Skip to main content
Log in

Can High-Mode Magnetohydrodynamic Waves Propagating in a Spinning Macrospicule Be Unstable Due to the Kelvin–Helmholtz Instability?

  • Published:
Solar Physics Aims and scope Submit manuscript

Abstract

We investigate the conditions at which high-mode magnetohydrodynamic (MHD) waves propagating in a spinning solar macrospicule can become unstable with respect to the Kelvin–Helmholtz instability (KHI). We consider the macrospicule as a weakly twisted cylindrical magnetic flux tube moving along and rotating around its axis. Our study is based on the dispersion relation (in complex variables) of MHD waves obtained from the linearized MHD equations of an incompressible plasma for the macrospicule and cool (\(\beta = 0\), with \(\beta\) the plasma to the magnetic pressure) plasma for its environment. This dispersion equation is solved numerically using appropriate input parameters to find out an instability region or window that accommodates suitable unstable wavelengths on the order of the macrospicule width. It is established that an \(m = 52\) MHD mode propagating in a macrospicule with width of 6 Mm, axial velocity of \(75~\mbox{km}\,\mbox{s}^{-1}\), and rotating one of \(40~\mbox{km}\,\mbox{s}^{-1}\) can become unstable against the KHI with growth times of 2.2 and 0.57 min at 3 and 5 Mm unstable wavelengths, respectively. These growth times are much shorter than the macrospicule lifetime, which lasts about 15 min. An increase or decease in the width of the jet would change the KHI growth times, which remain more or less of the same order when they are evaluated at wavelengths equal to the width or radius of the macrospicule. It is worth noting that the excited MHD modes are super-Alfvénic. A change in the background magnetic field can lead to another MHD mode number, \(m\), that ensures the required instability window.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Figure 1
Figure 2
Figure 3
Figure 4

Similar content being viewed by others

References

  • Ajabshirizadeh, A., Ebadi, H., Vekalati, R.E., Molaverdikhani, K.: 2015, The possibility of Kelvin–Helmholtz instability in solar spicules. Astrophys. Space Sci. 367, 33. DOI .

    Article  ADS  Google Scholar 

  • Banerjee, D., O’Shea, E., Doyle, J.G.: 2000, Giant macro-spicule as observed by CDS on SOHO. Astron. Astrophys. 355, 1152. ADS .

    ADS  Google Scholar 

  • Bennett, S.M., Erdélyi, R.: 2015, On the statistics of macrospicules. Astrophys. J. 808, 135. DOI .

    Article  ADS  Google Scholar 

  • Bennett, K., Roberts, B., Narain, U.: 1999, Waves in twisted magnetic flux tubes. Solar Phys. 185, 41. DOI .

    Article  ADS  Google Scholar 

  • Bohlin, J.D., Vogel, S.N., Purcell, J.D., Sheeley, N.R. Jr., Tousey, R., VanHoosier, M.E.: 1975, A newly observed solar feature: macrospicules in He ii 304 Å. Astrophys. J. 197, L133. DOI .

    Article  ADS  Google Scholar 

  • Chandrasekhar, S.: 1961, Hydrodynamic and Hydromagnetic Stability, Clarendon Press, Oxford. Chap. 11.

    MATH  Google Scholar 

  • Culhane, J.L., Harra, L.K., James, A.M., et al.: 2007, The EUV Imaging Spectrometer for Hinode. Solar Phys. 243, 19. DOI .

    Article  ADS  Google Scholar 

  • De Pontieu, B., McIntosh, S., Hansteen, V.H., et al.: 2007, A tale of two spicules: the impact of spicules on the magnetic chromosphere. Proc. Astron. Soc. Japan 59, S655. DOI .

    Article  Google Scholar 

  • De Pontieu, B., Rouppe van der Voort, L., McIntosh, S.W., et al.: 2014, On the prevalence of small-scale twist in the solar chromosphere and transition region. Science 346, 1255732. DOI .

    Article  Google Scholar 

  • Domingo, V., Fleck, B., Poland, A.I.: 1995, SOHO: the Solar and Heliospheric Observatory. Space Sci. Rev. 72, 81. DOI .

    Article  ADS  Google Scholar 

  • Ebadi, H.: 2016, Kelvin–Helmholtz instability in solar spicules. Iranian J. Phys. Res. 16, 41. DOI .

