Skip to main content
Log in

Dissipative soliton dynamics of the Landau–Lifshitz–Gilbert equation

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

Abstract

We study ferromagnetic dissipative systems described by the isotropic LLG equation, from the standpoint of their spatially localized dynamical excitations. In particular, we focus on dissipative soliton solutions of a nonlocal NLS equation to which the LLG equation is transformed and use Melnikov’s theory to prove the existence of these solutions for sufficiently small dissipation. Next, we employ pseudospectral and PINN (physics-informed neural network) numerical techniques of machine learning to demonstrate the validity of our analytic results. Such localized structures have been detected experimentally in magnetic systems and observed in nano-oscillators, while dissipative magnetic droplet solitons have also been found theoretically and experimentally.

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.
Fig. 4.
Fig. 5.
Fig. 6.

Similar content being viewed by others

References

  1. M. Lakshmanan, Th. W. Ruijgrok, and C. J. Thompson, “On the dynamics of continuum spin systems,” Phys. A, 84, 577–590 (1976).

    Article  MathSciNet  Google Scholar 

  2. L. D. Landau and E. M. Lifshitz, “On the theory of the dispersion of magnetic permeability in ferromagnetic bodies,” Phys. Z. Sowjetunion, 8, 153–164 (1935).

    MATH  Google Scholar 

  3. T. L. Gilbert, “A phenomenological theory of damping in ferromagnetic materials,” IEEE Trans. Magn., 40, 3443–3449 (2004).

    Article  ADS  Google Scholar 

  4. D. C. Mattis, The Theory of Magnetism I: Statics and Dynamics (Springer Series in Solid-State Sciences, Vol. 17), Springer, Berlin (1988).

    Google Scholar 

  5. M. D. Stiles and J. Miltat, “Spin-transfer torque and dynamics,” in: Spin Dynamics in Confined Magnetic Structures III (Topics in Applied Physics, Vol. 101, B. Hillebrands and A. Thiaville, eds.), Springer, Berlin–Heidelberg (2006), pp. 225–308.

    Chapter  Google Scholar 

  6. M. Lakshmanan, “The fascinating world of the Landau–Lifshitz–Gilbert equation: an overview,” Philos. Trans. Roy. Soc. London Ser. A, 369, 1280–1300 (2011).

    ADS  MathSciNet  MATH  Google Scholar 

  7. O. Descalzi, M. Clerc, S. Residori, and G. Assanto (eds.), Localized States in Physics: Solitons and Patterns, Springer, Berlin–Heidelberg (2011).

    Book  Google Scholar 

  8. Z. Ahsan and K. R. Jayaprakash, “Evolution of a primary pulse in the granular dimers mounted on a linear elastic foundation: An analytical and numerical study,” Phys. Rev. E, 94, 043001, 15 pp. (2016).

    Article  ADS  Google Scholar 

  9. W. Królikowski, O. Bang, N. I. Nikolov, D. Neshev, J. Wyller, J. J. Rasmussen, and D. Edmundson, “Modulational instability, solitons and beam propagation in spatially nonlocal nonlinear media,” J. Opt. B: Quantum Semiclass. Opt., 6, S288–S294 (2004).

    Article  ADS  Google Scholar 

  10. A. N. Slavin and V. S. Tiberkevich, “Spin wave mode excited by spin-polarized current in a magnetic nanocontact is a standing self-localized wave bullet,” Phys. Rev. Lett., 95, 237201, 4 pp. (2005).

    Article  ADS  Google Scholar 

  11. M. A. Hoefer, T. J. Silva, and M. W. Keller, “Theory for a dissipative droplet soliton excited by a spin torque nanocontact,” Phys. Rev. B, 82, 054432, 14 pp. (2010).

    Article  ADS  Google Scholar 

  12. S. M. Mohseni, S. R. Sani, J. Persson et al., “Spin torque-generated magnetic droplet solitons,” Science, 339, 1295–1298 (2013).

    Article  ADS  Google Scholar 

  13. I. V. Barashenkov, S. R. Woodford, and E. V. Zemlyanaya, “Interactions of parametrically driven dark solitons. I. Néel–Néel and Bloch–Bloch interactions,” Phys. Rev. E, 75, 026604, 18 pp. (2007).

