} \left. \frac{x}{I_m}\frac{{\rm d}I_m}{{\rm d}x}\right|_{x=m_{0}a} \rightarrow \left. \frac{x}{J_m}\frac{{\rm d}J_m}{{\rm d}x}\right|_{x=\left| m_{0}\right|a}\cdot \label{aa2} \end{equation}(46)Here Jm(x) is the Bessel function of the first kind. In the long-wave approximation |m0|a ≈ |kz|a ≪ 1, therefore the Bessel functions satisfy following relations: xImIm|x=m0a1=xJmJm|x=|m0|a1=xKmKm|x=|kz|a1=m.\begin{equation} \left. \frac{xI_{m}^{\prime}}{I_{m}}\right|_{x=m_0 a \ll 1} = \left. \frac{xJ_{m}^{\prime}}{J_{m}}\right|_{x=\left|m_0\right| a \ll 1} = -\left. \frac{xK_{m}^{\prime}}{K_{m}}\right|_{x=\left|k_z\right| a \ll 1} =m. \label{a40} \end{equation}(47)By applying relations shown in Eq. (47) we arrive to the conclusion that, independent of the sign of m02\hbox{$m_{0}^2$}, Eq. (42) can be written in the form: (ω2ωAi2)+2ωAi(Bϕ(a)aρi)+ρeρi(ω2ωAe2)mBϕ2(a)a2ρi=0.\begin{equation} \left(\omega^2-\omega_{\rm Ai}^2\right)+2\omega_{\rm Ai}\left(\frac{B_{\varphi}\left(a\right)}{a\sqrt{\rho_{\rm i}}}\right)+\frac{\rho_{\rm e}}{\rho_{\rm i}}\left(\omega^2-\omega_{\rm Ae}^2\right)-m\frac{B_{\varphi}^2(a)}{a^2\rho_{\rm i}}=0. \label{a41} \end{equation}(48)After some algebra it can be shown that Eq. (48) coincides with the previously obtained dispersion relation in Eq. (38).

We now wish to further confirm our results by comparing the eigenfunctions of Eqs. (37) and (40) with those of Bennett et al. (1999). Bennett et al. obtained a solution (see Eq. (B.6)) for the total pressure perturbation δp1 when m02>0\hbox{$m_{0}^2>0$}: δp1={Im(m0r),raKm(|kz|r),r>a.\begin{equation} \delta p_1 = \left\{ \begin{array}{lr} I_m\left(m_0 r\right), & r\le a\\[2mm] K_m\left(\left|k_z\right|r\right), & r > a. \end{array} \right. \label{aa3} \end{equation}(49)If m02<0\hbox{$m_{0}^2<0$}, the expression for δp1 has the following form: δp1={Jm(|m0|r),raKm(|kz|r),r>a.\begin{equation} \delta p_1 = \left\{ \begin{array}{lr} J_m\left(\left|m_0\right|r\right), & r\le a\\[2mm] K_m\left(\left|k_z\right|r\right), & r > a. \end{array} \right. \label{aa4} \end{equation}(50)Since in the long wavelength approximation, functions Im(x), Jm(x) and Km(x) satisfy relations: Im(x),Jm(x)|x1~xm,Km(x)|x1~xm,\begin{equation} I_m (x), \, \left. J_m (x)\right|_{x \ll 1} \sim x^m, \, \left. K_{m}(x)\right|_{x \ll 1}\sim x^{-m}, \end{equation}(51)δp1, independent of the sign of m02\hbox{$m_{0}^2$}, is given by: δp1~{rm,rarm,r>a.\begin{equation} \delta p_1 \sim \left\{ \begin{array}{lr} r^m, & r\le a\\[2mm] r^{-m}, & r > a. \end{array} \right. \label{a42} \end{equation}(52)According to Eqs. (B.1), (B.4) and (B.5) the total pressure perturbation δp1 is associated with radial displacement ξr as: rξr=const.1rddrδp1+const.2δp1.\begin{equation} r\xi_r={\rm const.}_1r\frac{{\rm d}}{{\rm d}r}\delta p_1 +{\rm const.}_2 \delta p_1. \label{a43} \end{equation}(53)From Eqs. (52) and (53) the radial displacement ξr depends on r as follows: ξr~{rm1,rarm1,r>a.\begin{equation} \xi_r \sim \left\{ \begin{array}{lr} r^{m-1}, & r\le a\\[2mm] r^{-m-1}, & r > a. \end{array} \right. \label{aa5} \end{equation}(54)This is consistent with the radial dependence of the eigenfunctions previously derived from dispersion relation (38) and shown in Eqs. (37) and (40).

