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

Phase Mixing of Kink MHD Waves in the Solar Corona: Viscous Dissipation and Heating

Zanyar Ebrahimi1, Roberto Soler 2,3 and Kayoomars Karami 4
1Research Institute for Astronomy &\& Astrophysics of Maragha, University of Maragheh, P. O. Box 55136-553, Maragheh, Iran
2Departament de Física, Universitat de les Illes Balears, E-07122, Palma de Mallorca, Spain
3Institut d’Aplicacions Computacionals de Codi Comunitari (IAC3IAC^{3}), Universitat de les Illes Balears, E-07122, Palma de Mallorca, Spain
4Department of Physics, University of Kurdistan, Pasdaran Street, P.O. Box 66177-15175, Sanandaj, Iran
E-mail: [email protected]
Abstract

Magnetohydrodynamic (MHD) kink waves have been observed frequently in solar coronal flux tubes, which makes them a great tool for seismology of the solar corona. Here, the effect of viscosity is studied on the evolution of kink waves. To this aim, we solve the initial value problem for the incompressible linearized viscous MHD equations in a radially inhomogeneous flux tube in the limit of long wavelengths. Using a modal expansion technique the spatio-temporal behavior of the perturbations is obtained. We confirm that for large Reynolds numbers representative of the coronal plasma the decrement in the amplitude of the kink oscillations is due to the resonant absorption mechanism that converts the global transverse oscillation to rotational motions in the inhomogeneous layer of the flux tube. We show that viscosity suppresses the rate of phase mixing of the perturbations in the inhomogeneous region of the flux tube and prevents the continuous building up of small scales in the system once a sufficiently small scale is reached. The viscous dissipation function is calculated to investigate plasma heating by viscosity in the inhomogeneous layer of the flux tube. For Reynolds numbers of the order of 10610^{6}10810^{8}, the energy of the kink wave is transformed into heat in two to eight periods of the kink oscillation. For larger and more realistic Reynolds numbers, heating happens, predominantly, after the global kink oscillation is damped, and no significant heating occurs during the observable transverse motion of the flux tube.

Key words: Sun: corona – Sun: magnetic fields – Sun: oscillations.

1 Introduction

Aschwanden et al. (1999) and Nakariakov et al. (1999) were the first to identify the spatially resolved kink oscillations of coronal loops using the Transitional Region And Coronal Explorer (TRACE) observations in the 171 Å{\AA} Fe IX emission lines. Because of the large number of observations of kink oscillations in coronal loops these oscillations are a great seismological tool to estimate the parameters of the solar coronal plasma such as the magnetic field, the plasma density, and the transport coefficients.

An interesting characteristic of kink oscillations in coronal loops is that they damp fast usually within 3-5 periods. Nakariakov et al. (1999) speculated that an anomalously high viscosity or resistivity in the coronal plasma could be responsible for this rapid damping. However, among the proposed mechanisms to justify the rapid damping of the kink waves (see e.g. Goossens et al. 2002; Ofman & Aschwanden 2002; Ruderman & Roberts 2002; Ofman 2005, 2009; Morton & Erdélyi 2009, 2010) resonant absorption is the strongest candidate, since, unlike other theories, resonant absorption is the only proposed mechanism that is able to explain short damping times (of the order of a few periods of the kink oscillation) without invoking anomalous processes or values of the dissipative coefficients many orders of magnitude larger than those expected in the corona. Resonant absorption is an ideal process that does not need strong diffusion to work. However, there is still no observational evidence for resonant absorption. This mechanism was first proposed by Ionson (1978) as a heating mechanism in coronal loops. Since then, many studies have developed the theory of resonant absorption (see e.g. Davila 1987; Sakurai et al. 1991a, 1991b; Goossens et al. 1995; Goossens & Ruderman 1995; Erdélyi 1997; Cally & Andries 2010 among many others). The necessary condition in this mechanism is that the wave frequency lies in the local Alfvén and/or slow frequency continuum. In this situation the energy of the global mode oscillation transfers to the local perturbations in the inhomogeneous regions of the magnetic flux tube. As a result, the amplitude of the perturbations grows at the resonance point and the dissipation mechanisms become important in the resonance layer, where the oscillations make large gradients. Ruderman & Roberts (2002) studied damping of kink oscillations in coronal loops. Considering the effect of viscosity in their analysis, they confirmed the previously obtained numerical result (see e.g. Poedts & Kerner 1991; Tirry & Goossens 1996) that the decay rate of the transverse oscillation is independent of the Reynolds number RvR_{v} when Rv1R_{v}\gg 1. They concluded that Reynolds number affects only the perturbations in the resonance layer so that it is not possible to obtain the value of viscosity from the observations of decaying kink oscillations in coronal loops. Resonant absorption has been studied for various complex configurations of the magnetic flux tubes including curvature of the flux tube (Terradas et al. 2006), longitudinal density stratification (Andries et al. 2005; Karami & Asvar 2007; Soler et al. 2011), twisted magnetic field (Karami & Bahari 2010; Ebrahimi & Karami 2016; Ebrahimi & Bahari 2019) and magnetic field expansion (Shukhobodskiy et al. 2018; Howson et al. 2019). For a review on the theory of resonant absorption, see Goossens et al. (2011).

Another consequence of existing a continuum of Alfvén frequencies in the flux tubes may be the phase mixing of the Alfvén waves (Heyvaerts & Priest 1983; Ireland & Priest 1997; Karami & Ebrahimi 2009; Prokopyszyn & Hood 2019). In this mechanism due to inhomogeneity of the local Alfvén phase speed across the background magnetic field the perturbations on different magnetic surfaces become out of phase as travel in the case of propagating wave or oscillate in the case of standing wave. In the developed stage of phase mixing even with a small amount of viscosity or resistivity the dissipation mechanisms become important and could transform the wave energy to heat. Ofman & Aschwanden (2002) suggested that the loop oscillations are dissipated by phase mixing with viscosity of the order ν=105.3±3.5m2s1\nu=10^{5.3\pm 3.5}~{}m^{2}s^{-1} that is anomalously many order of magnitudes higher than the classical coronal value of the shear viscosity, ν=1m2s1\nu=1~{}m^{2}s^{-1} (Ofman et al. 1994). It is believed that some small-scale turbulence and structure enhance the viscosity in the coronal loop plasma. Ofman et al. (1994) investigated the effect of viscous stress tensor on the heating of the corona by the resonant absorption and showed that the shear viscosity has the dominant role in the heating process but the compressive viscosity does not have a significant contribution (see also Erdélyi & Goossens 1995, 1996).

Goossens et al. (2014) investigated the nature of kink waves and stated that kink waves do not only involve purely transverse motions of solar magnetic flux tubes, but the velocity field is a spatially and temporally varying sum of both transverse and rotational motion. In an axisymmetric cylindrical flux tube, wave modes can be classified according to the value of the azimuthal wavenumber, mm. In this paper we study modes with m=1m=1, i.e., kink modes. The global kink mode is the only mode that is able to displace transversely (i.e., laterally) the axis of the cylinder. The global mode with m=1m=1 is resonantly coupled to Alfvén modes with m=1m=1 in the nonuniform layer of the tube. These modes have both radial and azimuthal components of the displacement. Goossens et al. (2014) called the Alfvén modes with m0m\neq 0 as ”rotational modes”. The reason for calling these modes ”rotational” is that their streamlines follow a closed curve, so that a fluid element flowing along one of those streamlines would describe a ”rotation” around a certain point. In the case of m=1m=1, the center of the rotation is not located on the axis of the cylinder (as happens for torsional motions with ”m=0”), but at some place in between the axis and the boundary of the tube. Following Goossens et al. (2014), we use the term ”transverse motion” to refer the lateral displacement of the flux tube caused by the global kink mode and the term ”rotational motion” to refer the local Alfvén perturbations inside the tube. Soler & Terradas (2015, hereafter ST2015) investigated the resonant absorption of the kink MHD wave and phase mixing of its perturbations in coronal flux tubes. Using a modal expansion method they showed that the energy of the global kink oscillation of the flux tube is transformed into small-scale rotational motions in the nonuniform boundary of the tube which are eventually subject to the simultaneously occurring phase mixing process. However, ST2015 used the ideal MHD equations and did not consider the effect of dissipation terms in their analysis. At the developed stage of the phase mixing process where the perturbations are highly phase mixed, the dissipation mechanisms could suppress the rate of generating small scales in the system by coupling the perturbations on the neighboring magnetic surface and finally transform the kink wave energy to heat.

In this paper, our aim is to investigate the effect of viscosity on the kink MHD waves and show how viscosity modifies the previous results obtained by ST2015. Section 2 presents the model and the governing MHD equations of motion. In section 3 we apply and extend the mathematical method used by ST2015 and give a solution to the equation of motion. The results are discussed in section 4. Finally we conclude the paper in section 5.