    Article  ADS  MathSciNet  Google Scholar 

  14. V. Besse, H. Leblond, D. Mihalache, and B. A. Malomed, “Pattern formation by kicked solitons in the two-dimensional Ginzburg–Landau medium with a transverse grating,” Phys. Rev. E, 87, 012916, 12 pp. (2013).

    Article  ADS  Google Scholar 

  15. Y. Lan and Y. C. Li, “Chaotic spin dynamics of a long nanomagnet driven by a current,” Nonlinearity, 21, 2801–2823 (2008).

    Article  ADS  MathSciNet  MATH  Google Scholar 

  16. M. Daniel and M. Lakshmanan, “Perturbation of solitons in the classical continuum isotropic Heisenberg spin system,” Phys. A, 120, 125–152 (1983).

    Article  Google Scholar 

  17. St. Wiggins, Introduction to Applied Nonlinear Dynamical Systems and Chaos (Texts in Applied Mathematics, Vol. 2), Springer, New York (2005).

    MATH  Google Scholar 

  18. J. Gruendler, “The existence of homoclinic orbits and the method of Melnikov for systems in \(\mathbb R^{n\,}\),” SIAM J. Math. Anal., 16, 907–931 (1985).

    Article  MathSciNet  MATH  Google Scholar 

  19. M. Yamashita, “Melnikov vector in higher dimensions,” Nonlinear Anal. Theory Methods Appl., 18, 657–670 (1992).

    Article  MathSciNet  MATH  Google Scholar 

  20. S.-N. Chow and M. Yamashita, “Geometry of the Melnikov vector,” Math. Sci. Eng., 185, 79–148 (1992).

    Article  MathSciNet  MATH  Google Scholar 

  21. V. M. Rothos and T. C. Bountis, “Mel’nikov analysis of phase space transport in a \(N\)-degree- of- freedom Hamiltonian system,” Nonlinear Anal. Theory Methods Appl., 30, 1365–1374 (1997).

    Article  MATH  Google Scholar 

  22. J. Yang, Nonlinear Waves in Integrable and Nonintegrable Systems (Mathematical Modeling and Computation, Vol. 16), SIAM, Philadelphia, PA (2010).

    Book  MATH  Google Scholar 

  23. M. Raissi, P. Perdikaris, and G. E. Karniadakis, “Physics-informed neural networks: a deep learning framework for solving forward and inverse problems involving nonlinear partial differential equations,” J. Comput. Phys., 378, 686–707 (2019).

    Article  ADS  MathSciNet  MATH  Google Scholar 

  24. V. M. Rothos and A. F. Vakakis, “Dynamic interactions of traveling waves propagating in a linear chain with local essentially nonlinear attachment,” Wave Motion, 46, 174–188 (2009).

    Article  ADS  MathSciNet  MATH  Google Scholar 

  25. K. J. Palmer, “Exponential dichotomy and transversal homoclinic orbits,” J. Differ. Equ., 55, 225–256 (1984).

    Article  ADS  MATH  Google Scholar 

  26. P. Tsilifis, P. G. Kevrekidis, V. M. Rothos, “Cubic-quintic long-range interactions with double well potentials,” J. Phys. A: Math. Theor., 47, 035201, 21 pp. (2014); arXiv: 1207.6731.

    Article  ADS  MATH  Google Scholar 

Download references

Funding

T. Bountis acknowledges that his work on Sections 2, 3.1 and 3.2 of this paper was supported by the Russian Science Foundation project No. 21-71-30011. T. Bountis also acknowledges partial support for Section 3.3 by the grant No. AP08856381 of the Science Committee of the Ministry of Education and Science of the Republic of Kazakhstan, for the project of the Institute of Mathematics and Mathematical Modeling MES RK, Almaty, Kazakhstan.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to T. Bountis.

Ethics declarations

The authors declare no conflicts of interest.

Additional information

Prepared from an English manuscript submitted by the authors; for the Russian version, see Teoreticheskaya i Matematicheskaya Fizika, 2023, Vol. 215, pp. 190–206 https://doi.org/10.4213/tmf10404.

Appendix: Homoclinic Melnikov theory of integro-differential equations

Here, we briefly review the theory of a more general class of perturbed nonlocal dynamical systems [24], which we used in Sec. 2 to study the existence and stability of bright solitons in the nonlocal NLS (2.2).