Now we shall compare our results with those obtained by Ruderman (2007) who studied linear non-axisymmetric oscillations of a thin magnetic tube in presence of the weak internal magnetic twist, that is Bϕ(ra) ≪ Bzi and Bϕ(r>a) = 0. Ruderman obtained the dispersion relation for incompressible MHD perturbations in the form of ω2=2B02(ρi+ρe)kz2+A(msign(m))2B02(Am+2B0kz).\begin{equation} \omega^2=\frac{2B_{0}^2}{\left(\rho_{\rm i}+\rho_{\rm e}\right)}\left\{ k_{z}^2+\frac{A\left(m-{\rm sign}\left(m\right)\right)}{2B_{0}^2}\left(Am+2B_{0}k_z\right)\right\}. \label{aa6} \end{equation}(55)It can be seen that this equation is the same as Eq. (39) when Bϕ(a) /a = A and Bzi = Bze = B0.

Therefore, the dispersion relation of Eq. (39) and the resulting expressions for the eigenfunctions in Eqs. (37) and (40) are in excellent agreement with the previous results of Bennett et al. (1999), who studied more specialised cases of our more general plasma and magnetic field background configuration.

8. Conclusions

In this work we analysed the incompressible linear MHD modes of a twisted magnetic flux tube. Special attention was given to the problem of finding the eigenvalues and eigenfunctions of modes with m ≥ 1 in the long wavelength limit. The description of these modes leads to significant mathematical difficulties as noted in Bennett et al. (1999). However, in this current work, by using the long wavelength approximation we have made significant analytical progress. Notably, we have shown that the dispersion relation given in Eq. (34) for the m = 1 kink mode is completely unaffected by the radial profile of background internal magnetic twist. It was also found from the ξr eigenfunction that the kink mode is body-like in character (see Eq. (40)) and the higher order modes (m ≥ 2) are surface-like (see Eq. (37)). Hence, these results are the same as known previously in the case of no background magnetic twist for the particular choice of having Bϕr inside the tube. However, it was shown in Eq. (41), that a low frequency cut-off was introduced for fluting modes due to the presence of background twist, in contrast to the cut-off free propagation for the kink mode.

It can be seen by the derived dispersion relation shown in Eq. (34) that the phase speed of the kink mode, equal to the group speed in the long wavelength limit, will not be affected at all by the presence of internal background magnetic twist. When there is no internal twist, all surface modes tend to the kink speed in the long wavelength limit. However, this is not the case when twist is present, since it can either work for or against the propagation speed of these modes. For example, when m, kz and Bϕ are all the same sign, the resulting phase speed is increased relative to the kink speed. If the sign of Bϕ is opposite to that of m and kz, then the speed is reduced relative to the kink speed. This is analogous to MHD wave mode propagation along a magnetic flux tube in the presence of field-aligned flow. That is to say, if the wave is travelling in the same or opposite direction to that of the flow, the speed is increased/decreased relative to the case when no flow is present. Hence, this presents a challenge in interpreting what differences in observed counter-propagating MHD wave-mode speeds could be caused by, meaning a magnetic twist, field-aligned flow, or a combination of both. Futhermore, to our knowledge m ≥ 2 modes have still yet to be identified in solar atmospheric observations of thin twisted magnetic flux tubes, for example, chromospheric fibrils, mottles, and spicules. However, the current work suggests their sensitivity to magnetic twist would make them a very interesting future case study.