2 Equations of motion and model

We model a typical coronal loop by a straight cylinder that has a circular cross section of radius RR. The background plasma density in cylindrical coordinates (rr, φ\varphi, zz) is assumed to be as follows

ρ(r)={ρi,rr1,ρi2[(1+ρeρi)(1ρeρi)sin(πl(rR))],r1<r<r2,ρe,rr2,\rho(r)=\left\{\begin{array}[]{lll}\rho_{{\rm i}},&r\leqslant r_{1},\\ \frac{\rho_{{\rm i}}}{2}\left[\left(1+\frac{\rho_{{\rm e}}}{\rho_{{\rm i}}}\right)-\left(1-\frac{\rho_{{\rm e}}}{\rho_{{\rm i}}}\right)\sin\left(\frac{\pi}{l}(r-R)\right)\right],&r_{1}<r<r_{2},\\ \rho_{{\rm e}},&r\geqslant r_{2},\end{array}\right. (1)

where r1=Rl/2r_{1}=R-l/2 and r2=R+l/2r_{2}=R+l/2. Here, l=r2r1l=r_{2}-r_{1} is the width of the inhomogeneous region. The background magnetic field is assumed to be constant and aligned with the flux tube axis everywhere, i.e. 𝐁=B0z^\mathbf{B}=B_{{\rm 0}}\hat{z} where B0B_{0} is constant.

The linearized MHD equations for an incompressible plasma with viscosity are as follows

ρ(r)2𝝃t2=p+1μ0(×𝐁)×𝐁+ρ(r)ν2t𝝃,\rho(r)\frac{\partial^{2}\boldsymbol{\xi}}{\partial t^{2}}=-\nabla p^{\prime}+\frac{1}{\mu_{0}}(\nabla\times\mathbf{B^{\prime}})\times{\mathbf{B}}+\rho(r)\nu\nabla^{2}\frac{\partial}{\partial t}\boldsymbol{\xi}, (2)
𝐁=×(𝝃×𝐁),\mathbf{B^{\prime}}=\nabla\times(\boldsymbol{\xi}\times\mathbf{B}), (3)
𝝃=0,\nabla\cdot\boldsymbol{\xi}=0, (4)

where 𝝃\boldsymbol{\xi} is the lagrangian displacement of the plasma, 𝐁\mathbf{B^{\prime}} and pp^{\prime} are the Eulerian perturbations of the magnetic field and plasma pressure, respectively. Here μ0\mu_{0} is the magnetic permeability of the free space and ν\nu is coefficient of the kinematic shear viscosity which is assumed to be uniform. Using Eq. (4), we can rewrite Eq. (3) as

𝐁=(𝐁)𝝃.\mathbf{B^{\prime}}=(\mathbf{B}\cdot\nabla)\boldsymbol{\xi}. (5)

Substituting 𝐁\mathbf{B^{\prime}} from Eq. (5) into (2) yields

ρ(r)2𝝃t2=P+1μ0B022z2𝝃+ρ(r)ν2t𝝃,\rho(r)\frac{\partial^{2}\boldsymbol{\xi}}{\partial t^{2}}=-\nabla P^{\prime}+\frac{1}{\mu_{0}}B_{0}^{2}\frac{\partial^{2}}{\partial z^{2}}\boldsymbol{\xi}+\rho(r)\nu\nabla^{2}\frac{\partial}{\partial t}\boldsymbol{\xi}, (6)

or in a more compact form

A𝝃ρ(r)ν2t𝝃=P,\mathcal{L}_{A}\boldsymbol{\xi}-\rho(r)\nu\nabla^{2}\frac{\partial}{\partial t}\boldsymbol{\xi}=-\nabla P^{\prime}, (7)

where P=p+(𝐁𝐁)/μ0P^{\prime}=p^{\prime}+(\mathbf{B^{\prime}}\cdot\mathbf{B})/\mu_{0} is the Eulerian perturbation of the total (gas plus magnetic) pressure and

Aρ(r)2t21μ0B022z2,\mathcal{L}_{A}\equiv\rho(r)\frac{\partial^{2}}{\partial t^{2}}-\frac{1}{\mu_{0}}B_{0}^{2}\frac{\partial^{2}}{\partial z^{2}}, (8)

is the Alfvén operator. In the absence of viscosity, Eq. (7) reduces to equation (8) of ST2015. Since the equilibrium quantities are only functions of rr, the perturbations can be Fourier-analyzed with respect to the φ\varphi and zz coordinates. Hence,

P=P(r,t)ei(mφ+kzz),𝝃=𝝃(r,t)ei(mφ+kzz),\begin{split}P^{\prime}=&P^{\prime}(r,t)~{}e^{i(m\varphi+k_{z}z)},\\ \boldsymbol{\xi}=&\boldsymbol{\xi}(r,t)~{}e^{i(m\varphi+k_{z}z)},\end{split} (9)

where mm is the azimuthal mode number and kzk_{z} is the axial wave number. The three components of Eq. (7) and the incompressibility condition (Eq. 4) form a system of 4 independent equations for ξr\xi_{r}, ξφ\xi_{\varphi}, ξz\xi_{z} and PP^{\prime} as follows

Aξrρ(r)νt(2ξrξrr22r2φξφ)=rP,\displaystyle\mathcal{L}_{A}\xi_{r}-\rho(r)\nu\frac{\partial}{\partial t}\left(\nabla^{2}\xi_{r}-\frac{\xi_{r}}{r^{2}}-\frac{2}{r^{2}}\frac{\partial}{\partial\varphi}\xi_{\varphi}\right)=-\frac{\partial}{\partial r}P^{\prime}, (10)
Aξφρ(r)νt(2ξφξφr2+2r2φξr)=1rφP,\displaystyle\mathcal{L}_{A}\xi_{\varphi}-\rho(r)\nu\frac{\partial}{\partial t}\left(\nabla^{2}\xi_{\varphi}-\frac{\xi_{\varphi}}{r^{2}}+\frac{2}{r^{2}}\frac{\partial}{\partial\varphi}\xi_{r}\right)=-\frac{1}{r}\frac{\partial}{\partial\varphi}P^{\prime}, (11)
Aξzρ(r)νt2ξz=zP,\displaystyle\mathcal{L}_{A}\xi_{z}-\rho(r)\nu\frac{\partial}{\partial t}\nabla^{2}\xi_{z}=-\frac{\partial}{\partial z}P^{\prime}, (12)
1rr(rξr)+1rφξφ+zξz=0.\displaystyle\frac{1}{r}\frac{\partial}{\partial r}\left(r\xi_{r}\right)+\frac{1}{r}\frac{\partial}{\partial\varphi}\xi_{\varphi}+\frac{\partial}{\partial z}\xi_{z}=0. (13)

Here, we apply 1rφ\frac{1}{r}\frac{\partial}{\partial\varphi} and z\frac{\partial}{\partial z} from left on Eqs. (11) and (12), respectively and add the resulting equations together. After that with the help of Eqs. (9) and (13) we obtain PP^{\prime} in terms of ξr\xi_{r} and ξφ\xi_{\varphi} as

P=1m2/r2+kz2[A(1r(rξr)r)ρ(r)νt(2(1rr(rξr))+(2r32r2r)φξφ2r32φ2ξr)].\begin{split}P^{\prime}&=\frac{-1}{m^{2}/r^{2}+k_{z}^{2}}\left[\mathcal{L}_{A}\left(\frac{1}{r}\frac{\partial(r\xi_{r})}{\partial r}\right)-\right.\\ &\left.\rho(r)\nu\frac{\partial}{\partial t}\left(\nabla^{2}\left(\frac{1}{r}\frac{\partial}{\partial r}(r\xi_{r})\right)+\left(\frac{2}{r^{3}}-\frac{2}{r^{2}}\frac{\partial}{\partial r}\right)\frac{\partial}{\partial\varphi}\xi_{\varphi}-\frac{2}{r^{3}}\frac{\partial^{2}}{\partial\varphi^{2}}\xi_{r}\right)\right].\end{split} (14)

Now we use the thin tube (TT) approximation in which the wavelength of the kink waves, λ\lambda, is much larger than the radius of the cross section of the flux tube, R=(r1+r2)/2R=(r_{1}+r_{2})/2, i.e. Rkz1Rk_{z}\ll 1 or RλR\ll\lambda. The first two terms of Eq. (13) have magnitudes of the order ξ0/R\xi_{0}/R where ξ0\xi_{0} is a typical value for the Lagrangian displacement. Since the characteristic length scale in the zz direction is λ\lambda, the order of the magnitude of the third term of Eq. (13) is equal to ξ0/λ\xi_{0}/\lambda. Hence, in TT approximation neglecting the third term of Eq. (13) with respect to the first two terms, yields

φξφr(rξr).\frac{\partial}{\partial\varphi}\xi_{\varphi}\simeq-\frac{\partial}{\partial r}(r\xi_{r}). (15)