We start by considering a perturbed \(n\)-dimensional dynamical system of the integro-differential form

$$ \dot x=f(x)+\epsilon [\kern1pt\mathcal N\kern1.5pt] (x,t;\mu),\qquad x\in\mathbb{R}^{n},$$
(A.1)
where \(0\le|\epsilon|\ll 1\), \( [\kern1pt\mathcal N\kern1.5pt] \) are periodic functions in \(t\), and \(\mu\in\mathbb{R}\) is the set of external parameters of the problem. The nonlocal part \(\mathcal N\) is quite general and includes different classes of terms that arise frequently in applications.

We assume that the unperturbed system (A.1) (corresponding to \(\epsilon=0\)) has a hyperbolic singular point \(x_0\) and a homoclinic orbit \(\gamma\) to \(x_0\). The linear variational equation of unperturbed system (A.1) along the homoclinic orbit \(\gamma\), written in the form

$$ \dot z=A(t)z,\qquad A(t)=\mathrm Df(\gamma(t)),$$
(A.2)
is known to have dichotomies on \(\alpha\le t\le\infty\) and \(-\infty\le t\le\alpha\) for \(\alpha\in\mathbb{R}\). Clearly, when \(A(t)=A\) is constant, system (A.2) has an exponential dichotomy on the infinite interval if and only if \(\operatorname{spec}A=\{\lambda\in\mathbb{C}\colon\operatorname{Re}\lambda\neq 0\}\), whereas when \(A(t)=A(t+T)\), Eq. (A.2) has an exponential dichotomy if and only if the Floquet multipliers lie outside the unit circle [25].

We now let \(W^{\mathrm s}_0\) and \(W^{\mathrm u}_0\) denote the respective stable and unstable manifolds of the singular point \(x_0\), and define the following projections for a given \(\alpha\in\mathbb{R}\):

$$P(a)\colon T_{\gamma(\alpha)}\mathbb{R}^{n}\to T_{\gamma(\alpha)}W^{\mathrm s}_0,\qquad Q(a)\colon T_{\gamma(\alpha)}\mathbb{R}^{n}\to T_{\gamma(\alpha)}W^{\mathrm u}_0$$
satisfying,
$$(I-P(\alpha))T_{\gamma(\alpha)}\mathbb{R}^{n}=T^{\perp}_{\gamma(\alpha)}W^{\mathrm s}_0,\qquad (I-Q(\alpha))T_{\gamma(\alpha)}\mathbb{R}^{n}=T^{\perp}_{\gamma(\alpha)}W^{\mathrm u}_0.$$
Here, \(T_{\gamma(\alpha)}\mathbb{R}^{n}\) and \(T^{\perp}_{\gamma(\alpha)}W^{\mathrm s}_0\) are the tangent spaces to \(\mathbb{R}^{n}\) and to \(W^{\mathrm s}_0\) at \(\gamma(\alpha)\). System (A.2) has an exponential dichotomy on \(\alpha\le t\le\infty\), such that \(\Phi(t,\alpha)z_0\to 0\) exponentially as \(t\to\infty\) if \(z_0\in T_{\gamma(\alpha)}W^{\mathrm u}_0\), where \(\Phi(t,s)\) is the Floquet transition matrix of (A.2).

We now define the perturbed stable invariant manifold of the fixed point \(x_\epsilon\) as