Acknowledgments

All authors thank the referee and Prof. M. Ruderman for the constructive suggestions that improved the paper. O.C. and A.K. would like to thank the Ukrainian Scientific and Technical Center, PN 6060; Integrated ScientificProgrammes of the National Academy of Science of Ukraine on Space Research and Plasma Physics. V.F., G.V. thank the STFC and Royal Society-Newton Mobility Grant for the support received. V.F. also thanks the Newton Fund MAS/CONACyT Mobility Grants Program.

References

  1. Bateman, G. 1978, MHD instabilities (Cambridge, Mass.: MIT Press) [Google Scholar]
  2. Bennett, K., Roberts, B., & Narain, U. 1999, Sol. Phys., 185, 41 [NASA ADS] [CrossRef] [Google Scholar]
  3. Burdo, O. S., Cheremnykh, O. K., Revenchuk, S. M., & Pustovitov, V. D. 1994, Plasma Phys. Control. Fusion, 36, 641 [NASA ADS] [CrossRef] [Google Scholar]
  4. Cheremnykh, O. K. 1989, Nucl. Fusion, 29, 1899 [CrossRef] [Google Scholar]
  5. Cheremnykh, O. K. 2015, Kinemat. Phys. Celest. Bodies, 31, 213 [NASA ADS] [CrossRef] [Google Scholar]
  6. Cheremnykh, O. K., Andrushchenko, Z. M., Edenstrasser, J. W., & Taranov, V. B. 1994, Phys. Plasmas, 1, 2525 [NASA ADS] [CrossRef] [Google Scholar]
  7. Cheremnykh, O. K., Klimushkin, D. Yu., & Kostarev, D. V. 2014, Kinemat. Phys. Celest. Bodies, 30, 209 [NASA ADS] [CrossRef] [Google Scholar]
  8. Edwin, P. M., & Roberts, B. 1983, Sol. Phys., 88, 179 [NASA ADS] [CrossRef] [Google Scholar]
  9. Erdélyi, R., & Fedun, V. 2006, Sol. Phys., 238, 41 [NASA ADS] [CrossRef] [Google Scholar]
  10. Erdélyi, R., & Fedun, V. 2007, Sol. Phys., 246, 101 [NASA ADS] [CrossRef] [Google Scholar]
  11. Erdélyi, R., & Fedun, V. 2010, Sol. Phys., 263, 63 [NASA ADS] [CrossRef] [Google Scholar]
  12. Filippov, B. P. 2007, Eruptive process on the Sun (Moscow: Fizmatlit), 216 [in Russian] [Google Scholar]
  13. Galeev, A. A., & Sudan, R. N. 1989, Basic plasma physics. Handbook of plasma physics (Amsterdam: North-Holland) [Google Scholar]
  14. Giagkiozis, I., Fedun, V., Erdélyi, R., & Verth, G. 2015, ApJ, 810, 53 [NASA ADS] [CrossRef] [Google Scholar]
  15. Hain, K., & Lüst, R. 1958, Z. Naturforsh. A., 13, 936 [NASA ADS] [Google Scholar]
  16. Jackson, J. D. 1998, Classical Electrodynamics, 3rd edn. (John Wiley & Sons Ltd) [Google Scholar]
  17. Ladikov-Roev, Y. P., Cheremnykh, S. O., Yatsenko, V. A. 2013, J. Automat. Info. Sci., 45, 48 [CrossRef] [Google Scholar]
  18. Kadomtsev, B. B. 1966, Rev. Plasma Phys., 2, 153 [NASA ADS] [Google Scholar]
  19. Miyamoto, K. 2005, Plasma Physics and Controlled Nuclear Fusion, Springer Series on Atomic, Optical, and Plasma Physics, 38 (Berlin, Heidelberg: Springer-Verlag) [Google Scholar]
  20. Parker, E. N. 2007, Conversations on Electric and Magnetic Fields in the Cosmos (Princeton: Princeton University Press) [Google Scholar]
  21. Priest, E. R. 1982, Solar Magnetohydrodynamics (Dordrecht: D. Reidel Pub. Co.) [Google Scholar]
  22. Priest, E. 2014, Magnetohydrodynamics of the Sun (Cambridge, UK: Cambridge University Press) [Google Scholar]
  23. Priest, E. R., & Hood, A. W. 1991, Advances in Solar System Magnetohydrodynamics (CUP) [Google Scholar]
  24. Ruderman, M. S. 2007, Sol. Phys., 246, 119 [NASA ADS] [CrossRef] [Google Scholar]
  25. Ruderman, M. S. 2015, A&A, 575, A130 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  26. Ryutov, D. A., & Ryutova, M. P. 1976, Sov. J. Exp. Theor. Phys., 43, 491 [NASA ADS] [Google Scholar]
  27. Sewell, G. 2005, The Numerical Solution of Ordinary and Partial Differential Equations, 2nd edn. (New Jersey: Wiley) [Google Scholar]
  28. Solov’ev, A. A. 2011, Astron. Rep., 55, 1025 [NASA ADS] [CrossRef] [Google Scholar]
  29. Solov’ev, A. A. 2012, Astrophys. Space Sci. Proc., 30, 203 [NASA ADS] [CrossRef] [Google Scholar]
  30. Soloviev, L. S. 1975, Rev. Plasma Phys., 6, 239 [NASA ADS] [Google Scholar]
  31. Spruit, H. C. 1982, Sol. Phys., 75, 3 [NASA ADS] [CrossRef] [Google Scholar]
  32. Suydam, B. R. 1958, in Proc. of the Second UN Int. Conf. on the Peaceful Uses of Atomic Energy, Geneva, 31, 157 [Google Scholar]
  33. Terradas, J., & Goossens, M. 2012, A&A, 584, 112 [Google Scholar]
  34. Trubnikov, B. A. 1966, Plasma theory (Energoatomizdat) [in Russian] [Google Scholar]
  35. Vršnak, B. 2008, Annales Geophysicae, 26, 3089 [Google Scholar]
  36. Wesson, J. A. 1978, Nucl. Fusion, 18, 87 [Google Scholar]
  37. Zagorodny, A. G., & Cheremnykh, O. K. 2014, Introduction to the plasma physics, 696 (Kiev: Naukova Dumka) [in Russian] [Google Scholar]