We use Eq. (15) to obtain a single equation for ξr\xi_{r}. Since we use this approximation, it is not possible to obtain ξz\xi_{z} with respect to ξr\xi_{r} from Eqs. (10)-(13). However, in the TT approximation, Goossens et al. (2009) showed that the longitudinal component of the displacement, ξz\xi_{z}, is always much smaller than the other components and the dominant motion is in the horizontal plane normal to the background magnetic field. Substituting Eq. (15) in Eqs. (10) and (14) and eliminating PP^{\prime} from the resulting equations, gives the equation for ξr\xi_{r} as follows

Asξr+(m2r2+kz2)dρ(r)dr2t21rr(rξr)=νtdξr,\mathcal{L}_{A}\mathcal{L}_{s}\xi_{r}+\left(\frac{m^{2}}{r^{2}}+k_{z}^{2}\right)\frac{d\rho(r)}{dr}\frac{\partial^{2}}{\partial t^{2}}\frac{1}{r}\frac{\partial}{\partial r}(r\xi_{r})=\nu\frac{\partial}{\partial t}\mathcal{L}_{d}\xi_{r}, (16)

where s\mathcal{L}_{s} is the surface wave operator (see ST2015 for more details) and d\mathcal{L}_{d} is the viscous damping operator which are defined as follows

s(kz2+m2r2)2r2+1r(kz2+3m2r2)r1r2(kz2m2r2)(kz2+m2r2)2,\mathcal{L}_{s}\equiv\left(k_{z}^{2}+\frac{m^{2}}{r^{2}}\right)\frac{\partial^{2}}{\partial r^{2}}+\frac{1}{r}\left(k_{z}^{2}+\frac{3m^{2}}{r^{2}}\right)\frac{\partial}{\partial r}-\frac{1}{r^{2}}\left(k_{z}^{2}-\frac{m^{2}}{r^{2}}\right)-\left(k_{z}^{2}+\frac{m^{2}}{r^{2}}\right)^{2}, (17)
d(m2r2+kz2)2[ρ(r)(21r2+2r2(1+rr))r(ρ(r)m2/r2+kz2[3r3+4r2r2(m2r2+kz2)r1r(m2+1r2+kz2)])].\begin{split}\mathcal{L}_{d}\equiv&\left(\frac{m^{2}}{r^{2}}+k_{z}^{2}\right)^{2}\left[\rho(r)\left(\nabla^{2}-\frac{1}{r^{2}}+\frac{2}{r^{2}}\left(1+r\frac{\partial}{\partial r}\right)\right)\right.\\ &\left.-\frac{\partial}{\partial r}\left(\frac{\rho(r)}{m^{2}/r^{2}+k_{z}^{2}}\left[\frac{\partial^{3}}{\partial r^{3}}+\frac{4}{r}\frac{\partial^{2}}{\partial r^{2}}-\left(\frac{m^{2}}{r^{2}}+k_{z}^{2}\right)\frac{\partial}{\partial r}-\frac{1}{r}\left(\frac{m^{2}+1}{r^{2}}+k_{z}^{2}\right)\right]\right)\right].\end{split} (18)

In the absence of viscosity, Eq. (16) consistently reverts to Eq. (16) of ST2015. From Eq. (15) we find that ξφ\xi_{\varphi} is related to ξr\xi_{r} as

iξφ1mr(rξr).i\xi_{\varphi}\simeq-\frac{1}{m}\frac{\partial}{\partial r}(r\xi_{r}). (19)

In this relation, the factor ii accounts for a phase difference of π/2\pi/2 between ξφ\xi_{\varphi} and ξr\xi_{r}. So, for convenience, in order to avoid imaginary terms in the calculations, in the rest of the paper we redefine iξφi\xi_{\varphi} as ξφ\xi_{\varphi}.

3 Solution

3.1 Solution in the uniform regions (rr1r\leqslant r_{1}, rr2r\geqslant r_{2})

In the limit of small viscosity which is the case in the solar corona, we can neglect the effect of viscosity in the interior and exterior regions of the flux tube, because viscous effects are only important in the inhomogeneous regions where phase mixing operates (Heyvaerts & Priest 1983). So, following ST2015 in TT approximation (Rkz1Rk_{z}\ll 1) solutions of ξr\xi_{r} representing the kink (m=1m=1) waves in the constant density regions, i.e. rr1r\leqslant r_{1} and rr2r\geqslant r_{2}, are as follows

ξr(r,t)Ai(t),rr1,\displaystyle\xi_{r}(r,t)\approx A_{\rm i}(t),~{}~{}~{}~{}~{}~{}~{}r\leqslant r_{1}, (20)
ξr(r,t)Ae(t)r2,rr2,\displaystyle\xi_{r}(r,t)\approx A_{\rm e}(t)r^{-2},~{}~{}r\geqslant r_{2}, (21)

where Ai(t)A_{\rm i}(t) and Ae(t)A_{\rm e}(t) are the time-dependent amplitudes.

3.2 Solution in the nonuniform region (r1<r<r2r_{1}<r<r_{2})

In the nonuniform region r1<r<r2r_{1}<r<r_{2}, following ST2015, we perform a modal expansion of the radial component of the Lagrangian displacement ξr(r,t)\xi_{r}(r,t) as

ξr(r,t)=n=1an(t)ψn(r).\xi_{r}(r,t)=\sum_{n=1}^{\infty}a_{n}(t)\psi_{n}(r). (22)

In cylindrical geometry, it is appropriate to set functions ψn(r)\psi_{n}(r) as the orthogonal eigenfunctions of the regular Sturm-Liouville system defined by the following Bessel differential equation

d2ψdr2+1rdψdr+(α21r2)ψ=0.\frac{{\rm d}^{2}\psi}{{\rm d}r^{2}}+\frac{1}{r}\frac{{\rm d}\psi}{{\rm d}r}+\left(\alpha^{2}-\frac{1}{r^{2}}\right)\psi=0. (23)

The boundary conditions are the continuity of the radial displacement of the plasma, ξr\xi_{r}, and Lagrangian perturbation of total pressure, δP=P+ξrdP0/dr\delta P=P^{\prime}+\xi_{r}dP_{0}/dr, at r=r1r=r_{1} and r=r2r=r_{2}. Here, P0P_{0} is the equilibrium total (gas + magnetic) pressure. For the equilibrium presented in this paper, dP0/dr=(1/μ0)(𝐁)𝐁=0dP_{0}/dr=(1/\mu_{0})(\mathbf{B}\cdot\nabla)\mathbf{B}=0. Hence δP=P\delta P=P^{\prime} and the continuity of the Lagrangian perturbation of total pressure, δP\delta P, is satisfied by the continuity of the Eulerian perturbation of total pressure, PP^{\prime}. Neglecting the viscous term in Eq. (14) at the boundaries of the inhomogeneous region i.e. r=r1r=r_{1} and r=r2r=r_{2}, one can find that the remaining terms are proportional to ξr\xi_{r} or ξr/r\partial\xi_{r}/\partial r. Hence, the continuity of δP\delta P at the boundaries is satisfied with the continuity of ξr\xi_{r} and ξr/r\partial\xi_{r}/\partial r. From the continuity of ξr\xi_{r} at r=r1r=r_{1} and r=r2r=r_{2} we obtain amplitudes of ξr\xi_{r} inside and outside the tube, i.e. Ai(t)A_{\rm i}(t) and Ae(t)A_{\rm e}(t), respectively as follows

Ai(t)=n=1an(t)ψn(r)|r=r1,A_{\rm i}(t)=\left.\sum_{n=1}^{\infty}a_{n}(t)\psi_{n}(r)\right|_{r=r_{1}}, (24)
Ae(t)=r2n=1an(t)ψn(r)|r=r2.A_{\rm e}(t)=r^{2}\left.\sum_{n=1}^{\infty}a_{n}(t)\psi_{n}(r)\right|_{r=r_{2}}. (25)

From the continuity of ξr/r\partial\xi_{r}/\partial r at r=r1r=r_{1} and r=r2r=r_{2} we get

n=1an(t)dψn(r)dr|r=r1=0,\left.\sum_{n=1}^{\infty}a_{n}(t)\frac{d\psi_{n}(r)}{dr}\right|_{r=r_{1}}=0, (26)
Ae(t)=r32n=1an(t)dψn(r)dr|r=r2.A_{\rm e}(t)=-\frac{r^{3}}{2}\left.\sum_{n=1}^{\infty}a_{n}(t)\frac{d\psi_{n}(r)}{dr}\right|_{r=r_{2}}. (27)

Subtracting Eq. (27) from Eq. (25) and multiplying the resulting equation by 2r32r^{-3} we get

n=1an(t)(2rψn(r)+dψn(r)dr)|r=r2=0.\left.\sum_{n=1}^{\infty}a_{n}(t)\left(\frac{2}{r}\psi_{n}(r)+\frac{d\psi_{n}(r)}{dr}\right)\right|_{r=r_{2}}=0. (28)

Since the functions an(t)a_{n}(t) in Eqs. (26) and (28) are linearly independent, their coefficients must be zero, namely,

dψn(r)dr|r=r1=0,\displaystyle\left.\frac{d\psi_{n}(r)}{dr}\right|_{r=r_{1}}=0, (29)
(2rψn(r)+dψn(r)dr)|r=r2=0.\displaystyle\left.\left(\frac{2}{r}\psi_{n}(r)+\frac{d\psi_{n}(r)}{dr}\right)\right|_{r=r_{2}}=0. (30)

We use these equations as the boundary conditions governing ψn(r)\psi_{n}(r) at r=r1r=r_{1} and r=r2r=r_{2}. Functions ψn(r)\psi_{n}(r) also satisfy the following orthogonality condition

1lr1r2ψn(r)ψn(r)rdr=δnnn,n{1,2,3,}.\frac{1}{l}\int_{r_{1}}^{r_{2}}\psi_{n}(r)\psi_{n^{\prime}}(r)r{\rm d}r=\delta_{nn^{\prime}}~{}~{}~{}\forall\ n,n^{\prime}\in\{1,2,3,\ldots\}. (31)

For more details on the solution of ψn(r)\psi_{n}(r) see section 3.2 of ST2015.