$$ W^{\mathrm s}_{\mathrm{loc}}(x_\epsilon)= \bigcup_{\alpha\in\mathbb{R}}\bigl\{\gamma(\alpha)+\epsilon\kern1pt G^{\kern1pt\mathrm s}_2(\alpha,\eta^{\mathrm s};\epsilon;\mu)\bigr\},$$
(A.3)
where
$$G^{\kern1pt\mathrm s}_2(\alpha,\eta^{\mathrm s};\epsilon;\mu)= \eta^{\mathrm s}+(I-P(\alpha))\int^{\alpha}_{\infty}\Phi(\alpha,\tau) [\kern1pt\mathcal N\kern1.5pt] (\gamma(\tau),\tau-\alpha;\mu)\,d\tau$$
and \(\eta^{\mathrm s}\in T_{\gamma(\alpha)}W^{\mathrm s}_0/T_{\gamma(\alpha)}\gamma(\alpha)\), \(|\eta^{\mathrm s}|\ll 1\). Similarly, for the perturbed unstable manifold, we have
$$ W^{\mathrm u}_{\mathrm{loc}}(x_\epsilon)= \bigcup_{\alpha\in\mathbb{R}}\bigl\{\gamma(\alpha)+\epsilon\kern1pt G^{\mathrm u}_2(\alpha,\eta^{\mathrm u};\epsilon;\mu)\bigr\}.$$
(A.4)
Here,
$$G^{\mathrm u}_2(\alpha,\eta^{\mathrm u};\epsilon;\mu)= \eta^{\mathrm u}+(I-Q(\alpha))\int^{\alpha}_{-\infty}\Phi(\alpha,\tau) [\kern1pt\mathcal N\kern1.5pt] (\gamma(\tau),\tau-\alpha;\mu)\,d\tau$$
where \(\eta^{\mathrm u}\in T_{\gamma(\alpha)}W^{\mathrm u}_0/T_{\gamma(\alpha)}\gamma(\alpha)\), \(|\eta^{\mathrm u}|\ll 1\). Choosing
$$ x(t)=\gamma(t+\alpha)+\epsilon z_2(t+\alpha),\qquad \alpha\in\mathbb{R}$$
(A.5)
as a perturbation of the homoclinic orbit of (A.1), we obtain the variational equations
$$ \dot z_2=A(t)z_2+ [\kern1pt\mathcal N\kern1.5pt] (\gamma(t), t-\alpha;\mu)=h_2(t,x,z,\epsilon),$$
(A.6)
where
$$ h_2(t,x,z,\epsilon)= \frac{1}\epsilon\bigl\{f(\gamma(t)+\epsilon z_2)-f(\gamma(t))z_2\big\}+ [\kern1pt\mathcal N\kern1.5pt] (\gamma(t)+\epsilon z_2)- [\kern1pt\mathcal N\kern1.5pt] (\gamma(t)).$$
(A.7)
Let \(z_2(t;\alpha,z_0)\) be a solution of system (A.6) with \(z_2(0;\alpha,z_0)=z_{0,2}\). Following [20], it is straightforward to conclude that \(z_2(t;\alpha,z_0)\) is bounded for \(\alpha\le t<\infty\) if and only if it satisfies
$$\begin{aligned} \, z_2(t;\alpha,z_0)={}& \Phi(t,\alpha) \biggl[P(\alpha)z_{0,2}+P(\alpha)\int^t_{\alpha}\Phi(\alpha,\tau)\{ [\kern1pt\mathcal N\kern1.5pt] (\gamma(\tau),\tau-\alpha;\mu)+h_2(\tau,\alpha,z_2;\epsilon)\}\,d\tau+{} \nonumber\\ &+(I-P(\alpha))\int^t_{\infty}\Phi(\alpha,\tau)\{ [\kern1pt\mathcal N\kern1.5pt] (\gamma(\tau),\tau-\alpha;\mu)+h_2(\tau,\alpha,z_2;\epsilon)\}\,d\tau\biggr]. \end{aligned}$$
(A.8)

Letting \(\eta^{\mathrm s}_2=P(\alpha)z_{0,2}\), we can prove, using the contraction mapping theorem, that system (A.8) has the unique solution

