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

On the derivation of guiding center dynamics without coordinate dependence

Zhi YU University of Science and Technology of China, Hefei, People’s Republic of China, 230027 Institute of Plasma Physics, Hefei Institutes of Physical Science, Chinese Academy of Sciences, Hefei, People’s Republic of China, 230031
Abstract

The fundament of the classical guiding center theory is gyro-phase averaging, which cannot be well defined over a non-trivial magnetic field topology. The local gyro-phase coordinate frame hides the geometric nature of gyro-symmetry. A coordinate-free geometric representation should be a more appropriate alternative for a deeper understanding of the guiding center dynamics. In this paper, the motion of a charged particle is described by a Lagrangian one-form on a seven-dimensional phase space. The Lagrangian one-form is geometrically decomposed by constructing a coordinate-free gyro-averaging method. As a result, we obtain the coordinate-free expression of the non-relativistic guiding-center dynamics in the time-dependent slow-varying electromagnetic field.

I Introduction

The guiding center dynamics has been a subject of interest to researchers for decades [1]. The purpose of guiding center theory is to decompose particle motion into vertical gyro-motion and horizontal drift motion. The classical decomposition method is based on the averaging of the gyro-phase. The gyro-phase is defined on a predefined local orthogonal coordinate frame, which may not exist globally over a non-trivial magnetic field. This issue was raised and discussed by Sugiyama [2, 3] and Krommes [4] in 2009. Soon after, Burdy and Qin [5] recognized that the magnetic field inhomogeneity obstructs the existence of the global gyro-phase. The global gyro-phase is not a necessary condition for gyro-symmetry. The gyro-symmetry depends only on the homogeneity of the electromagnetic field within the range of gyro-motion. The local gyro-phase coordinate frame hides the geometric nature of gyro-symmetry and prevents us from identifying this issue. A coordinate-free geometric representation should be a more appropriate alternative for a deeper understanding of the guiding center dynamics. A series of recent works by Burby on slow manifolds of near-periodic Hamiltonian systems has shown the importance of geometric tools for understanding gyro-symmetry [6, 7, 8].

The base of guiding-center theory is the symmetry of trajectories of the charged particles in electromagnetic fields, called gyro-symmetry. Kruskal [9] pointed out that the set of charged particles with a common guiding center constitutes a topological ring in phase space, called Kruskal’s ring [10]. In a slowly varying electromagnetic field, the charged particles on the same Kruskal’s ring have similar phase space trajectories. As long as the particles do not resonate with the field, Kruskal’s ring will not break, but will only be slightly deformed. In other words, the guiding center is the center of Kruskal’s ring. The decomposition of gyro-motion is to decompose the Kruskal’s ring from the phase space of particle trajectories.

This paper aims to construct coordinate-free guiding-center dynamics, or rather, Kruskal’s ring dynamics. The description of Kruskal’s ring relies on two vector fields, the roto-rate vector and the gyro-radius. The roto-rate vector is the generator of gyro-symmetry, named by Kruskal [9]. Omohundro[11] showed the coordinate-free expression of the roto-rate vector. The gyro-radius is the vector from the guiding center to the ringmates, representing the mapping between Kruskal’s ring and the guiding center. In an inhomogeneous electromagnetic field, gyro-symmetry is not absolute. The gyro-decomposition is an asymptotic approximation to the exact particle motion. The form of gyro-radius is not unique – different forms of gyro-radius correspond to different decompositions [12]. We will show that a proper definition of the rote rate vector and the gyro-radius may yield a concise expression of the guiding center dynamics.

In this paper, we use the Lagrangian one-form (or called Poincaré-Cartan one-form) to represent the motion of charge particle in a time-dependent slow-varying electromagnetic field. The extended phase space of particle is seven-dimensional contact manifold. The Lagrangian formalism provides simple and explicit expressions for the variational principle and Noether’s theorem [13]. From the geometric point of view, the Lagrangian one-form is the dual counterpart of the trajectory. The decomposition of the trajectory can be achieved by decomposing the Lagrangian one-from. We will show that the non-existence of global gyro-phase is a natural conclusion of the Lagrangian one-form decomposition. As the result go geometric descomposition, we obtain the coordinate-free expression of the non-relativistic guiding-center dynamics in the time-dependent slow-varying electromagnetic field.

The derivation in this paper uses knowledge of elementary differential geometry and Lie groups. For interested readers, Marsden and Ratiu’s book [14] would be a good reference.