In order to calculate the time dependent coefficients an(t)a_{n}(t), we must truncate the infinite series of Eq. (22) to a finite number N of terms. Substituting Eq. (22) into (16) we obtain

[ρ(r)sψn(r)+(m2r2+kz2)dρdr(ψn(r)r+dψn(r)dr)]d2an(t)dt2νDψn(r)dan(t)dt+B2kz2μsψn(r)an(t)=0.\begin{split}&\left[\rho(r)\mathcal{L}_{s}\psi_{n}(r)+\left(\frac{m^{2}}{r^{2}}+k_{z}^{2}\right)\frac{d\rho}{dr}\left(\frac{\psi_{n}(r)}{r}+\frac{d\psi_{n}(r)}{dr}\right)\right]\frac{d^{2}a_{n}(t)}{dt^{2}}\\ &-\nu\mathcal{L}_{D}\psi_{n}(r)\frac{da_{n}(t)}{dt}+\frac{B^{2}k_{z}^{2}}{\mu}\mathcal{L}_{s}\psi_{n}(r)a_{n}(t)=0.\end{split} (32)

Multiplying Eq. (32) by ψn(r)\psi_{n^{\prime}}(r) and integrating the resulting equation over the interval [r1,r2][r_{1},r_{2}] we obtain the following matrix equation

𝕄a¨(t)+𝔾a˙(t)+a(t)=0,\mathbb{M}\ddot{\vec{a}}(t)+\mathbb{G}\dot{\vec{a}}(t)+\mathbb{H}\vec{a}(t)=0, (33)

where 𝕄\mathbb{M}, 𝔾\mathbb{G} and \mathbb{H} are square matrices of order NN defined as follows

Mnn=1lr1r2[ρ(r)sψn(r)+dρdr(kz2+m2r2)(ψn(r)r+dψn(r)dr)]ψn(r)rdr,\displaystyle M_{nn^{\prime}}=\frac{1}{l}\int_{r_{1}}^{r_{2}}\left[\rho(r)\mathcal{L}_{s}\psi_{n^{\prime}}(r)+\frac{{\rm d}\rho}{{\rm d}r}\left(k_{z}^{2}+\frac{m^{2}}{r^{2}}\right)\left(\frac{\psi_{n^{\prime}}(r)}{r}+\frac{{\rm d}\psi_{n^{\prime}}(r)}{{\rm d}r}\right)\right]\psi_{n}(r)r{\rm d}r, (34)
Gnn=ν1lr1r2ψn(r)dψn(r)r𝑑r,\displaystyle G_{nn^{\prime}}=-\nu\frac{1}{l}\int_{r_{1}}^{r_{2}}\psi_{n}(r)\mathcal{L}_{d}\psi_{n^{\prime}}(r)rdr, (35)
Hnn=kz2B02μ1lr1r2ψn(r)sψn(r)rdr,\displaystyle\begin{split}&H_{nn^{\prime}}=k_{z}^{2}\frac{B_{0}^{2}}{\mu}\frac{1}{l}\int_{r_{1}}^{r_{2}}\psi_{n}(r)\mathcal{L}_{s}\psi_{n^{\prime}}(r)r{\rm d}r,\end{split} (36)

and a(t)\vec{a}(t) is a column vector defined as

a(t)=[a1(t),a2(t),,aN(t)]T,\vec{a}(t)=\left[a_{1}(t),a_{2}(t),\ldots,a_{N}(t)\right]^{T}, (37)

in which, the superscript TT denotes the transpose. The dot and double dot signs in Eq. (33) represent the first and the second derivative with respect to tt, respectively. We rewrite Eq. (33) in the following form

[𝕄𝕄𝔾][a¨(t)a˙(t)]+[𝕄][a˙(t)a(t)]=0,\left[\begin{matrix}\emptyset&\mathbb{M}\\ \mathbb{M}&\mathbb{G}\end{matrix}\right]\left[\begin{matrix}\ddot{\vec{a}}(t)\\ \dot{\vec{a}}(t)\end{matrix}\right]+\left[\begin{matrix}-\mathbb{M}&\emptyset\\ \emptyset&\mathbb{H}\end{matrix}\right]\left[\begin{matrix}\dot{\vec{a}}(t)\\ \vec{a}(t)\end{matrix}\right]=0, (38)

where \emptyset is the zero square matrix of order NN. Using the following definitions

𝔸[𝕄𝕄𝔾],𝔹[𝕄],b(t)(a˙(t)a(t)),\mathbb{A}\equiv\left[\begin{matrix}\emptyset&\mathbb{M}\\ \mathbb{M}&\mathbb{G}\end{matrix}\right],~{}~{}\mathbb{B}\equiv\left[\begin{matrix}-\mathbb{M}&\emptyset\\ \emptyset&\mathbb{H}\end{matrix}\right],~{}~{}\vec{b}(t)\equiv\binom{\dot{\vec{a}}(t)}{\vec{a}(t)}, (39)

we can rewrite Eq.(38) as

𝔸b˙(t)+𝔹b(t)=0,\mathbb{A}\dot{\vec{b}}(t)+\mathbb{B}\vec{b}(t)=0, (40)

where 𝔸\mathbb{A} and 𝔹\mathbb{B} are the square matrices of order 2N2N. By setting the temporal dependence of bn(t)b_{n}(t) as exp(σt)\exp(\sigma t), Eq. (40) can be cast in the form of a generalized eigenvalue problem, namely,

σ𝔸b+𝔹b=0.\sigma\mathbb{A}\vec{b}+\mathbb{B}\vec{b}=0. (41)

By solving Eq. (41) we obtain a set of 2N2N eigenvalues, σ\sigma, and the corresponding eigenvectors, b\vec{b}. The time-dependent coefficients, bn(t)b_{n}(t), can be expressed as a superposition of the eigenvectors as

bn(t)=n=12NCnβnneσnt,b_{n}(t)=\sum_{n^{\prime}=1}^{2N}C_{n^{\prime}}\beta_{nn^{\prime}}e^{\sigma_{n^{\prime}}t}, (42)

where, βnn\beta_{nn^{\prime}} is the nnth component of the nn^{\prime}th eigenvector and σn\sigma_{n^{\prime}} is the nn^{\prime}th eigenvalue. The constant coefficients CnC_{n^{\prime}} are obtained from the initial conditions. From the definition of b(t)\vec{b}(t) in Eq. (39) we can see that the desired coefficients an(t)a_{n}(t) correspond to the last NN components of b(t)\vec{b}(t), i.e.

an(t)=n=12NCnβN+n,neσnt,n=1,2,,N.a_{n}(t)=\sum_{n^{\prime}=1}^{2N}C_{n^{\prime}}\beta_{N+n,n^{\prime}}e^{\sigma_{n^{\prime}}t},~{}~{}~{}n=1,2,\ldots,N. (43)

Hence the expression for ξr(r,t)\xi_{r}(r,t) takes the following form

ξr(r,t)=n=1Nn=12NCnβN+n,neσntψn(r).\xi_{r}(r,t)=\sum_{n=1}^{N}\sum_{n^{\prime}=1}^{2N}C_{n^{\prime}}\beta_{N+n,n^{\prime}}e^{\sigma_{n^{\prime}}t}\psi_{n}(r). (44)

3.3 Initial Conditions

As in ST2015 we take the initial conditions for ξr\xi_{r} as

ξr(r,t=0)={ξ0,rr1,ξ0ψ1(r)ψ1(r1),r1<r<r2,ξ0ψ1(r2)ψ1(r1)(r2r)2,rr2,\displaystyle\xi_{r}(r,t=0)=\left\{\begin{array}[]{lll}\xi_{0},&r\leqslant r_{1},&\\ \xi_{0}\frac{\psi_{1}(r)}{\psi_{1}(r_{1})},&r_{1}<r<r_{2},&\\ \xi_{0}\frac{\psi_{1}(r_{2})}{\psi_{1}(r_{1})}\left(\frac{r_{2}}{r}\right)^{2},&r\geqslant r_{2},\end{array}\right. (48)
ξrt|(r,t=0)=0,\displaystyle\frac{\partial\xi_{r}}{\partial t}\Big{|}_{(r,t=0)}=0, (49)

where ξ0\xi_{0} is a constant. We choose these initial conditions in purpose in order to be sure that at time t=0t=0 the entire energy of the perturbations is in the generalized Fourier mode with largest spatial scale i.e. ψ1(r)\psi_{1}(r). This enables us to investigate the process of phase mixing of the perturbations in which the energy of the wave transfers from large spatial scales to smaller ones. In other words, with these initial conditions, as time goes on, the Fourier modes with smaller and smaller spatial scales contribute in the evolution of the kink wave. So, the larger the number of the available modes, NN, the larger the evolution time that we are allowed to proceed before the solutions become inaccurate (for more details see Cally 1991).

