Skip to main content
Log in

Evolution of the magnetic field in spatially inhomogeneous axion structures

  • Research Articles
  • Published:
Theoretical and Mathematical Physics Aims and scope Submit manuscript

Abstract

We study the time evolution of magnetic fields in various configurations of spatially inhomogeneous pseudoscalar fields that are a coherent superposition of axions. For such systems, we derive a new induction equation for the magnetic field, which takes this inhomogeneity into account. Based on this equation, we study the evolution of a pair of Chern–Simons waves interacting with a linearly decreasing pseudoscalar field. The nonzero gradient of the pseudoscalar field leads to the mixing of these waves. We then consider the problem in a compact domain in the case where the initial Chern–Simons wave is mirror symmetric. The pseudoscalar field inhomogeneity then leads to an effective change in the \(\alpha\) dynamo parameter. Thus, the influence of a spatially inhomogeneous pseudoscalar field on the magnetic field evolution bears a strong dependence on the system geometry.

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.

Fig. 1.
Fig. 2.
Fig. 3.

Similar content being viewed by others

References

  1. R. D. Peccei and H. R. Quinn, “CP conservation in the presence of pseudoparticles,” Phys. Rev. Lett., 38, 1440–1443 (1977).

    Article  ADS  Google Scholar 

  2. L. D. Duffy and K. van Bibber, “Axions as dark matter particles,” New J. Phys., 11, 105008, 21 pp. (2009); arXiv: 0904.3346.

    Article  ADS  Google Scholar 

  3. Y. K. Semertzidis and S. Youn, “Axions dark matter: how to see it?,” Sci. Adv., 8, eabm9928, 13 pp. (2022); arXiv: 2104.14831.

    Article  Google Scholar 

  4. G. Galanti and M. Roncadelli, “Axion-like particles implications for high-energy astrophysics,” Universe, 8, 253, 72 pp. (2022); arXiv: 2205.00940.

    Article  ADS  Google Scholar 

  5. D. J. E. Marsh, “Axion cosmology,” Phys. Rep., 643, 1–79 (2016); arXiv: 1510.07633.

    Article  ADS  MathSciNet  Google Scholar 

  6. J. Barranco and A. Bernal, “Self-gravitating system made of axions,” Phys. Rev. D, 83, 043525, 5 pp. (2011); arXiv: 1001.1769.

    Article  ADS  Google Scholar 

  7. E. W. Kolb and I. I. Tkachev, “Axion miniclusters and Bose stars,” Phys. Rev. Lett., 71, 3051–3054 (1993); arXiv: hep-ph/9303313.

    Article  ADS  Google Scholar 

  8. J. Enander, A. Pargner, and T. Schwetz, “Axion minicluster power spectrum and mass function,” J. Cosmol. Astropart. Phys., 2017, 038 (2017); arXiv: 1708.04466.

    Article  Google Scholar 

  9. L. Visinelli and J. Redondo, “Axion miniclusters in modified cosmological histories,” Phys. Rev. D, 101, 023008, 28 pp. (2020); arXiv: 1808.01879.

    Article  ADS  Google Scholar 

  10. M. Yu. Khlopov, A. S. Sakharov, and D. D. Sokoloff, “The nonlinear modulation of the density distribution in standard axionic CDM and its cosmological impact,” Nucl. Phys. B, Proc. Suppl., 72, 105–109 (1999); arXiv: hep-ph/9812286.

    Article  ADS  Google Scholar 

  11. A. V. Sokolov and A. Ringwald, “Electromagnetic couplings of axions,” arXiv: 2205.02605.

  12. K. Choi, S. H. Im, and C. S. Shin, “Recent progress in the physics of axions and axion-like particles,” Ann. Rev. Nucl. Part. Sci., 71, 225–252 (2021); arXiv: 2012.05029.

    Article  ADS  Google Scholar 

  13. A. Long and T. Vachaspati, “Implications of a primordial magnetic field for magnetic monopoles, axions, and Dirac neutrinos,” Phys. Rev. D, 91, 103522, 12 pp. (2015); arXiv: 1504.03319.

    Article  ADS  Google Scholar 

  14. M. Dvornikov and V. B. Semikoz, “Evolution of axions in the presence of primordial magnetic fields,” Phys. Rev. D, 102, 123526, 9 pp. (2020); arXiv: 2011.12712.