$$z^{\mathrm s}_2=z_2(t;\alpha, z_{0,2}(\eta^{\mathrm s}_2)),\qquad |\eta^{\mathrm s}_1|\ll 1,\quad |\eta^{\mathrm s}_2|\ll 1.$$
Setting now \(t=\alpha\) in (A.8), we obtain
$$ z_{0,2}(\eta^{\mathrm s}_2)=\eta^{\mathrm s}_2+ (I-P(\alpha))\int^{\alpha}_{\infty}\Phi(\alpha,\tau)\{ [\kern1pt\mathcal N\kern1.5pt] (\gamma(\tau),\tau-\alpha;\mu)+h_2(\tau,\alpha,z_2;\epsilon)\}\,d\tau,$$
(A.9)
whence
$$ x(0)=\gamma(\alpha)+\epsilon \biggl[\eta^{\mathrm s}_2+(I-P(\alpha))\int^{\alpha}_{\infty}\Phi(\alpha,\tau) \{ [\kern1pt\mathcal N\kern1.5pt] (\gamma(\tau),\tau-\alpha;\mu)+h_2(\tau,\alpha,z_2;\epsilon;\mu)\}\,d\tau\biggr].$$
(A.10)
It follows that we can obtain the perturbed local stable and unstable manifolds of the Poincaré map of the perturbed dynamical system (A.1), \(W^{\mathrm s}_{\mathrm{loc}}(x_\epsilon)\) and \(W^{\mathrm u}_{\mathrm{loc}}(x_\epsilon)\), as the graphs on the unperturbed stable and unstable manifolds \(W^{\mathrm s}_0\) and \(W^{\mathrm u}_0\). The functions \(G^{\mathrm s}(\alpha,\eta^{\mathrm s},\epsilon;\mu) \) and \(G^{\mathrm u}(\alpha,\eta^{\mathrm u},\epsilon;\mu) \) decompose as (see [20])
$$\begin{aligned} \, &G^{\mathrm s}(\alpha,\eta^{\mathrm s},\epsilon;\mu)= (\alpha,\nu,\eta^{\mathrm s}_2,m^{\mathrm s}_2(\alpha,\nu,\eta^{\mathrm s}_1,\eta^{\mathrm s}_2,\epsilon;\mu)), \\ &G^{\mathrm u}(\alpha,\eta^{\mathrm u},\epsilon;\mu)= (\alpha,\nu,\eta^{\mathrm u}_2,m^{\mathrm u}_2(\alpha,\nu,\eta^{\mathrm u}_1,\eta^{\mathrm u}_2,\epsilon;\mu)), \end{aligned}$$
where \((\alpha,\nu)\in\operatorname{Range}P(\alpha)\cap\operatorname{Range}Q(\alpha)\) and \(\eta^{\mathrm s,\mathrm u}=(\nu,\eta^{\mathrm s,\mathrm u}_2)\) with
$$\eta^{\mathrm s}_2\in\operatorname{Range}P(\alpha)\cap\operatorname{Range}(I-Q(\alpha)),\qquad \eta^{\mathrm u}_2\in\operatorname{Range}(I-P(\alpha))\cap\operatorname{Range}Q(\alpha)$$
(see [20]).

To measure the distance between the stable and unstable manifolds of the perturbed Poincaré map, it suffices to measure the separation of these manifolds in the subspace \(\operatorname{Range}(I-P(\alpha))\cap\operatorname{Range}(I-Q(\alpha))\). Let \(\mathcal D\) denote the distance between \(W^{\mathrm s}_{\mathrm{loc}}(x_{\epsilon,\mu})\) and \(W^{\mathrm u}_{\mathrm{loc}}(x_{\epsilon,\mu})\):

$$ \mathcal D=m^{\mathrm u}_2(\alpha,\nu,\eta^{\mathrm u}_2,\epsilon;\mu)-m^{\mathrm s}_2(\alpha,\nu,\eta^{\mathrm s}_2,\epsilon;\mu).$$
(A.11)