    Article  Google Scholar 

  • Golub, L., Deluca, E., Austin, G., et al.: 2007, The X-Ray Telescope (XRT) for the Hinode Mission. Solar Phys. 243, 63. DOI .

    Article  ADS  Google Scholar 

  • Goossens, M., Hollweg, J., Sakurai, T.: 1992, Resonant behaviour of MHD waves on magnetic flux tubes. III. Effect of equilibrium flow. Solar Phys. 138, 233. DOI .

    Article  ADS  Google Scholar 

  • Harrison, R.A., Fludra, A., Pike, C.D., Payne, J., Thompson, W.T., Poland, A.I., Breeveld, E.R., Breeveld, A.A., Culhane, J.L., Kjeldseth-Moe, O., Huber, M.C.E., Aschenbach, B.: 1997, High-resolution observations of the extreme ultraviolet Sun. Solar Phys. 170, 123. DOI .

    Article  ADS  Google Scholar 

  • Howard, T.A., Moses, J.D., Vourlidas, A., et al.: 2008, Sun Earth Connection Coronal and Heliospheric Investigation (SECCHI). Space Sci. Rev. 136, 67. DOI .

    Article  ADS  Google Scholar 

  • Iijima, H., Yokoyama, T.: 2017, Three-dimensional magnetohydrodynamic simulation of the formation of solar chromospheric jets with twisted magnetic field lines. Astrophys. J. 848, 38. DOI .

    Article  ADS  Google Scholar 

  • Kaiser, M.L., Kucera, T.A., Davila, J.M., et al.: 2008, The STEREO Mission: an introduction. Space Sci. Rev. 136, 5. DOI .

    Article  ADS  Google Scholar 

  • Kamio, S., Curdt, W., Teriaca, L., Inhester, B., Solanki, S.K.: 2010, Observations of a rotating macrospicule associated with an X-ray jet. Astron. Astrophys. 510, L1. DOI .

    Article  ADS  Google Scholar 

  • Kayshap, P., Srivastava, A.K., Murawski, K., Tripathi, D.: 2013, Origin of macrospicule and jet in polar corona by a small-scale kinked flux tube. Astrophys. J. 770, L3. DOI .

    Article  ADS  Google Scholar 

  • Kiss, T.S., Gyenge, N., Erdélyi, R.: 2017, Systematic variations of macrospicule properties observed by SDO/AIA over half a decade. Astrophys. J. 835, 47. DOI .

    Article  ADS  Google Scholar 

  • Kosugi, T., Matsuzaki, K., Sakao, T., et al.: 2007, The Hinode (Solar-B) Mission: an overview. Solar Phys. 243, 3. DOI .

    Article  ADS  Google Scholar 

  • Kuridze, D., Zaqarashvili, T.V., Henriques, V., Mathioudakis, M., Keenan, F.P., Hanslmeier, A.: 2016, Kelvin–Helmholtz instability in solar chromospheric jets: theory and observation. Astrophys. J. 830, 133. DOI .

    Article  ADS  Google Scholar 

  • Lemen, J.R., Title, A.M., Akin, D.J., et al.: 2012, The Atmospheric Imaging Assembly (AIA) on the Solar Dynamics Observatory (SDO). Solar Phys. 275, 17. DOI .

    Article  ADS  Google Scholar 

  • Loboda, I.P., Bogachev, S.A.: 2017, Plasma dynamics in solar macrospicules from high-cadence extreme-UV observations. Astron. Astrophys. 597, A78. DOI .

    Article  ADS  Google Scholar 

  • Madjarska, M.S., Vanninathan, K., Doyle, J.G.: 2011, Can coronal hole spicules reach coronal temperatures? Astron. Astrophys. 532, L1. DOI .

    Article  ADS  Google Scholar 

  • Miura, A.: 1984, Anomalous transport by magnetohydrodynamic Kelvin–Helmholtz instabilities in the solar wind–magnetosphere interaction. J. Geophys. Res. 89, 801. DOI .

    Article  ADS  Google Scholar 

  • Murawski, K., Srivastava, A.K., Zaqarashvili, T.V.: 2011, Numerical simulations of solar macrospicules. Astron. Astrophys. 535, A58. DOI .

    Article  ADS  Google Scholar 

  • Parenti, S., Bromage, B.J.I., Bromage, G.E.: 2002, An erupting macrospicule: characteristics derived from SOHO-CDS spectroscopic observations. Astron. Astrophys. 384, 303. DOI .