Appendix A: Obtaining an expression for δp1

From the following equations:

δp1=δp+B·δB,δB=×[ξ×B]=(B·)ξ(ξ·)BB(B·ξ).\appendix \setcounter{section}{1} \begin{eqnarray} && \delta p_1=\delta p + \vec{B}\cdot \delta \vec{B}, \nonumber \\ && \delta p=-\bfxi \cdot \vec{\nabla} p -\gamma p \bfbigtriangledown \cdot \bfxi, \label{BB1} \\ && \delta \vec{B}=\bfbigtriangledown \times \left[\bfxi \times \vec{B}\right]=\left(\vec{B}\cdot\bfbigtriangledown \right)\bfxi-\left(\bfxi\cdot\bfbigtriangledown\right)\vec{B}-\vec{B}\left(\vec{B}\cdot\bfxi\right). \nonumber \end{eqnarray}(A.1)By using the equation of equilibrium p=(B·)B(B2/2),\appendix \setcounter{section}{1} \begin{equation} \bfbigtriangledown p=\left(\vec{B}\cdot \bfbigtriangledown\right)\vec{B}-\bfbigtriangledown\left(B^2/2\right), \label{BB11} \end{equation}(A.2)after some algebra from Eq. (A.1) we obtain δp1=ξ·(B·)Bγp(·ξ)+B·(B·)ξB2·ξ.\appendix \setcounter{section}{1} \begin{equation} \delta p_1=-\bfxi\cdot\left(\vec{B}\cdot\bfbigtriangledown\right)\vec{B}-\gamma p \left(\bfbigtriangledown\cdot\bfxi\right)+\vec{B}\cdot\left(\vec{B} \cdot \bfbigtriangledown \right)\bfxi- B^2\bfbigtriangledown\cdot\bfxi. \label{BB2} \end{equation}(A.3)By using the vector relations