Here, we rewrite Eq. (42) in its matrix form, namely,

b(t)=β^eΣtC,\vec{b}(t)=\hat{\beta}e^{\Sigma t}\vec{C}, (50)

where Σ=diag(σ1,σ2,,σ2N)\Sigma={\rm diag}(\sigma_{1},\sigma_{2},\ldots,\sigma_{2N}), C=[C1,C2,,C2N]T\vec{C}=\left[C_{1},C_{2},\ldots,C_{2N}\right]^{T} and β^\hat{\beta} is a 2N2N by 2N2N matrix that its columns are the eigenvectors of Eq. (41). Evaluating Eq. (50) at t=0t=0 yields

b(t=0)=β^I^C=β^C,\vec{b}(t=0)=\hat{\beta}\hat{I}\vec{C}=\hat{\beta}\vec{C}, (51)

where I^\hat{I} is the identity matrix of size 2N×2N2N\times 2N. Assuming that the matrix β^\hat{\beta} has a non-zero determinant, we obtain the coefficient vector C\vec{C} as follows

C=β^1b(t=0)=β^1[a˙(t=0)a(t=0)].\vec{C}=\hat{\beta}^{-1}\vec{b}(t=0)=\hat{\beta}^{-1}\left[\begin{matrix}\dot{\vec{a}}(t=0)\\ \vec{a}(t=0)\end{matrix}\right]. (52)

Setting t=0t=0 in Eq. (22) and using the initial conditions presented in Eqs. (48) and (49) one can easily find that the only non-zero component of b(t=0)\vec{b}(t=0) is bN+1(t=0)=a1(t=0)=ξ0/ψ1(r1)b_{N+1}(t=0)=a_{1}(t=0)=\xi_{0}/\psi_{1}(r_{1}). Hence the coefficients CnC_{n} are obtained as

Cn=βn,N+11ξ0ψ1(r1).C_{n}=\beta^{-1}_{n,N+1}\frac{\xi_{0}}{\psi_{1}(r_{1})}. (53)

4 Numerical results

In order to solve Eq. (41), the density ratio of the flux tube is considered to be ρi/ρe=5\rho_{\rm i}/\rho_{\rm e}=5. The observational values of this parameter has been reported to be in the range [2, 10] (Aschwanden et al. 2003). The azimuthal mode number representing the kink waves is m=±1m=\pm 1. For the model considered in this paper, the results for both m=+1m=+1 and m=1m=-1 are the same. Here, we take m=+1m=+1. Since we use the TT approximation, the longitudinal wavenumber is assumed to be kz=π100Rk_{z}=\frac{\pi}{100R}. For the thickness of the inhomogeneous layer we consider two cases l/R=0.2l/R=0.2 and l/R=1l/R=1 that represent a thin and a thick layer, respectively. To consider the effect of viscosity, it is appropriate to use a dimensionless quantity, the viscous Reynolds number which is defined as

RvlvAiν.R_{v}\equiv\frac{lv_{\rm Ai}}{\nu}. (54)

Here, the Alfvén speed vAi=B0/μ0ρiv_{\rm Ai}=B_{0}/\sqrt{\mu_{0}\rho_{\rm i}} is the characteristic speed for the propagation of kink waves in the flux tube. Traditional value of the coefficient of shear viscosity in the solar corona is of the order ν=1m2/s\nu=1~{}m^{2}/s (see Ofman et al. 1994 and references therein). With l=106ml=10^{6}m and vAi=106m/sv_{\rm Ai}=10^{6}m/s, the corresponding Reynolds number is Rv=1012R_{v}=10^{12}. Due to computational limitations, we are forced to consider smaller RvR_{v} in our results. However, the conclusions we obtain can be easily generalized to the case of larger RvR_{v}. When appropriate, we shall stress what differences would appear if more realistic values of RvR_{v} were considered.

Solving the generalized eigenvalue problem (41) results to 2N eigenvalues, σiω~\sigma\equiv-i\tilde{\omega} where ω~ω+iγ\tilde{\omega}\equiv\omega+i\gamma is the complex eigenfrequency. These eigenvalues are real or come in pairs (σ,σ)(\sigma,\sigma^{*}) where σ\sigma^{*} is the complex conjugate of σ\sigma. Figure 1 shows the ωn0\omega_{n}\geqslant 0 part of the spectrum of the complex eigenvalues σniωn+γn\sigma_{n}\equiv-i\omega_{n}+\gamma_{n} with N=101N=101 for l/R=0.2l/R=0.2 and l/R=1l/R=1 and Reynolds numbers Rv=106,107R_{v}=10^{6},10^{7}. Here, ωn\omega_{n} and γn\gamma_{n} are the frequency and the damping rate of the nn’th eigenmode, respectively. In the figure, ωAi=kzB0/μ0ρi\omega_{\rm Ai}=k_{z}B_{0}/\sqrt{\mu_{0}\rho_{\rm i}} and ωAe=kzB0/μ0ρe\omega_{\rm Ae}=k_{z}B_{0}/\sqrt{\mu_{0}\rho_{\rm e}} are the Alfvén frequencies inside and outside of the tube, respectively. Note that the frequencies and damping rates are in units of the internal Alfvén frequency, ωAi\omega_{\rm Ai}. The complex spectrum has the typical three-branch structure found in previous papers that computed the resistive Alfvén spectrum in similar configurations (see Poedts & Kerner 1991). It is clear from Figure 1 that the smaller the value of the Reynolds number (larger coefficient of viscosity) the more number of modes with ω=0\omega=0 are present in the complex spectrum. This result is in well agreement with previous results obtained in resistive MHD (see e.g. Van Doorsselaere and Poedts 2007). An interesting result is that similar to the resistive MHD analysis (see e.g. Poedts and Kerner 1991) in viscous MHD, one of the eigenvalues of the complex spectrum could be identified as the damped quasi-mode solution (global mode) of ideal MHD. The real part of the frequency of this solution is the kink mode frequency, and the imaginary part is its corresponding damping rate due to the resonant absorption mechanism. Following Ruderman and Roberts (2002) and Goossens et al. (2002), the quasi-mode frequency ωqm\omega_{\rm qm} and damping rate γqm\gamma_{\rm qm} for MHD kink waves in the TT and thin boundary (TB) (l/R1l/R\ll 1) approximations are

ωqmωk=2ρiρi+ρeωAi,\omega_{\rm qm}\simeq\omega_{\rm k}=\sqrt{\frac{2\rho_{\rm i}}{\rho_{\rm i}+\rho_{\rm e}}}\omega_{\rm Ai}, (55)
γqm=ωkl4Rρiρeρi+ρe.\gamma_{\rm qm}=-\omega_{\rm k}\frac{l}{4R}\frac{\rho_{\rm i}-\rho_{\rm e}}{\rho_{\rm i}+\rho_{\rm e}}. (56)

From these equations for ρi/ρe=5\rho_{\rm i}/\rho_{\rm e}=5 the quasi-mode frequency is obtained as ωqm=1.29ωAi\omega_{qm}=1.29\omega_{\rm Ai}. Also the quasi-mode damping rates for l/R=0.2l/R=0.2 and l/R=1l/R=1 are γqm=0.043ωAi\gamma_{\rm qm}=-0.043\omega_{\rm Ai} and γqm=0.215ωAi\gamma_{\rm qm}=-0.215\omega_{\rm Ai}, respectively. As illustrated in Figure 1, the singled out mode matches the quasi-mode solution especially for the thin transitional layer case, l/R=0.2l/R=0.2, and larger values of the Reynolds number. For instance, for l/R=0.2l/R=0.2 and Rv=107R_{v}=10^{7} the singled out eigenfrequency is ω~=1.290.044i\tilde{\omega}=1.29-0.044i that is in well agreement with the result obtained from the above analytic approximations. For l/R=1l/R=1 and Rv=107R_{v}=10^{7} the corresponding eigenfrequency is ω~=1.330.269i\tilde{\omega}=1.33-0.269i. Hence, in the case of thick transitional layer, the numerical results and the analytic approximations deviate more from each other. This is consistent with the fact that the analytic approximations are only strictly valid when l/R1l/R\ll 1. For a study of the validity of the approximations beyond its theoretical range of applicability see Soler et al. (2014). These results show that for the kink waves, the quasi-mode solution of the ideal MHD can be identified as an eigenmode of the viscous MHD spectrum. This correspondence is very clear in the case of thin nonuniform layers and not very large Reynolds numbers. However, as the thickness of the layer or the Reynolds numbers increase, the identification of the quasi-mode in the dissipative spectrum is more confusing since the quasi-mode gets embedded in one of the branches of the spectrum and becomes indistinguishable from an ordinary dissipative Alfvén mode (see discussions on this issue in Van Doorsselaere & Poedts 2007 and Soler et al. 2013). A detailed analysis of the peculiar behaviour of the quasi-mode in the complex spectrum is beyond the purpose of the present paper.