We consider the problem adjoint to (A.2),

$$ \dot\phi+(\mathrm Df(\gamma(t)))^*\phi=0$$
(A.12)
and let \(\{\phi_i\}\), \(i=1,\ldots m\), be a set of linearly independent bounded solutions on \(\mathbb{R}\). Then
$$\operatorname{Range}(I-P^*(\alpha))\cap\operatorname{Range}(I-Q^*(\alpha))=\operatorname{span}\{\phi_i^*(\alpha)\}^{m}_{i=1}.$$
In terms of \(\phi_1(\alpha),\ldots,\phi_m(\alpha)\), the manifold separation \(\mathcal D\) has the coordinate expression
$$\begin{aligned} \, \phi^*_i(\alpha)\mathcal D={}&\phi^*_i(\alpha)(m^{\mathrm u}_2-m^{\mathrm s}_2) \biggl[\phi^*_i(\alpha)(I-Q(\alpha)) \int^{\alpha}_{-\infty}\!\!\Phi(\alpha,\tau) \{ [\kern1pt\mathcal N\kern1.5pt] (\gamma(\tau),\tau-\alpha;\mu)+h_2(\tau,\alpha,z^{\mathrm u}_1,z^{\mathrm u}_2;\epsilon)\}\,d\tau\biggr]-{} \\ &-\biggl[\phi^*_i(\alpha)(I-P(\alpha)) \int^{\alpha}_{\infty}\!\!\Phi(\alpha,\tau) \{ [\kern1pt\mathcal N\kern1.5pt] (\gamma(\tau),\tau-\alpha;\mu)+h_2(\tau,\alpha,z^{\mathrm s}_1,z^{\mathrm s}_2;\epsilon)\}\,d\tau\biggr]= \\ ={}&\biggl[\,\int^{\alpha}_{-\infty}\!\!\phi^*_i(\tau) \{ [\kern1pt\mathcal N\kern1.5pt] (\gamma(\tau),\tau-\alpha;\mu)+h_2(\tau,\alpha,z^{\mathrm u}_1,z^{\mathrm u}_2;\epsilon)\}\,d\tau\biggr]-{} \\ &-\biggl[\,\int^{\alpha}_{\infty}\!\!\phi^*_i(\tau) \{ [\kern1pt\mathcal N\kern1.5pt] (\gamma(\tau),\tau-\alpha;\mu)+h_2(\tau,\alpha,z^{\mathrm s}_1,z^{\mathrm s}_2;\epsilon)\}\,d\tau\biggr]= \\ ={}&\int^{\infty}_{-\infty}\!\!\phi^*(t) [\kern1pt\mathcal N\kern1.5pt] (\gamma(\tau),\tau-\alpha;\mu)\,dt+{} \\ &+\biggl[\,\int^{0}_{-\infty}\!\!\phi^*_i(\tau) h_2(\tau,\alpha,z^{\mathrm u}_1,z^{\mathrm u}_2;\epsilon)\,d\tau- \int^{0}_{\infty}\!\!\phi^*_i(\tau)h_2(\tau,\alpha,z^{\mathrm s}_1,z^{\mathrm s}_2;\epsilon)\,d\tau\biggr] \end{aligned}$$
for \(i=1,2,\ldots,m\).

We define \(\mathcal D=\epsilon M_2+O(\epsilon^2)\), where the components of the Melnikov vector \(M_2\) are given by

$$ M_{2,i}(\alpha,\nu)=\int^{\infty}_{-\infty}\phi^*_i(t) [\kern1pt\mathcal N\kern1.5pt] (\gamma(\tau),\tau-\alpha;\mu)\,dt,\qquad i=1,2,$$
(A.13)
where \(z_1\) is the solution of the first variational problem \(\dot z_1=\mathrm Df(\gamma(t))z_1\).

Thus, the conditions for the existence of transverse intersections between the stable and unstable manifolds of a Poincaré map associated with the system of integro-differential equations (A.1) can be summarized by the following theorem [24].

Theorem.

If there exist \(\alpha\), \(\nu=(\nu_1,\ldots,\nu_{m-1})\), and \(\epsilon\) such that the Melnikov vector \(M_1(\alpha,\nu)=0\) and the \(m\) vectors \(\frac{\partial M_1}{\partial\alpha},\frac{\partial M_1}{\partial\nu_1},\ldots,\frac{\partial M_1}{\partial\nu_{m-1}}\) are all nonzero at the point \((\alpha,\nu)\), then the local stable and unstable manifolds of the Poincaré map associated with system (A.1) intersect transversally. Moreover, in the case where the Melnikov vector \(M_1(\alpha,\nu)\equiv 0\), if there exist \(\alpha\), \(\nu=(\nu_1,\ldots,\nu_{m-1})\), and \(\epsilon\) such that \(M_2(\alpha,\nu)=0\) and the \(m\) vectors \(\frac{\partial M_2}{\partial\alpha},\frac{\partial M_2}{\partial\nu_1},\ldots,\frac{\partial M_2}{\partial\nu_{m-1}}\) are all nonzero at the point \((\alpha,\nu)\), then \(W^{\mathrm s}_{\mathrm{loc}}(x_\epsilon)\) and \(W^{\mathrm u}_{\mathrm{loc}}(x_\epsilon)\) intersect transversally.

We remark that the stability of bright solitons of the nonlocal NLS with cubic and quintic nonlocal terms was studied in [26] using a different approach to analyze their dynamical behavior.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rothos, V.M., Mylonas, I.K. & Bountis, T. Dissipative soliton dynamics of the Landau–Lifshitz–Gilbert equation. Theor Math Phys 215, 622–635 (2023). https://doi.org/10.1134/S0040577923050033

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S0040577923050033

Keywords

Navigation