ξ·(B·)B+B·(B·)ξ=2ξ·(B·)B+(B·)(ξ·B)=2ξ·(B·)B+·[B(ξ·B)]\appendix \setcounter{section}{1} \begin{eqnarray} && -\bfxi\cdot\left(\vec{B}\cdot \bfbigtriangledown\right)\vec{B}+\vec{B}\cdot\left(\vec{B}\cdot \bfbigtriangledown\right)\bfxi=-2\bfxi\cdot\left(\vec{B\cdot\bfbigtriangledown}\right)\vec{B} +\left(\vec{B}\cdot\bfbigtriangledown\right)\left(\bfxi\cdot\vec{B}\right) \nonumber \\ &&\qquad \qquad \quad\qquad =-2\bfxi\cdot\left(\vec{B}\cdot\bfbigtriangledown\right)\vec{B}+\bfbigtriangledown\cdot\left[\vec{B}\left(\bfxi\cdot\vec{B}\right)\right] \nonumber \\ &&\qquad \qquad \quad\qquad= -2 B\bfxi\cdot\left(\vec{\tau}\cdot\bfbigtriangledown\right)\vec{\tau}B+\bfbigtriangledown\cdot \left[B^2\left(\bfxi\cdot\vec{\tau}\right)\tau\right], \label{BBB12} \end{eqnarray}(A.4)where τ = B/B we can rewrite Eq. (A.3) as

δp1=2Bξ·(τ·)τB+·[B2(ξ·τ)]\appendix \setcounter{section}{1} \begin{eqnarray} && \delta p_1= -2B\bfxi\cdot\left(\tau\cdot\bfbigtriangledown\right)\tau B+\bfbigtriangledown\cdot\left[B^2\left(\bfxi\cdot\tau\right)\right] \nonumber \\ && \qquad\,\, -~ \gamma p\bfbigtriangledown\cdot \bfxi - B^2\bfbigtriangledown\cdot \bfxi. \label{BB3} \end{eqnarray}(A.5)The first term on the RHS of Eq. (A.5) can be transformed as follows:

2Bξ·(τ·)τB=2B2ξ·(τ·)τ2(ξ·τ)B(τ·B)=2B2ξ·κ2ξB(τ·B)\appendix \setcounter{section}{1} \begin{eqnarray} -2B\bfxi\cdot\left(\tau\cdot\bfbigtriangledown\right)\tau B\, &&= -2B^2\bfxi\cdot\left(\tau\cdot \bfbigtriangledown\right)\tau - 2\left(\bfxi\cdot\tau\right)B\left(\tau\cdot\bfbigtriangledown B\right)\nonumber \\ && = -2B^2\bfxi\cdot\bfkappa-2\xi_{\parallel}B\left(\tau\cdot\bfbigtriangledown B\right) \nonumber \\ && =-2B^2\bfxi\cdot\bfkappa-\xi_{\parallel}\left(\tau\cdot\bfbigtriangledown B^2\right). \label{BB4} \end{eqnarray}(A.6)Here κ = (τ· )τ is the vector of curvature of the magnetic field lines. To obtain Eq. (A.6) we take into account that displacement vector ξ is represented in the form ξ = ξ + ξτ, where subscripts and correspond to the perpendicular and parallel components to the equilibrium magnetic field. Therefore, Eq. (A.5) can be rewritten in the form:

δp1=2B2ξ·κξ(τ·B2)+·(ξB2τ)\appendix \setcounter{section}{1} \begin{eqnarray} && \delta p_1 =-2B^2\bfxi\cdot\bfkappa-\xi_{\parallel}\left(\tau\cdot\bfbigtriangledown B^2\right)+\bfbigtriangledown \cdot\left(\xi_{\parallel} B^2 \tau\right) \nonumber \\ &&\quad \,\,\,\, \quad -\,\gamma p \bfbigtriangledown\cdot\xi -B^2 \bfbigtriangledown \bfxi. \label{BB5} \end{eqnarray}(A.7)From the second and third terms in Eq. (A.7) we can obtain

ξ(τ·B2)+·(ξB2τ)=ξ(τ·B2)+ξ(τ·B2)+B2(ξτ)\appendix \setcounter{section}{1} \begin{eqnarray} &&-\xi_{\parallel}\left(\tau\cdot\bfbigtriangledown B^2\right)+\bfbigtriangledown\cdot\left(\xi_{\parallel} B^2 \vec{\tau}\right)=-\xi_{\parallel}\left(\tau\cdot \bfbigtriangledown B^2\right) \nonumber \\ &&\qquad \qquad+\,\xi_{\parallel}\left(\tau\cdot B^2\right)+ B^2 \bfbigtriangledown\left(\xi_{\parallel} \vec{\tau}\right)\notag\\&&\qquad\qquad =B^2\bfbigtriangledown \left(\xi_{\parallel}\vec{\tau}\right). \label{BB6} \end{eqnarray}(A.8)From Eqs. (A.7) and (A.8) it follows that:

δp1=2B2ξ·κ+B2·(ξτ)γp·ξB2ξ=\appendix \setcounter{section}{1} \begin{eqnarray} \delta p_1\,&=&\,-2B^2\bfxi\cdot\bfkappa+B^2\bfbigtriangledown\cdot\left(\xi_{\parallel}\tau\right)-\gamma p \bfbigtriangledown\cdot \bfxi -B^2\bfbigtriangledown\bfxi\notag\\\,& =&\,-2B^2\bfxi\cdot\bfkappa-\gamma p\bfbigtriangledown\cdot \bfxi -B^2\bfbigtriangledown\cdot \bfxi_{\perp}. \label{BB7} \end{eqnarray}(A.9)The first term on the RHS of Eq (A.9) is equal to:

2B2ξ·κ=2B2(ξ+ξτ)(τ·)τ=2B2ξ·κ2B2ξτ·(τ·)τ=2B2ξ·κB2ξ(τ·)1=2B2ξ·κ.\appendix \setcounter{section}{1} \begin{eqnarray} -2B^2\bfxi\cdot\bfkappa&&\,=-2B^2\left(\bfxi_{\perp}+\bfxi_{\parallel}\tau\right)\left(\tau\cdot\bfbigtriangledown\right)\vec{\tau} \nonumber \\ &&\,= -2B^2\bfxi_{\perp}\cdot\bfkappa-2B^2\bfxi_{\parallel}\vec{\tau}\cdot\left(\vec{\tau}\cdot \bfbigtriangledown\right)\vec{\tau} \nonumber \\ &&\,= -2B^2\bfxi_{\perp}\cdot\bfkappa-B^2\bfxi_{\parallel}\left(\vec{\tau}\cdot\bfbigtriangledown\right)1=-2B^2\bfxi_{\perp}\cdot\bfkappa. \end{eqnarray}(A.10)Finally, we obtain the second expression in Eq. (2), that is, δp1=γp·ξB2(·ξ+2ξ·κ).\appendix \setcounter{section}{1} \begin{equation} \delta p_1=-\gamma p \bfbigtriangledown\cdot \bfxi-B^2\left(\bfbigtriangledown\cdot\bfxi_{\perp}+2\bfxi_{\perp}\cdot\bfkappa\right). \label{BB8} \end{equation}(A.11)

Appendix B: Exact dispersion relation for the magnetic field described as Eq. (35) for m02<0\hbox{$m_{0}^ 2< 0$}