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 1: Spectrum of the eigenvalues for kz=π/100k_{z}=\pi/100 and ρi/ρe=5\rho_{\rm i}/\rho_{\rm e}=5. Left and right panels are for l/R=0.2l/R=0.2 and l/R=1l/R=1, respectively. Top and bottom panels are for Rv=106R_{v}=10^{6} and Rv=107R_{v}=10^{7}, respectively.

Once the dissipative spectrum is computed, the time-dependent behavior of the perturbations is obtained by the superposition of all the modes in the spectrum according to the prescribed initial condition, which represents a transverse, i.e., kink displacement of the whole tube (see Section 3.3). In short, the evolution of the subsequent global kink oscillation is determined by two simultaneously working mechanisms: resonant absorption and phase mixing. On the one hand, resonant absorption is responsible for a radial flux of energy towards the nonuniform layer of the flux tube, and its net effect is producing the damping of the global kink oscillation. As a result, the amplitude of the displacement at the tube axis decreases in time. On the other hand, the energy accumulated at the nonuniform layer because of resonant absorption drives local Alfvén waves with the same azimuthal symmetry as the original kink oscillation, i.e., m=1m=1. Although these Alfvén waves have both radial and azimuthal components of the displacement, they are largely polarized in the azimuthal direction. The Alfvén waves undergo the process of phase mixing because of the spatially-dependent Alfvén velocity. This causes the building up of small scales in the nonuniform layer and the subsequent energy cascade to these small scales. Then, viscous dissipation becomes important. In the following paragraphs, we analyze the dynamics we have just summarized.

Figures 2 and 3 show the evolution of ξr\xi_{r} and ξφ\xi_{\varphi}, respectively. Time is in units of the period of the kink oscillation in TTTB approximations, Pk=2π/ωkP_{\rm k}=2\pi/\omega_{\rm k}. In Figures 2 and 3 the solid black curves represent the results previously obtained by ST2015 in the absence of viscosity, i.e. Rv=R_{v}=\infty. The blue dashed and red dashed-dotted curves are for Rv=107R_{v}=10^{7} and Rv=106R_{v}=10^{6}, respectively. Figures reveal that in the presence of viscosity, unlike the results obtained in ideal MHD (see Soler and Terradas 2015; Ebrahimi et al. 2017) perturbations are not allowed to be phase mixed indefinitely. The smaller the Reynolds number, the quicker the dissipative solution departs from the ideal solution. The existence of viscosity cause to coupling of the perturbations on the neighboring magnetic surfaces. This effect suppresses the phase mixing when a certain spatial scale is reached and transforms the total (kinetic plus magnetic) energy of the perturbations to heat via dissipation.

Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption
Figure 2: Evolution of ξr\xi_{r} in a nonuniform tube with l/R=0.2l/R=0.2 (top panels) and l/R=1l/R=1 (bottom panels) for Rv=,106R_{v}=\infty,10^{6} and 10710^{7}. Left, center and right panels denote t/Pk=0t/P_{k}=0, 33 and 1010, respectively. The left and right vertical dashed lines locate r=r1r=r_{1} and r=r2r=r_{2}, respectively. Other auxiliary parameters are as in Figure 1. An animation of this Figure is available.
Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption
Figure 3: Same as Figure 2 but for ξφ\xi_{\varphi}. An animation of this Figure is available.

To illustrate the flux of the total energy of the kink wave from the internal and external regions to the inhomogeneous region where it is finally dissipated, we calculate the total energy density of the perturbations as

E(r,t)=12(ρ|𝝃t|2+1μ|𝐁|2),E(r,t)=\frac{1}{2}\left(\rho\left|\frac{\partial\boldsymbol{\xi}}{\partial t}\right|^{2}+\frac{1}{\mu}\left|\mathbf{B^{\prime}}\right|^{2}\right), (57)

where 𝐁\mathbf{B^{\prime}} is obtained from Eq. (5). Figure 4 illustrates the evolution of the total energy density as a function of rr. In the absence of viscosity after a few time periods the whole energy of the kink wave concentrates in a narrow layer in the inhomogeneous region of the flux tube. The role of viscosity is to dissipate this concentrated energy. Although, these two processes i.e. flow of energy to the inhomogeneous layer and the dissipation occur simultaneously but they are caused by independent and physically different mechanisms. In order to have a better illustration of the flow of energy of the kink wave from internal and external regions to the inhomogeneous layer and its dissipation, it is appropriate to calculate the integrated total energy of the perturbations in the interior, transitional layer and exterior of the flux tube as a function of time, respectively, as follows

Ein=0r112(ρ|𝝃t|2+1μ|𝐁|2)rdr,\displaystyle E_{\rm{in}}=\int_{0}^{r_{1}}\frac{1}{2}\left(\rho\left|\frac{\partial\boldsymbol{\xi}}{\partial t}\right|^{2}+\frac{1}{\mu}\left|\mathbf{B^{\prime}}\right|^{2}\right)r~{}{\rm d}r,
Etr=r1r212(ρ|𝝃t|2+1μ|𝐁|2)rdr,\displaystyle E_{\rm{tr}}=\int_{r_{1}}^{r_{2}}\frac{1}{2}\left(\rho\left|\frac{\partial\boldsymbol{\xi}}{\partial t}\right|^{2}+\frac{1}{\mu}\left|\mathbf{B^{\prime}}\right|^{2}\right)r~{}{\rm d}r, (58)
Eex=r212(ρ|𝝃t|2+1μ|𝐁|2)rdr.\displaystyle E_{\rm{ex}}=\int_{r_{2}}^{\infty}\frac{1}{2}\left(\rho\left|\frac{\partial\boldsymbol{\xi}}{\partial t}\right|^{2}+\frac{1}{\mu}\left|\mathbf{B^{\prime}}\right|^{2}\right)r~{}{\rm d}r.

Note that in order to compute the third integral of Eq. (4) we must replace the upper limit of the integral, \infty, with a sufficiently large radius where the amplitudes of the perturbations are already negligible. Hence, we check the convergence of the total energy by taking a series of large radii as the upper limit of the third integral in Eq. (4). Results show that at r=20Rr=20R the total energy converges to the desired level of accuracy. Figure 5 shows the integrated energy in these three regions as a function of time for Rv=,107&106R_{v}=\infty,~{}10^{7}~{}\&~{}10^{6}. The solid, dashed and dotted lines denote the integrated energies of the internal, inhomogeneous and external regions, respectively. The black, blue and red colors represent Rv=R_{v}=\infty (ideal MHD), Rv=107R_{v}=10^{7} and Rv=106R_{v}=10^{6}, respectively. Note that in the figure, the plots of the internal and external energies (i.e. the solid and dotted lines) for Rv=R_{v}=\infty and Rv=107&106R_{v}=10^{7}\&10^{6} coincide with each other which reveals that for large Reynolds numbers the existence of viscosity does not affect the energy flow from internal and external regions to the inhomogeneous layer. In other words, the resonant absorption process is not affected by viscosity. Figure 5 also shows that in the presence of viscosity, the energy in the inhomogeneous layer increases in the initial stage of the evolution and reaches a maximum after a few time periods. After that, the energy decreases, since phase mixing in the inhomogeneous layer enhances the viscous dissipation mechanism. The temporal behaviour of the total energy of the kink wave has been illustrated in Figure 6 for l/R=0.2&1l/R=0.2~{}\&~{}1 and Rv=,107&106R_{v}=\infty,~{}10^{7}~{}\&~{}10^{6}. As figure shows, the energy of the kink wave in the absence of viscosity (black line in the figure) is conserved. The results show that, considering a flux tube with thin transitional layer (l/R=0.2l/R=0.2), for Rv=107&106R_{v}=10^{7}~{}\&~{}10^{6} the energy of the kink wave decreases to 1/e1/e of its initial energy after t5.8Pkt\simeq 5.8P_{k} and t4.4Pkt\simeq 4.4P_{k}, respectively. For thick transitional layer (l/R=1l/R=1) the corresponding values are t5.2Pkt\simeq 5.2P_{k} and t2.7Pkt\simeq 2.7P_{k}.