    Article  ADS  MathSciNet  Google Scholar 

  15. J.-C. Hwang and H. Noh, “Axion electrodynamics and magnetohydrodynamics,” Phys. Rev. D, 106, 023503, 8 pp. (2022); arXiv: 2203.03124.

    Article  ADS  MathSciNet  Google Scholar 

  16. M. Dvornikov, “Interaction of inhomogeneous axions with magnetic fields in the early universe,” Phys. Lett. B, 829, 137039, 7 pp. (2022); arXiv: 2201.10586.

    Article  MathSciNet  Google Scholar 

  17. F. Anzuini, J. A. Pons, A. Gómez-Bañón, P. D. Lasky, F. Bianchini, and A. Melatos, “Magnetic dynamo caused by axions in neutron stars,” Phys. Rev. Lett., 130, 071001, 7 pp. (2023); arXiv: 2211.10863.

    Article  ADS  MathSciNet  Google Scholar 

  18. E. Braaten and H. Zhang, “Colloquium: The physics of axion stars,” Rev. Mod. Phys., 91, 041002, 22 pp. (2019); arXiv: 2211.10863.

    Article  ADS  MathSciNet  Google Scholar 

  19. V. I. Arnold and B. A. Khesin, Topological Methods in Hydrodynamics (Applied Mathematical Sciences, Vol. 125), Springer, New York (2021).

    Google Scholar 

  20. P. M. Akhmet’ev, S. Candelaresi, and A. Y. Smirnov, “Minimum quadratic helicity states,” J. Plasma Phys., 84, 775840601, 16 pp. (2018); arXiv: 1806.07428.

    Article  Google Scholar 

  21. A. M. Kamchatnov, “Topological solitons in magnetohydrodynamics,” Sov. Phys. JETP, 55, 69–73 (1982).

    ADS  MathSciNet  Google Scholar 

  22. G. Hornig and C. Mayer, “Towards a third-order topological invariant for magnetic fields,” J. Phys. A: Math. Gen., 35, 3945–3959 (2002); arXiv: physics/0203048.

    Article  ADS  MathSciNet  Google Scholar 

  23. P. M. Akhmet’ev and I. V. Vyugin, “Dispersion of the Arnold’s asymptotic ergodic Hopf invariant and a formula for its calculation,” Arnold Math. J., 6, 199–211 (2020); arXiv: 1906.12131.

    Article  MathSciNet  Google Scholar 

  24. E. A. Kudryavtseva, “Helicity is the only invariant of incompressible flows whose derivative is continuous in the \(C^1\) topology,” Math. Notes, 99, 611–615 (2016); arXiv: 1511.03746.

    Article  MathSciNet  Google Scholar 

  25. V. Z. Grines and L. M. Lerman, “Nonautonomous dynamics: classification, invariants, and implementation,” Contemp. Math. Fund. Directions, 68, 596–620 (2022).

    Article  MathSciNet  Google Scholar 

  26. A. Brandenburg, K. Enqvist, and P. Olesen, “Large-scale magnetic fields from hydromagnetic turbulence in the very early universe,” Phys. Rev. D, 54, 1291–1300 (1996); arXiv: astro-ph/9602031.

    Article  ADS  Google Scholar 

  27. L. Husdal, “On effective degrees of freedom in the early Universe,” Galaxies, 4, 78, 28 pp. (2016); arXiv: 1609.04979.

    Article  ADS  Google Scholar 

  28. L. Visinelli, “Boson stars and oscillatons: A review,” Internat. J. Modern Phys. D, 30, 2130006, 48 pp. (2021); arXiv: 2109.05481.

    Article  ADS  MathSciNet  Google Scholar 

  29. P. Charbonneau, “Dynamo models of the solar cycle,” Living Rev. Solar Phys., 17, 4, 104 pp. (2020).

    Article  ADS  Google Scholar 

  30. A. V. Gruzinov and P. H. Diamond, “Self-consistent theory of mean-field electrodynamics,” Phys. Rev. Lett., 72, 1651–1653 (1994).

    Article  ADS  Google Scholar 

  31. Y. Bai and Y. Hamada, “Detecting axion stars with radio telescopes,” Phys. Lett. B, 781, 187–194 (2018); arXiv: 1709.10516.

    Article  ADS  Google Scholar 

  32. M. Buschmann, C. Dessert, J. W. Foster, A. J. Long, and B. R. Safdi, “Upper limit on the QCD axion mass from isolated neutron star cooling,” Phys. Rev. Lett., 128, 091102, 9 pp. (2022); arXiv: 2111.09892.