To obtain the exact dispersion relation for the case (35) we have used the following set of Eqs: (13) and (19), (20) and (21). From Eqs. (19) and (21) we have:

ρ(ω2ωA2)p1dr=2mr2Bϕ(mrBϕ+kzBz)δp1+ξr{[ρ(ω2ωA2)]2+2ρBϕ(ω2ωA2)ddr(Bϕr)\appendix \setcounter{section}{2} \begin{eqnarray} && \rho\left(\omega^2-\omega_{\rm A}^2\right)\frac{d\delta p_1}{{\rm d}r}=2\frac{m}{r^2}B_{\varphi}\left(\frac{m}{r}B_{\varphi}+k_{z}B_{z}\right)\delta p_1 \nonumber \\ && \qquad \qquad +\,\,\xi_{r}\left\{\left[\rho\left(\omega^2-\omega_{\rm A}^2\right)\right]^2+2\rho B_{\varphi}\left(\omega^2-\omega_{\rm A}^2\right)\frac{{\rm d}}{{\rm d}r}\left(\frac{B_{\varphi}}{r}\right) \right. \nonumber \\ && \qquad \qquad - \left. 4\rho \omega_{\rm A}^2\left(\frac{B_{\varphi}^2}{r^2}\right)\right\}\cdot \label{AA1} \end{eqnarray}(B.1)Equation (B.1) together with Eqs. (19) and (13) represent a system of equations with respect to ξr and δp1. From this system it is easy to obtain the equation for δp1:

d2dr2δp1+[C3rDddr(rDC3)]ddrδp1+[C3rDddr(rC1C3)\appendix \setcounter{section}{2} \begin{eqnarray} && \frac{{\rm d}^2}{{\rm d}r^2}\delta p_1+\left[\frac{C_3}{rD}\frac{{\rm d}}{{\rm d}r}\left(\frac{rD}{C_3}\right)\right]\frac{{\rm d}}{{\rm d}r}\delta p_1+ \left[\frac{C_3}{rD}\frac{{\rm d}}{{\rm d}r}\left(\frac{rC_1}{C_3}\right)\right. \nonumber \\ &&\qquad \qquad\qquad\qquad \qquad+\left. \frac{1}{D^2}\left(C_2C_3-C_{1}^2\right)\right]\delta p_1=0, \label{AA2} \end{eqnarray}(B.2)where

D=ρ(ω2ωA2),C1=2mBϕr2(mrBϕ+kzBz),C2=(m2r2+kz2),\appendix \setcounter{section}{2} \begin{eqnarray} && D=\rho\left(\omega^2-\omega_{\rm A}^2\right), \quad C_1=-2\frac{mB_{\varphi}}{r^2}\left(\frac{m}{r}B_{\varphi}+k_{z}B_{z}\right), \nonumber \\ && C_2=-\left(\frac{m^2}{r^2}+k_{z}^2\right), \nonumber \\ && C_3=D^2+2DB_{\varphi}\frac{{\rm d}}{{\rm d}r}\left(\frac{B_{\varphi}}{r}\right)-4\rho\omega_{\rm A}^2\left(\frac{B_{\varphi}^2}{r^2}\right)\cdot \label{AA3} \end{eqnarray}(B.3)From Eq. (B.3) it follows that, for the magnetic field (35), the following relations are satisfied: ρωA2=const.,Bϕr=const.,mrBϕ+kzBz=const.\appendix \setcounter{section}{2} \begin{equation} \rho\omega_{\rm A}^2={\rm const.}, \, \frac{B_{\varphi}}{r}={\rm const.}, \, \frac{m}{r}B_{\varphi}+k_z B_z={\rm const.} \label{AA4} \end{equation}(B.4)Therefore, D=const.,rC1=const.,C3=const.\appendix \setcounter{section}{2} \begin{equation} D={\rm const.}, \, rC_1={\rm const.}, \, C_3={\rm const.} \label{AA5} \end{equation}(B.5)By taking into account Eqs. (B.4) and Eq. (B.5), from (B.2) we obtain