Figure 7 shows the radial component of the displacement at the axis of the flux tube. Interestingly the results for Rv=,107&106R_{v}=\infty,10^{7}\&10^{6} are the same. Note that in the observations of kink waves in coronal flux tubes, we measure the displacement of the axis but detecting the rotational motions related to the kink waves in the inhomogeneous boundary of the flux tubes is not an easy task. Comparing the results illustrated in Figures 6 and 7, it is clear that although the temporal behaviour of the displacement of the flux tube axis is the same in both cases of ideal and viscous MHD (with large Reynolds number), the total energy of the kink waves which is conserved for Rv=R_{v}=\infty, decays in a few periods for Rv=107R_{v}=10^{7} and 10610^{6}. Hence, one can conclude that assuming a large Reynolds number for the coronal plasma the global damping of kink waves reported in the observations is due to converting the transverse motion to rotational perturbations in the inhomogeneous layer of the flux tube, i.e. the resonant absorption process, and is not related to the existence of viscosity.

In order to illustrate the rate of plasma heating in the inhomogeneous region due to viscosity we use the viscous dissipation function that for an incompressible plasma with constant viscosity is as follows (White 1991)

Φ=ρν[2(εrr2+εφφ2+εzz2)+εrφ2+εrz2+εφz2],\Phi=\rho\nu\left[2\left(\varepsilon_{rr}^{2}+\varepsilon_{\varphi\varphi}^{2}+\varepsilon_{zz}^{2}\right)+\varepsilon_{r\varphi}^{2}+\varepsilon_{rz}^{2}+\varepsilon_{\varphi z}^{2}\right], (59)

where εij\varepsilon_{ij} with i,j=r,φ,zi,j=r,\varphi,z are components of the strain rate tensor defined as follows

εrr\displaystyle\varepsilon_{rr} =\displaystyle= 2ξrrt,\displaystyle\frac{\partial^{2}\xi_{r}}{\partial r\partial t},
εφφ\displaystyle\varepsilon_{\varphi\varphi} =\displaystyle= 1r2ξφφt+1rξrt,\displaystyle\frac{1}{r}\frac{\partial^{2}\xi_{\varphi}}{\partial\varphi\partial t}+\frac{1}{r}\frac{\partial\xi_{r}}{\partial t},
εzz\displaystyle\varepsilon_{zz} =\displaystyle= 2ξzzt,\displaystyle\frac{\partial^{2}\xi_{z}}{\partial z\partial t},
εrφ\displaystyle\varepsilon_{r\varphi} =\displaystyle= 1r2ξrφt+2ξφrt1rξφt,\displaystyle\frac{1}{r}\frac{\partial^{2}\xi_{r}}{\partial\varphi\partial t}+\frac{\partial^{2}\xi_{\varphi}}{\partial r\partial t}-\frac{1}{r}\frac{\partial\xi_{\varphi}}{\partial t},
εrz\displaystyle\varepsilon_{rz} =\displaystyle= 2ξrzt+2ξzrt,\displaystyle\frac{\partial^{2}\xi_{r}}{\partial z\partial t}+\frac{\partial^{2}\xi_{z}}{\partial r\partial t},
εφz\displaystyle\varepsilon_{\varphi z} =\displaystyle= 1r2ξzφt+2ξφzt.\displaystyle\frac{1}{r}\frac{\partial^{2}\xi_{z}}{\partial\varphi\partial t}+\frac{\partial^{2}\xi_{\varphi}}{\partial z\partial t}.

The dissipation function, Φ\Phi, is the work done by the viscous stresses on an element of plasma per unit volume per unit time. Here, we remove the φ\varphi and zz dependency of the dissipation function by integrating Φ\Phi in these directions, i.e.

Φ=0λ02πΦr𝑑φ𝑑z=πλrΦ,\Phi^{\prime}=\int_{0}^{\lambda}\int_{0}^{2\pi}\Phi rd\varphi dz=\pi\lambda r\Phi, (60)

where, λ=2π/kz\lambda=2\pi/k_{z} is the wavelength. Figure 8 shows the contour plot of Φ\Phi^{\prime} in rtr-t plane for l/R=0.2l/R=0.2 and l/R=1l/R=1 with Rv=106R_{v}=10^{6} and Rv=107R_{v}=10^{7}. As figure shows, the heating rate has an oblique oscillatory pattern. The obliquity of the contours of Φ\Phi^{\prime} is due to the phase mixing of the perturbations. In order to obtain the time that the dissipation in the inhomogeneous region reaches its maximum we have calculated the integral of Φ\Phi^{\prime} over the range [r1,r2][r_{1},r_{2}] that is a function of time. For Rv=106R_{v}=10^{6}, considering l/R=0.2l/R=0.2 and l/R=1l/R=1 we obtain the peak time of the integrated dissipation as tpeak=2.75Pkt_{peak}=2.75P_{k} and tpeak=2.1Pkt_{peak}=2.1P_{k}, respectively. The corresponding values for Rv=107R_{v}=10^{7} are tpeak=4.5Pkt_{peak}=4.5P_{k} and tpeak=4.1Pkt_{peak}=4.1P_{k}. For l/R=0.2l/R=0.2 and Rv=108R_{v}=10^{8} we get tpeak=8Pkt_{peak}=8P_{k}. So, for the larger value of the Reynolds number the system needs to be more phase mixed for viscosity to reach its maximum efficiency as a heating mechanism for the plasma. Obviously, if the Reynolds number is further increased to the expected value in the solar corona, the peak efficiency of dissipation happens in a much later time corresponding to many periods of the original kink oscillation. Since the observationally reported damping time of kink oscillations corresponds to a few periods (as the theory of resonant absorption correctly predicts), we conclude that viscous heating becomes of significance only after the global kink oscillation is damped, i.e., no heating is expected during the damping of the global kink oscillation in realistic coronal conditions.

Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption
Figure 4: Energy density as a function of rational radius in unit of the total initial energy with l/R=0.2l/R=0.2 (top panels) and l/R=1l/R=1 (bottom panels) for Rv=,106R_{v}=\infty,10^{6} and 10710^{7}. Left, center and right panels denote t/Pk=0t/P_{k}=0, 33 and 1010, respectively. The left and right vertical dashed lines locate r=r1r=r_{1} and r=r2r=r_{2}, respectively. Other auxiliary parameters are as in Figure 1.

5 Summary and Conclusions

Here, we investigated the effect of viscosity on the evolution of MHD kink waves in coronal flux tubes. We modelled a magnetic flux tube by a straight magnetic cylinder. Plasma density inside and outside the tube is constant with different values. The interior and exterior of the tube are connected by an inhomogeneous transitional layer in which the plasma density varies smoothly with a sinusoidal profile from the internal value to the external one. The background magnetic field is aligned with tube axis and has constant magnitude everywhere. We neglected the role of viscosity in the constant density regions since in the limit of small viscosities the dissipation is only important in the inhomogeneous region where the phase mixing process is at work. Using the modal expansion technique (Cally 1991; Soler & Terradas 2015) we solved the viscous MHD equations of motion in thin tube approximation and obtained the spatio-temporal behaviour of the perturbations in the flux tube. We considered both the cases of thin and thick inhomogeneous layers in our analysis.

We obtained the spectrum of the complex eigenfrequencies of the Alfvén discrete modes in the inhomogeneous layer. In the spectrum, one of the eigenfrequencies could be identified as the quasi-mode solution of the kink waves by the resonant absorption mechanism.

Refer to caption
Refer to caption
Figure 5: Integrated energy of the kink wave in the interior, inhomogeneous region and exterior of the flux tube in unit of the total initial energy versus time for Rv=,106R_{v}=\infty,10^{6} and 10710^{7}. Top and bottom panel are for l/R=0.2l/R=0.2 and l/R=1l/R=1, respectively. Other auxiliary parameters are as in Figure 1.
Refer to caption
Refer to caption
Figure 6: Total energy of the kink wave versus time for Rv=,106&107R_{v}=\infty,10^{6}\&10^{7}. Top and bottom panel are for l/R=0.2l/R=0.2 and l/R=1l/R=1, respectively. Other auxiliary parameters are as in Figure 1.
Refer to caption
Refer to caption
Figure 7: Temporal behavior of ξr\xi_{r} on the axis of the flux tube for Rv=,106R_{v}=\infty,10^{6} and 10710^{7}. Top and bottom panel are for l/R=0.2l/R=0.2 and l/R=1l/R=1, respectively. Other auxiliary parameters are as in Figure 1.
Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 8: Contours of the dissipation function in normalized units in rtr-t plane for Rv=106R_{v}=10^{6} (top panels) and 10710^{7} (bottom panels). Left and right panels are for l/R=0.2l/R=0.2 and l/R=1l/R=1, respectively. Other auxiliary parameters are as in Figure 1.