    Article  ADS  Google Scholar 

  • Pesnell, W.D., Thompson, B.J., Chamberlin, P.C.: 2012, The Solar Dynamics Observatory (SDO). Solar Phys. 275, 3. DOI .

    Article  ADS  Google Scholar 

  • Pike, C.D., Harrison, R.A.: 1997, Euv observations of a macrospicule: evidence for solar wind acceleration? Solar Phys. 175, 457. DOI .

    Article  ADS  Google Scholar 

  • Pike, C.D., Mason, H.E.: 1998, Rotating transition region features observed with the SOHO Coronal Diagnostic Spectrometer. Solar Phys. 182, 333. DOI .

    Article  ADS  Google Scholar 

  • Ryu, D., Jones, T.W., Frank, A.: 2000, The magnetohydrodynamic Kelvin–Helmholtz instability: a three-dimensional study of nonlinear evolution. Astrophys. J. 545, 475. DOI .

    Article  ADS  Google Scholar 

  • Scullion, E., Doyle, J.G., Erdélyi, R.: 2010, A spectroscopic analysis of macrospicules. Mem. Soc. Astron. Ital. 81, 737. ADS .

    ADS  Google Scholar 

  • Vernazza, J.E., Avrett, E.H., Loeser, R.: 1981, Structure of the solar chromosphere. III – models of the EUV brightness components of the quiet-sun. Astrophys. J. Suppl. 45, 635. DOI .

    Article  ADS  Google Scholar 

  • Wilhelm, K., Curdt, W., Marsch, E., et al.: 1995, SUMER – Solar Ultraviolet Measurements of Emitted Radiation. Solar Phys. 162, 189. DOI .

    Article  ADS  Google Scholar 

  • Zank, G.P., Matthaeus, W.H.: 1993, Nearly incompressible fluids. II: magnetohydrodynamics, turbulence, and waves. Phys. Fluids 5, 257. DOI .

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Zaqarashvili, T.V., Erdélyi, R.: 2009, Oscillations and waves in solar spicules. Space Sci. Rev. 149, 355. DOI .

    Article  ADS  Google Scholar 

  • Zaqarashvili, T.V., Vörös, Z., Zhelyazkov, I.: 2014, Kelvin–Helmholtz instability of twisted magnetic flux tubes in the solar wind. Astron. Astrophys. 561, A62. DOI .

    Article  ADS  Google Scholar 

  • Zaqarashvili, T.V., Zhelyazkov, I., Ofman, L.: 2015, Stability of rotating magnetized jets in the solar atmosphere. I. Kelvin–Helmholtz instability. Astrophys. J. 813, 123. DOI .

    Article  ADS  Google Scholar 

  • Zaqarashvili, T.V., Díaz, A.J., Oliver, R., Ballester, J.L.: 2010, Instability of twisted magnetic tubes with axial mass flows. Astron. Astrophys. 516, A84. DOI .

    Article  ADS  MATH  Google Scholar 

  • Zhelyazkov, I.: 2012, Magnetohydrodynamic waves and their stability status in solar spicules. Astron. Astrophys. 537, A124. DOI .

    Article  ADS  MATH  Google Scholar 

  • Zhelyazkov, I.: 2013, Kelvin–Helmholtz instability of kink waves in photospheric, chromospheric, and X-ray solar jets. In: Zhelyazkov, I., Mishonov, T. (eds.) Space Plasma Physics: Proceedings of the 4th School and Workshop on Space Plasma Physics, AIP Conf. Proc. 1551, 150. DOI .

    Chapter  MATH  Google Scholar 

  • Zhelyazkov, I., Chandra, R.: 2018, High mode magnetohydrodynamic waves propagation in a twisted rotating jet emerging from a filament eruption. Mon. Not. Roy. Astron. Soc. 478, 5505. DOI .

    Article  ADS  Google Scholar 

  • Zhelyazkov, I., Chandra, R., Srivastava, A.K.: 2016, Kelvin–Helmholtz instability in an active region jet observed with Hinode. Astrophys. Space Sci. 361, 51. DOI .

    Article  ADS  Google Scholar 

  • Zhelyazkov, I., Zaqarashvili, T.V.: 2012, Kelvin–Helmholtz instability of kink waves in photospheric twisted flux tubes. Astron. Astrophys. 547, A14. DOI .