    Article  ADS  Google Scholar 

  33. F. Chadha-Day, J. Ellis, and D. J. E. Marsh, “Axion dark matter: What is it and why now?,” Sci. Adv., 8, eabj3618, 20 pp. (2022); arXiv: 2105.01406.

    Article  Google Scholar 

  34. P. M. Akhmet’ev, “Magnetic helicity flux for mean magnetic field equations,” Theoret. and Math. Phys., 204, 947–956 (2020).

    Article  ADS  MathSciNet  Google Scholar 

  35. P. M. Akhmet’ev, S. Candelaresi, and A. Yu. Smirnov, “Calculations for the practical applications of quadratic helicity in MHD,” Phys. Plasmas, 24, 102128, 9 pp. (2017); arXiv: 1710.08833.

    Article  ADS  Google Scholar 

  36. D. G. Levkov, A. G. Panin, and I. I. Tkachev, “Gravitational Bose–Einstein condensation in the kinetic regime,” Phys. Rev. Lett., 121, 151301, 5 pp. (2018); arXiv: 1804.05857.

    Article  ADS  Google Scholar 

  37. P. Jain, S. Panda, and S. Sarala, “Electromagnetic polarization effects due to axion-photon mixing,” Phys. Rev. D, 66, 085007, 9 pp. (2002); arXiv: hep-ph/0206046.

    Article  ADS  Google Scholar 

  38. G. G. Raffelt, “Axion-photon oscillations. Mixing equations,” in: Stars as Laboratories for Fundamental Physics. Stars as Laboratories for Fundamental Physics: The Astrophysics of Neutrinos, Axions, and Other Weakly Interacting Particles, University of Chicago Press, Chicago (1996), pp. 179–181.

    Google Scholar 

  39. D. Harari and P. Sikivie, “Effects of a Nambu–Goldstone boson on the polarization of radio galaxies and the cosmic microwave background,” Phys. Lett. B, 289, 67–72 (1992).

    Article  ADS  Google Scholar 

Download references

Acknowledgments

We are grateful to E. M. Maslov for the useful discussions.

Funding

This work was supported by ongoing institutional funding. No additional grants to carry out or direct this particular research were obtained.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to P. M. Akhmet’ev.

Ethics declarations

The authors of this work declare that they have no conflicts of interest.

Additional information

Translated from Teoreticheskaya i Matematicheskaya Fizika, 2024, Vol. 218, pp. 601–618 https://doi.org/10.4213/tmf10574.

Publisher’s note. Pleiades Publishing remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix Kelvin transformation

We suppose that the Kelvin transformation is a stereographic projection of the sphere \(S^3\) outside the indicated point \(pt\) onto the Euclidean space \(\mathbb{R}^3\). This transformation is conformal, i.e., the angle between two vectors remains unchanged. In [21], the Kelvin transformation was used to construct the MHD soliton in Euclidean space with regular conditions at infinity.

It is noteworthy that the Kelvin transformation \(T\colon S^3 \setminus \{pt\} \to \mathbb{R}^3\) does not change the equations. In the case where the selected point \(pt\) is the north pole on the sphere (\(\psi=\theta=0\)), the magnetic mode \(\mathbf{B}_\mathrm{A}\) transforms into the (generalized) toroidal mode \(\mathbb{B}_\mathrm{A}\); and the magnetic mode \(\mathbf{B}_\mathrm {B}\), into the (generalized) poloidal field \(\mathbb{B}_\mathrm{B}\). We calculate only the transformation of magnetic modes.

The Kelvin transformation \(T\) is defined as

$$ (\phi,\theta,\psi) \mapsto (r=\cot(\theta),\Psi,\phi),$$
(A.1)
where, in the image, the spherical system of coordinates with the latitude \(\Psi\) is used and the longitude \(\phi\) is taken into account. The image of the coordinate \(\psi\) is calculated explicitly. It is independent of the coordinate \(\phi\) of the image. The absolute value of the gradient of \(\cot(\theta)\) determines the modulus of the conformal transformation \(T\). This scaling coefficient, which is a function in the image, is denoted by \(K\).