To investigate the effect of viscosity on the kink waves, we obtained the temporal and spatial behaviour of the perturbations in the interior, inhomogeneous region and exterior of the flux tube. Our results showed that for both cases of thin and thick inhomogeneous layers, considering the viscosity in the system does not affect the transverse motion of the flux tube axis and its decay rate for large Reynolds numbers (Rv106R_{v}\geq 10^{6}) confirming the previously results obtained by e.g. Ruderman & Roberts (2002) and Goossens et al. (2002). This result confirms that the fast damping of the kink oscillations in coronal loops could be a consequence of resonant absorption mechanism which despite the existence of viscosity naturally results to changing the behaviour of kink mode from mainly transverse motion to rotational motion (Goossens et al. 2014). However, even for small viscosities (relevant for the coronal plasma) the viscous dissipation is important in the developed stage of phase mixing of the perturbations in the inhomogeneous region. Our results show that viscosity eventually suppresses the rate of phase mixing of the perturbations by coupling the neighboring magnetic surfaces and transforming their energy to heat.

In order to investigate the effect of viscosity on cascading the total (kinetic plus magnetic) energy of the kink wave to the inhomogeneous layer of the flux tube, we obtained the total energy density as well as the integrated total energy as a function of rr and tt. As in the ideal MHD case, in presence of viscosity the energy of the kink wave tends to concentrate in a narrow region in the inhomogeneous layer of the flux tube but it is not allowed to achieve its maximum value obtained in the ideal case since the dissipation mechanism is at work and decreases the energy in time. However, the small amount of viscosity considered in our analysis does not affect the energy flow from interior and exterior of the tube to the inhomogeneous region. Temporal behavior of the total energy showed that for both the thin and thick inhomogeneous region with Reynolds numbers of the order 10610^{6}, 10710^{7} and 10810^{8} the energy of the kink wave decays to heat the plasma within a few periods. However, if larger and more realistic values of the Reynolds numbers were used, heating would happen much after the observable kink oscillation is completely damped. These conclusions agree with the simple estimations provided in the research note by Terradas & Arregui (2018), although these authors considered resistive heating instead of viscous heating.

We also studied the efficiency of heating duo to viscosity by calculating the dissipation function in the inhomogeneous region. The obtained results show that at initial stage of the evolution the dissipation increases with time and reaches a maximum level after a few periods (2-8 periods for Rv=106108R_{v}=10^{6}-10^{8} but much later for realistic RvR_{v}) in a narrow layer near the boundary of the flux tube. After that the dissipation decreases since the energy budget provided by the initial value problem considered in this paper is finite.

In summary, we showed that viscosity, even in a small amount, can have a significant impact on the later evolution of phase mixing by the suppressing the generation of small scales and transforming the energy of the wave to heat. Reynolds number larger than the values considered in this paper needs more massive numerical computation with the mathematical approach presented here to obtain the correct set of complex eigenvalues of the damped Alfvén discrete modes which could be a subject of future work.

We finally note that this work is based on linear theory. Nonlinear effects may modify somehow these results, since presumably important ingredients as, e.g., the triggering of Kelvin-Helmholtz instabilities are absent from our study (see e.g. Terradas et al. 2008). The nonlinear evolution should be necessarily investigated with high-resolution dissipative MHD simulations.

Acknowledgements

RS acknowledges the support from grant AYA2017- 85465-P (MINECO/AEI/FEDER, UE) and from the Ministerio de Economía, Industria y Competitividad and the Conselleria d’Innovació, Recerca i Turisme del Govern Balear (Pla de ciència, tecnologia, innovació i emprenedoria 2013−2017) for the Ramón y Cajal grant RYC-2014-14970.

References

  • [1] Andries, J., Goossens, M., Hollweg, J. V., Arregui, I., & Van Doorsselaere, T. 2005, A&A, 430, 1109
  • [2] Aschwanden, M. J., Fletcher, L., Schrijver, C. J., & Alexander D. 1999, ApJ, 520, 880
  • [3] Aschwanden, M. J., Nightingale, R. W., Andries, J., Goossens, M., & Van Doorsselaere, T. 2003, ApJ, 598, 1375
  • [4] Cally, P. S. 1991, JPlPh, 45, 453
  • [5] Cally, P. S., & Andries, J. 2010, Sol.Phys., 266, 17
  • [6] Davila, J. M. 1987, ApJ, 317, 514
  • [7] Ebrahimi, Z., & Bahari, K. 2019, MNRAS, 490, 1644
  • [8] Ebrahimi, Z., & Karami, K. 2016, MNRAS, 462, 1002
  • [9] Ebrahimi, Z., Karami, K., & Soler, R. 2017, ApJ, 845, 86
  • [10] Erdélyi, R. 1997, Sol.Phys., 171, 49
  • [11] Erdélyi, R., Goossens, M. 1995, A&A, 294, 575
  • [12] Erdélyi, R., Goossens, M. 1996, A&A, 313, 664
  • [13] Goossens, M., & Ruderman, M. S. 1995, PhSc, 157, 75
  • [14] Goossens, M., Ruderman, M. S., & Hollweg, J. V. 1995, Sol.Phys., 157, 75
  • [15] Goossens, M., Andries, J., Aschwanden, M. J. 2002, A&A, 394, L39
  • [16] Goossens, M., Terradas, J., Andries, J., Arregui, I., & Ballester, J. L. 2009, A&A, 503, 213
  • [17] Goossens, M., Erdélyi, R., & Ruderman, M. S. 2011, Space Sci. Rev., 158, 289
  • [18] Goossens, M., Soler, R., Terradas, J., Van Doorsselaere, T., & Verth, G. 2014, ApJ, 788, 9
  • [19] Heyvaerts, J., & Priest, E. R., 1983, A&A, 117, 220
  • [20] Howson, T. A., De Moortel, I., Antolin, P., Van Doorsselaere, T., & Wright A. N. 2019, A&A, 631, A105
  • [21] Ionson, J. A. 1978, ApJ, 226, 650
  • [22] Ireland, J., & Priest, E. 1997, Sol.Phys., 173, 31
  • [23] Karami, K., & Asvar, A. 2007, MNRAS, 381, 97
  • [24] Karami, K., & Bahari, K. 2010, Sol.Phys., 263, 87
  • [25] Karami, K., & Ebrahimi, Z. 2009, PASA, 26, 448
  • [26] Morton, R. J., & Erdélyi, R. 2009, ApJ, 707, 750
  • [27] Morton, R. J., & Erdélyi, R. 2010, A&A, 519, A43
  • [28] Nakariakov, V. M., Ofman, L., DeLuca, E. E., Roberts, B., & Davila, J. M. 1999, Sci., 285, 862
  • [29] Ofman, L. 2005, Adv. Space Res., 36, 1572
  • [30] Ofman, L. 2009, ApJ, 694, 502
  • [31] Ofman, L., Aschwanden, M. 2002, ApJL, 576, L153
  • [32] Ofman, L., Davila, J. M., & Steinolfson, R. S. 1994, ApJ, 421, 360
  • [33] Poedts, S., Kerner, W. 1991, PRL, 66, 2871
  • [34] Prokopyszyn, A. P. K., & Hood, A. W. 2019, A&A, 632, A93
  • [35] Ruderman, M. S., & Roberts, B. 2002, ApJ, 577, 475
  • [36] Sakurai, T., Goossens, M., & Hollweg J. V. 1991a, Sol.Phys., 133, 227
  • [37] Sakurai, T., Goossens, M., & Hollweg J. V. 1991b, Sol.Phys., 133, 247
  • [38] Shukhobodskiy, A. A., Ruderman, M. S., & Erdélyi, R. 2018, A&A, 619, A173
  • [39] Soler, R., Goossens, M., Terradas, J., & Oliver, R. 2013, ApJ, 777, 158
  • [40] Soler, R., & Terradas, J. 2015, ApJ, 803, 43
  • [41] Soler, R., Goossens, M., Terradas, J., & Oliver, R. 2014, ApJ, 781, 111
  • [42] Soler, R., Terradas, J., Verth, G., & Goossens, M. 2011, ApJ, 736, 10
  • [43] Tirry, W. J., & Goossens, M. 1996, ApJ, 471, 501
  • [44] Terradas, J., Andries, J., Goossens, M., Arregui, I., Oliver, R., & Ballester, J. L. 2008, ApJL, 687, L115
  • [45] Terradas, J., & Arregui, I. 2018, Res. Notes AAS, 2, 196
  • [46] Terradas, J., Oliver, R., & Ballester, J. L. 2006, ApJL, 650, L91
  • [47] Van Doorsselaere, T., & Poedts, S. 2007, PPCF, 49, 261
  • [48] White, F.M., 1991, Viscous fluid flow (2nd ed), McGraw-Hill, Inc.