    Article  ADS  Google Scholar 

  • Zhelyazkov, I., Zaqarashvili, T.V., Ofman, L., Chandra, R.: 2018, Kelvin–Helmholtz instability in a twisting solar polar coronal hole jet observed by SDO/AIA. Adv. Space Res. 61, 628. DOI .

    Article  ADS  Google Scholar 

Download references

Acknowledgements

Our work was supported by the Bulgarian Science Fund under project DNTS/INDIA 01/7. The authors are indebted to the anonymous reviewer for pointing out a mathematical error and for the helpful and constructive comments and suggestions that contributed to improving the final version of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to I. Zhelyazkov.

Ethics declarations

Disclosure of Potential Conflicts of Interest

The authors declare that they have no conflicts of interest.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix A: Derivation of the Wave Dispersion Relation

Appendix A: Derivation of the Wave Dispersion Relation

We recall that we modeled the spinning macrospicule as a rotating and axially moving twisted magnetic flux tube of radius \(r = a\). In a cylindrical coordinate system, the magnetic and velocity fields inside the jet are assumed to be

$$\bigl( 0, B_{\mathrm{i}\phi }(r), B_{\mathrm{i}z} \bigr) \quad \mbox{and} \quad \bigl( 0, U_{\phi }(r), U_{z} \bigr), $$

respectively. Linearized ideal MHD equations, which govern the incompressible dynamics of perturbations in the rotating jet, are

$$\begin{aligned} &\frac{\partial }{\partial t}\boldsymbol{v} + (\boldsymbol{U}\cdot \nabla ) \boldsymbol{v} + (\boldsymbol{v}\cdot \nabla ) \boldsymbol{U} = - \frac{\nabla p_{\mathrm{tot}}}{\rho _{\mathrm{i}}} + \frac{ ( \boldsymbol{B}_{\mathrm{i}} \cdot \nabla )\boldsymbol{b}}{ \rho _{\mathrm{i}} \mu } + \frac{ ( \boldsymbol{b} \cdot \nabla ) \boldsymbol{B}_{\mathrm{i}}}{\rho _{\mathrm{i}} \mu }, \end{aligned}$$
(8)
$$\begin{aligned} & \frac{\partial }{\partial t}\boldsymbol{b} - \nabla \times ( \boldsymbol{v} \times \boldsymbol{B}_{\mathrm{i}} ) - \nabla \times ( \boldsymbol{U} \times \boldsymbol{b} ) = 0, \end{aligned}$$
(9)
$$\begin{aligned} & \nabla \cdot \boldsymbol{v} = 0, \end{aligned}$$
(10)
$$\begin{aligned} & \nabla \cdot \boldsymbol{b} = 0, \end{aligned}$$
(11)

where \(\boldsymbol{v} = (v_{r}, v_{\phi }, v_{z})\) and \(\boldsymbol{b} = (b_{r}, b_{\phi }, b_{z})\) are the perturbations of fluid velocity and magnetic field, respectively, and \(p_{\mathrm{tot}}\) is the perturbation of the total pressure \(p_{\mathrm{t}}\).

Assuming that all perturbations are proportional to \(g(r)\exp [\mathrm{i} ( -\omega t + m \phi + k_{z} z ) ]\), with \(g(r)\) being just a function of \(r\), we obtain from the above set of equations the following ones:

$$\begin{aligned} & -\mathrm{i}\sigma v_{r} -2\frac{U_{\phi }}{r}v_{\phi }- \mathrm{i}\frac{f _{B}}{\mu \rho _{\mathrm{i}}}b_{r} + 2\frac{B_{\mathrm{i}\phi }}{ \mu \rho _{\mathrm{i}} r}b_{\phi }= -\mathrm{i}\frac{1}{\rho _{ \mathrm{i}}}\frac{\mathrm{d}p_{\mathrm{tot}}}{\mathrm{d}r}, \end{aligned}$$
(12)
$$\begin{aligned} & -\mathrm{i}\sigma v_{\phi }+ \frac{1}{r} \frac{\mathrm{d} ( r U _{\phi } )}{\mathrm{d}r}v_{r} - \mathrm{i}\frac{f_{B}}{\mu \rho _{\mathrm{i}}}b_{\phi }- \frac{1}{\mu \rho _{\mathrm{i}}} \frac{1}{r}\frac{\mathrm{d} ( r B_{\mathrm{i}\phi } )}{ \mathrm{d}r}b_{r} = - \mathrm{i} \frac{1}{\rho _{\mathrm{i}}} \frac{m}{r}p_{\mathrm{tot}}, \end{aligned}$$
(13)
$$\begin{aligned} & -\mathrm{i}\sigma v_{z} - \mathrm{i} \frac{f_{B}}{\mu \rho _{\mathrm{i}}}b_{z} = -\mathrm{i}\frac{1}{ \rho _{\mathrm{i}}}k_{z} p_{\mathrm{tot}}, \end{aligned}$$
(14)
$$\begin{aligned} & -\mathrm{i}\sigma b_{r} - \mathrm{i}f_{B} v_{r} = 0, \end{aligned}$$
(15)
$$\begin{aligned} & -\mathrm{i}\sigma b_{\phi }- r\frac{\mathrm{d}}{\mathrm{d}r} \biggl( \frac{U _{\phi }}{r} \biggr)b_{r} - \mathrm{i}f_{B} v_{\phi }+ r\frac{ \mathrm{d}}{\mathrm{d}r} \biggl( \frac{B_{\mathrm{i}\phi }}{r} \biggr)v _{r} = 0, \end{aligned}$$
(16)
$$\begin{aligned} & -\mathrm{i}\sigma b_{z} - \mathrm{i}f_{B} v_{z} = 0, \end{aligned}$$
(17)
$$\begin{aligned} & \biggl( \frac{\mathrm{d}}{\mathrm{d}r} + \frac{1}{r} \biggr)v_{r} + \mathrm{i}\frac{m}{r}v_{\phi }+ \mathrm{i}k_{z} v_{z} = 0, \end{aligned}$$
(18)

where

$$ \sigma = \omega - \frac{m}{r}U_{\phi }- k_{z} U_{z} $$
(19)

is the Doppler-shifted frequency and

$$ f_{B} = \frac{m}{r}B_{\mathrm{i}\phi } + k_{z} B_{\mathrm{i}z}. $$
(20)

It is now convenient to introduce the Lagrangian displacement, \(\boldsymbol{\xi }\), and express it via the fluid velocity perturbation through the relation (Chandrasekhar, 1961)

v= ξ t +(U)ξ(ξ)U,

which yields

$$ v_{r} = -\mathrm{i}\sigma \xi _{r}, \qquad v_{\phi }= -\mathrm{i}\sigma \xi _{\phi }- r\frac{\mathrm{d}}{ \mathrm{d}r} \biggl( \frac{U_{\phi }}{r} \biggr)\xi _{r}, \qquad v_{z} = - \mathrm{i}\sigma \xi _{z}. $$
(21)

In terms of \(\boldsymbol{\xi }\), Equations 12 – 18 can be rewritten as

$$\begin{aligned} &\biggl[ \sigma ^{2} - \omega _{\mathrm{Ai}}^{2} - r\frac{\mathrm{d}}{ \mathrm{d}r} \biggl( \frac{U_{\phi }^{2}}{r^{2}} \biggr) + \frac{1}{ \mu \rho _{\mathrm{i}}}r \frac{\mathrm{d}}{\mathrm{d}r} \biggl( \frac{B _{\mathrm{i}\phi }^{2}}{r^{2}} \biggr) \biggr]\xi _{r} - 2\mathrm{i} \biggl( \sigma \frac{U_{\phi }}{r} + \frac{1}{r} \frac{B_{\mathrm{i} \phi }f_{B}}{\mu \rho _{\mathrm{i}}} \biggr)\xi _{\phi } \\ &\quad = \frac{1}{ \rho _{\mathrm{i}}} \frac{\mathrm{d}p_{\mathrm{tot}}}{\mathrm{d}r}, \end{aligned}$$
(22)
$$\begin{aligned} & \bigl( \sigma ^{2} - \omega _{\mathrm{Ai}}^{2} \bigr)\xi _{\phi }+ 2 \mathrm{i} \biggl( \sigma \frac{U_{\phi }}{r} + \frac{1}{r}\frac{B_{ \mathrm{i}\phi }f_{B}}{\mu \rho _{\mathrm{i}}} \biggr)\xi _{r} = \mathrm{i} \frac{1}{\rho _{\mathrm{i}}}\frac{m}{r}p_{\mathrm{tot}}, \end{aligned}$$
(23)
$$\begin{aligned} & \bigl( \sigma ^{2} - \omega _{\mathrm{Ai}}^{2} \bigr)\xi _{z} = \mathrm{i} \frac{1}{\rho _{\mathrm{i}}}k_{z} p_{\mathrm{tot}}, \end{aligned}$$
(24)
$$\begin{aligned} & \biggl( \frac{\mathrm{d}}{\mathrm{d}r} + \frac{1}{r} \biggr)\xi _{r} + \mathrm{i}\frac{m}{r}\xi _{\phi }+ \mathrm{i}k_{z} \xi _{z} = 0, \end{aligned}$$
(25)