The Kelvin transformation maps the magnetic field \(\mathbf{B}\) (\(\mathbf{B}=\mathbf{B}_\mathrm {A}\) or \(\mathbf{B}=\mathbf{B}_\mathrm {B}\)), which is a divergence-free field on the initial sphere, into the magnetic field \(\mathbb{B}= T_{\ast}(\mathbf{B})/K^3\) in Euclidean space, where \(K\) is the scaling coefficient defined above. This function behaves asymptotically as \(\propto r^2\), where \(T_{\ast}\) is the translation of the vector via the differential of \(T\). The vector field \(\mathbb{B}\) is a magnetic field, because the flux of \(\mathbb{B}\) that passes through the surface \(T(\Sigma) \subset \mathbb{R}^3\) is equal to the flux of \(\mathbf{B}\) through the surface \(\Sigma \subset S^3\).

It can be seen that \(\mathbb{B}_\mathrm{B} = T_{\ast}(\mathbf{B}_\mathrm{B})K^{-3}\) varies with the longitude \(\phi\) and has the form of a toroidal magnetic mode. The mode \(\mathbb{B}_\mathrm{A}=T_{\ast}(\mathbf{B}_\mathrm{A})K^{-3}\) is directed perpendicular to the coordinates \((r,\Psi)\) and is a poloidal one. The scaling coefficient \(K\) is related to the form of the magnetic volume \(K^3 \mathbf{d} x\) in \(\mathbb{R}^3\) and has the asymptotic behavior \(\propto r^6\) along the radius. Both fields \(\mathbb{B}_\mathrm{A}\) and \(\mathbb{B}_\mathrm{B}\) share the asymptotic behavior \(\propto r^{-4}\).

The operator \(\operatorname{curl}\) in \(\mathbb{R}^3\) can be rewritten in the form

$$ \operatorname{curl}(K^{-2}T_{\ast}(F)) = K^{-3}T_{\ast}(\operatorname{curl}_\mathrm{S3}(F)),$$
(A.2)
where the operator \(\operatorname{curl}_\mathrm{S3}\) and the vector field \(F\) lie at the origin of coordinates of \(S^3\). In particular, at infinity, where \(K=K(r) \to +\infty\), the magnetic modes are almost potential and do not produce currents.

The magnetic helicity is preserved by the transformations \(T\). The helicity is calculated as an improper integral of the form

$$ \chi_{\mathbb{B}} =\int_{\mathbb{R}^3} (\mathbb{A} \cdot \mathbb{B})\,\mathbf{d} x = \int_{\mathbb{R}^3} (K^{-2}T_{\ast}(\mathbf{A}),K^{-3}T_{\ast}(\mathbf{B}))\, \mathbf{d} x = \int_{S^3} ((\mathbf{A} \cdot \mathbf{B}))\, \mathbf{d} V = \chi_{\mathbf{B}},$$
(A.3)
where \(\operatorname{curl}_\mathrm{S3}(\mathbf{A})=\mathbf{B}\). Here, the formulas \( K^3\mathbf{d} V = \mathbf{d} x \) and \( K^2((\dots,\dots)) = (\dots, \dots)\) relating the volume forms and the scalar products in the original image and the image are used.

Arnold’s inequality

Arnold’s inequality [19, Theorem 1.5, Ch. III] for the magnetic field on \(S^3\) is written as

$$ \int_{S^3} (\mathbf{B} \cdot \mathbf{B})\, \mathbf{d} V \ge\frac{1}{2R}\int_{S^3} (\mathbf{A} \cdot \mathbf{B})\, \mathbf{d} V = \frac{1}{2R}\chi_{\mathbf{B}},$$
(A.4)
where \(1/2R\) is the quantity that is inverse to the smallest eigenvalue of the operator \(\operatorname{curl}\); \(R\) is the sphere radius, which is the scaling coefficient; and \(\chi_\mathbf{B}\) is the magnetic helicity.

Arnold’s inequality is generalized to the case of an unbounded conducting domain with the variable magnetic permeability \(K(r)\) as

$$ \int_{\mathbb{R}^3} (\mathbb{B})^2 K^3\, \mathbf{d} x = \int_{S^3} (\mathbf{B})^2 K^{-1}\, \mathbf{d} V \ge \max_{S^3}(K^{-1}) \int_{S^3} (\mathbf{B})^2\, \mathbf{d} V \ge \frac{1}{2R}\chi_{\mathbf{B}},$$
(A.5)
where \(r\) is the distance to the origin and \(R\) is the scale of the magnetic permeability inhomogeneity.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dvornikov, M.S., Akhmet’ev, P.M. Evolution of the magnetic field in spatially inhomogeneous axion structures. Theor Math Phys 218, 515–529 (2024). https://doi.org/10.1134/S0040577924030103

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S0040577924030103

Keywords

Navigation