d2dr2δp1+1rddrδp1(m2r2+kz2)δp1\appendix \setcounter{section}{2} \begin{eqnarray} && \frac{{\rm d}^2}{{\rm d}r^2}\delta p_1+\frac{1}{r}\frac{{\rm d}}{{\rm d}r}\delta p_1-\left(\frac{m^2}{r^2}+k_{z}^2\right)\delta p_1 \nonumber \\ &&\qquad\qquad\qquad + \frac{4\left(B_{\varphi}^2/r^2\right)\omega_{\rm A}^2k_{z}^2}{\rho\left(\omega^2-\omega_{\rm A}^2\right)^2}\delta p_1=0. \label{AA6} \end{eqnarray}(B.6)Equation (B.6) coincides with the Eq. (13) of work Bennett et al. (1999) and satisfies the boundary conditions (23) and (25).

Following Bennett et al. (1999), we introduce m02\hbox{$m_{0}^2$} (see (43)). We assume m02<0\hbox{$m_{0}^2<0$}, that gives a solution (B.6) in the form δp1={AiJm(|m0|r),raAeKm(|kz|r),r>a.\appendix \setcounter{section}{2} \begin{equation} \delta p_1 = \left\{ \begin{array}{lr} A_{\rm i} J_m\left(\left|m_0\right|r\right), & r\le a\\ A_{\rm e} K_m\left(\left|k_z\right|r\right), & r > a. \end{array} \right. \label{AA7} \end{equation}(B.7)To obtain Eq. (B.7) we take into account that m0 = |kz| for r>a. From Eqs. (B.1), (23), (25) and (B.7), we obtain desired dispersion relation: (ω2ωA2)xJm(x)dJm(x)dx|x=|m0|a2mBϕ(a)ωAiaρi(ω2ωAi2)24ωAi2Bϕ2(a)a2ρi=\appendix \setcounter{section}{2} \begin{eqnarray} && \frac{\left. \left(\omega^2-\omega_{\rm A}^2\right)\frac{x}{J_{m}(x)}\frac{{\rm d}J_{m}(x)}{{\rm d}x}\right|_{x=\left|m_0\right|a}-\frac{2mB_{\varphi}(a)\omega_{\rm Ai}}{a\sqrt{\rho_{\rm i}}}}{{\left(\omega^2-\omega_{\rm Ai}^2\right)^2-4\omega_{\rm Ai}^2}\frac{B_{\varphi}^2(a)}{a^2\rho_{\rm i}}}= \nonumber \\ && \qquad \frac{\left.\frac{y}{K_{m}(y)}\frac{{\rm d}K_{m}(y)}{{\rm d}y}\right|_{y=\left|k_z\right|a}}{\frac{\rho_{\rm e}}{\rho_{\rm i}}\left(\omega^2-\omega_{\rm Ae}^2\right)+\left. \frac{B_{\varphi}^2(a)}{a^2\rho_{\rm i}}\frac{y}{K_{m}(y)}\frac{{\rm d}K_{m}(y)}{{\rm d}y}\right|_{y=\left|k_z\right|a}}\cdot \label{AA8} \end{eqnarray}(B.8)Used in Eq. (B.8) variables and parameters are explained in Eq. (43).

All Figures

thumbnail Fig. 1

Magneto-hydrostatic equilibrium of the twisted magnetic flux tube.

In the text
thumbnail Fig. 2

Radially dependent eigenfunctions ξr(r) for axially symmetric magnetic flux tube with a free boundary are shown for azimuthal wave numbers m = 1, 2, 3, and 4.

In the text

Current usage metrics show cumulative count of Article Views (full-text article views including HTML views, PDF and ePub downloads, according to the available data) and Abstracts Views on Vision4Press platform.

Data correspond to usage on the plateform after 2015. The current usage metrics is available 48-96 hours after online publication and is updated daily on week days.

Initial download of the metrics may take a while.