where

$$ \omega _{\mathrm{Ai}} = \frac{f_{B}}{\sqrt{\mu \rho _{\mathrm{i}}}} $$
(26)

is the local Alfvén frequency inside the jet.

Excluding \(\xi _{\phi }\) and \(\xi _{z}\) from these equations, we obtain

$$\begin{aligned} & \rho _{\mathrm{i}} \bigl( \sigma ^{2} - \omega _{\mathrm{Ai}}^{2} \bigr) \biggl( \frac{\mathrm{d}\xi _{r}}{\mathrm{d}r} + \frac{\xi _{r}}{r} \biggr) + 2\rho _{\mathrm{i}}d_{2} \frac{m}{r}\xi _{r} = \biggl( \frac{m ^{2}}{r^{2}} + k_{z}^{2} \biggr)p_{\mathrm{tot}}, \end{aligned}$$
(27)
$$\begin{aligned} & \rho _{\mathrm{i}}d_{1} \xi _{r} = \bigl( \sigma ^{2} - \omega _{ \mathrm{Ai}}^{2} \bigr) \frac{\mathrm{d}p_{\mathrm{tot}}}{\mathrm{d}r} - 2\frac{m}{r}d_{2} p _{\mathrm{tot}}, \end{aligned}$$
(28)

where

$$d_{1} = ( \sigma ^{2} - \omega _{\mathrm{Ai}}^{2} )^{2} - r ( \sigma ^{2} - \omega _{\mathrm{Ai}}^{2} ) [ \frac{ \mathrm{d}}{\mathrm{d}r} ( \frac{U_{\phi }^{2}}{r^{2}} ) - \frac{1}{\mu \rho _{\mathrm{i}}}\frac{\mathrm{d}}{\mathrm{d}r} ( \frac{B _{\mathrm{i}\phi }^{2}}{r^{2}} ) ] - 4d_{2}^{2}, \; d _{2} = \sigma \frac{U_{\phi }}{r} + \frac{B_{\mathrm{i}\phi } f_{B}}{ \mu \rho _{\mathrm{i}}r}. $$

By presenting \(\xi _{r}\) from Equation 28 in terms of \(p_{\mathrm{tot}}\) and inserting it into Equation 27, we obtain the following equation for the total pressure perturbation:

$$\begin{aligned} &\biggl[ \bigl( \sigma ^{2} - \omega _{\mathrm{Ai}}^{2} \bigr)\frac{ \mathrm{d}}{\mathrm{d}r} + \frac{\sigma ^{2} - \omega _{\mathrm{Ai}} ^{2} + 2m d_{2}}{r} \biggr] \\ &\quad {}\times \biggl[ \frac{\sigma ^{2} - \omega _{\mathrm{Ai}}^{2}}{d_{1}}\frac{ \mathrm{d}p_{\mathrm{tot}}}{\mathrm{d}r} - \frac{2md_{2}}{d_{1}} \frac{p _{\mathrm{tot}}}{r} \biggr] - \biggl( \frac{m^{2}}{r^{2}} + k_{z} ^{2} \biggr)p_{\mathrm{tot}} = 0. \end{aligned}$$
(29)

This equation can be significantly simplified by considering that the rotation and the magnetic twist of the jet are uniform, that is,

$$ U_{\phi }(r) = \Omega r \quad \mbox{and} \quad B_{\phi }(r) = A r, $$
(30)

where \(\Omega \) and \(A\) are constants. In this case, Equation 29 takes the form of the Bessel equation

$$ \frac{\mathrm{d}^{2}p_{\mathrm{tot}}}{\mathrm{d}r^{2}} + \frac{1}{r}\frac{ \mathrm{d}p_{\mathrm{tot}}}{\mathrm{d}r} - \biggl( \frac{m^{2}}{r^{2}} + \kappa _{\mathrm{i}}^{2} \biggr)p_{\mathrm{tot}} = 0, $$
(31)

where

$$ \kappa _{\mathrm{i}}^{2} = k_{z}^{2} \biggl[ 1 - 4 \biggl( \frac{\sigma \Omega + A\omega _{\mathrm{Ai}}/\!\sqrt{\mu \rho _{\mathrm{i}}}}{ \sigma ^{2} - \omega _{\mathrm{Ai}}^{2}} \biggr)^{2} \biggr]. $$
(32)

The equation governing plasma dynamics outside the jet (without twist and velocity field, i.e. \(A = 0\), \(\Omega = 0\), and \(U_{z} = 0\)) is the same Bessel equation, but \(\kappa _{\mathrm{i}}\) is replaced by \(k_{z}\).

Inside the jet, the solution to Equation 31 bounded on the jet axis is the modified Bessel function of the first kind,

$$ p_{\mathrm{tot}}(r \leqslant a) = \alpha _{\mathrm{i}}I_{m}( \kappa _{\mathrm{i}}r), $$
(33)

where \(\alpha _{\mathrm{i}}\) is a constant.

Outside the jet, the solution bounded at infinity is the modified Bessel function of the second kind,

$$ p_{\mathrm{tot}}(r > a) = \alpha _{\mathrm{e}}K_{m}(k_{z} r), $$
(34)

where \(\alpha _{\mathrm{e}}\) is a constant.

To obtain the dispersion equation governing the propagation of MHD modes along the jet, the solutions at the jet surface need to be merged through boundary conditions.

It is well known that for non-rotating and untwisted magnetic flux tubes, the boundary conditions are the continuity of the Lagrangian radial displacement and total pressure perturbation at the tube surface (Chandrasekhar, 1961), that is,

$$ \xi _{\mathrm{i}r}|_{r=a} = \xi _{\mathrm{e}r}|_{r=a} \quad \mbox{and} \quad p_{\mathrm{tot\,i}} \bigg| _{r=a} = p_{\mathrm{tot\,e}}|_{r=a}, $$
(35)

where total pressure perturbations \(p_{\mathrm{tot\,i}}\) and \(p_{\mathrm{tot\,e}}\) are given by Equations 33 and 34, respectively.

When the magnetic flux tube is twisted and still non-rotating and the twist has a discontinuity at the tube surface, then the Lagrangian total pressure perturbation is continuous and the boundary conditions are (see, e.g., Bennett, Roberts, and Narain, 1999; Zaqarashvili et al., 2010; Zaqarashvili, Vörös, and Zhelyazkov, 2014)

$$ \xi _{\mathrm{i}r}|_{r=a} = \xi _{\mathrm{e}r}|_{r=a} \quad \mbox{and} \quad p_{\mathrm{tot\,i}} - \frac{B_{\mathrm{ i}\phi }^{2}}{\mu a} \xi _{\mathrm{i}r} \bigg| _{r=a} = p_{\mathrm{tot\,e}}|_{r=a}. $$
(36)

The second term in the second boundary condition stands for the pressure from the magnetic tension force.

If a non-twisted, \(B_{\mathrm{i}\phi } = 0\), tube rotates and the rotation has discontinuity at the tube surface, the boundary condition for the Lagrangian total pressure perturbation has a similar form as the second boundary condition in Equation 36,

$$ \xi _{\mathrm{i}r}|_{r=a} = \xi _{\mathrm{e}r}|_{r=a} \quad \mbox{and} \quad p_{\mathrm{tot\,i}} + \frac{\rho _{\mathrm{i}} U_{\phi }^{2}}{a}\xi _{\mathrm{i}r}\bigg| _{r=a} = p_{\mathrm{tot\,e}}|_{r=a}. $$
(37)

Here, the second term that describes the contribution of the centrifugal force to the pressure balance can be derived from Equation 28 for \(B_{\mathrm{i}\phi } = 0\) by multiplying that equation by \(\mathrm{d}r\), and considering the limit of \(\mathrm{d}r \to 0\) through the boundary \(r = a\), one obtains the relation \(\mathrm{d} [ p_{\mathrm{tot}} + (\rho _{\mathrm{i}} U _{\phi }^{2}/a)\xi _{r} ] = 0\), or, equivalently, the second boundary condition in the above equation. Hence, the boundary condition for the Lagrangian total pressure perturbation in rotating and magnetically twisted flux tubes has the form

$$ p_{\mathrm{tot\,i}} + \biggl( \frac{\rho _{\mathrm{i}} U_{ \phi }^{2}}{a} - \frac{B_{\mathrm{i}\phi }^{2}}{\mu a} \biggr) \xi _{\mathrm{i}r} \bigg| _{r=a} = p_{\mathrm{tot\,e}}|_{r=a}. $$
(38)

In the case of uniform rotation and magnetic field twist, Equation 30, the boundary conditions for the Lagrangian radial displacement \(\xi _{r}\) and the total pressure perturbation \(p_{\mathrm{tot}}\) are

$$ \xi _{\mathrm{i}r}|_{r=a} = \xi _{\mathrm{e}r}|_{r=a} \quad \mbox{and} \quad p_{\mathrm{tot\,i}} + a \biggl( \rho _{\mathrm{i}} \Omega ^{2} - \frac{A^{2}}{\mu } \biggr)\xi _{\mathrm{i}r} \bigg| _{r=a} = p_{\mathrm{tot\,e}}|_{r=a}. $$
(39)

Using these boundary conditions, we obtain the dispersion equation of normal MHD modes propagating in rotating and axially moving twisted magnetic flux tubes (Zaqarashvili, Zhelyazkov, and Ofman, 2015)

$$\begin{aligned} & \frac{ ( \sigma ^{2} - \omega _{\mathrm{Ai}}^{2} )F_{m}( \kappa _{\mathrm{i}}a) - 2m ( \sigma \Omega + A\omega _{ \mathrm{Ai}}/\! \sqrt{\mu \rho _{\mathrm{i}}} )}{\rho _{ \mathrm{i}} ( \sigma ^{2} - \omega _{\mathrm{Ai}}^{2} )^{2} - 4\rho _{\mathrm{i}} ( \sigma \Omega + A\omega _{\mathrm{Ai}}/\! \sqrt{ \mu \rho _{\mathrm{i}}} )^{2}} \\ &\quad {}= \frac{P_{m}(k_{z} a)}{\rho _{\mathrm{e}} ( \sigma ^{2} - \omega _{\mathrm{Ae}}^{2} ) - ( \rho _{\mathrm{i}}\Omega ^{2} - A^{2}/\mu )P_{m}(k_{z} a)}, \end{aligned}$$
(40)

where

$$F_{m}(\kappa _{\mathrm{i}}a) = \frac{\kappa _{\mathrm{i}}aI_{m}^{\prime }(\kappa _{\mathrm{i}}a)}{I_{m}(\kappa _{\mathrm{i}}a)}, \qquad P_{m}(k _{z} a) = \frac{k_{z} aK_{m}^{\prime }(k_{z} a)}{K_{m}(k_{z} a)}, \qquad \omega _{\mathrm{Ae}} = \frac{k_{z} B_{\mathrm{e}}}{\sqrt{ \mu \rho _{\mathrm{e}}}}. $$

We note that in the case of a non-rotating twisted flux tube, \(\Omega = 0\), the above Equation 40 recovers the well-known dispersion relation of normal MHD modes propagating in cylindrical twisted jets (see, e.g., Zhelyazkov and Zaqarashvili, 2012). If the environment medium is a cool plasma, as is the case of our macrospicule, where the thermal pressure \(p_{\mathrm{e}} = 0\), the \(k_{z} a\) in \(P_{m}(k_{z} a)\) must be replaced by \(k_{z} a[ 1 - ( \omega / \omega _{\mathrm{Ae}} )^{2} ]^{1/2}\), which yields the wave dispersion relation we used in Equation 3.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhelyazkov, I., Chandra, R. Can High-Mode Magnetohydrodynamic Waves Propagating in a Spinning Macrospicule Be Unstable Due to the Kelvin–Helmholtz Instability?. Sol Phys 294, 20 (2019). https://doi.org/10.1007/s11207-019-1408-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11207-019-1408-8

Keywords

Navigation