This paper is organized as follows. In Section II, we recall the general geometric setting of the Lagrangian formalism for a time-dependent system and discuss the relation between Poincaré-Cartan integral invariant and Noether’s theorem. In Section III, we describe the decomposition of Kruskal’s ring and the general expression of the guiding center dynamics. The Section IV shows the coordinate-free expression of non-relativistic guiding center dynamics in the time-dependent electromagnetic field. And, the Section V is the summary and discussion.

II Geometric s etting

Considering a time-dependent Hamiltonian system with the extended phase space PP, the action integral is an line integral along phase space trajectory λ\lambda,

𝒜[λ]=λη,\mathcal{A}\left[\lambda\right]=\int_{\lambda}\eta\;, (1)

where one-form η=pdqHdt\eta=pdq-Hdt is called the Lagrangian one-form. From Hamiltonian principle, the variation of trajectory λ\lambda gives the Hamiltonian equations

ιτdη=0,\iota_{\tau}d\eta=0\,, (2)

where ιτ\iota_{\tau} means interior product with vector field τ\tau, and dηd\eta is the exterior derivative of one-form η\eta. The Hamiltonian flow is an one-parameter group of tt whose infinitesimal generator is τ\tau, ΨtH=exp(tτ).\Psi_{t}^{H}=\exp\left(t\tau\right). If we add a closed one-form to the Lagrangian one-form η=η+α\eta^{\prime}=\eta+\alpha and dα=0d\alpha=0, the result of Hamiltonian equations Eq.(2) does not change. The extended phase space PP is an 2n+12n+1 dimensional manifold endowed with a one-form η\eta that satisfies the nonintegrable condition η(dη)n0\eta\wedge\left(d\eta\right)^{n}\neq 0. This structure (P,η)\left(P,\eta\right) is a contact structure, and η\eta is also called contact form. The Lagrangian one-form (contact one-form) η\eta plays an important role in time-dependent mechanics, which provides simple and explicit expressions for the variational principles and Noether’s theorem [13, 14].

II.1 Poincaré-Cartan integral invariant

Consider a curve 𝒪\mathcal{O} encircles a tube of phase trajectories in extended phase space PP, the action integral on 𝒪\mathcal{O} is denoted as μP[𝒪]𝒪η\mu_{P}\left[\mathcal{O}\right]\equiv\oint_{\mathcal{O}}\eta. Let 𝒪\mathcal{O} move along the same tube of phase trajectories, the action integral on the image of Hamiltonian flow ΨtH\Psi_{t}^{H} looks like

μP[ΨtH𝒪]\displaystyle\mu_{P}\left[\Psi_{t}^{H}\circ\mathcal{O}\right] =ΨtH𝒪η=𝒪ΨtHη\displaystyle=\oint_{\Psi_{t}^{H}\circ\mathcal{O}}\eta=\oint_{\mathcal{O}}\Psi_{t}^{H*}\eta
=𝒪(η+tτη+)\displaystyle=\oint_{\mathcal{O}}\left(\eta+t\mathcal{L}_{\tau}\eta+\cdots\right)
=μP[𝒪]+t𝒪(iτdη+diτη)+.\displaystyle=\mu_{P}\left[\mathcal{O}\right]+t\oint_{\mathcal{O}}\left(\cancel{i_{\tau}d\eta}+di_{\tau}\eta\right)+\cdots\,. (3)

Because ιτdη=0\iota_{\tau}d\eta=0 and the circle integral of closed form diτηdi_{\tau}\eta is zero, the second and higher order terms of Eq.(3) will vanish. Then, we can say that the action integral on a closed phase space curve, μP[𝒪]\mu_{P}\left[\mathcal{O}\right], is a constant of motion. The one-form η\eta is also called the of Poincaré’s relative integral invariant or Poincaré-Cartan one-form[15]. The action integral μP[𝒪]\mu_{P}\left[\mathcal{O}\right] is indepedent with the shape of 𝒪\mathcal{O}, which captures the topological property of the bundle of trajectories. The Hamiltonian flow preserves the action integral over arbitrarily closed loop in phase space. However, without additional constraints, the Hamiltonian flow would not preserve the compactness of the loop 𝒪\mathcal{O}.

II.2 Noether’s theorem

Consider an one-parameter Lie group Φθs=exp(sθ),s\Phi_{\theta s}=\exp\left(s\partial_{\theta}\right),s\in\mathbb{R}, which is generated by a vector field θ\partial_{\theta}. The action integral of the infinitesimally transformed trajectory is

𝒜[Φθsλ]\displaystyle\mathcal{A}\left[\Phi_{\theta s}\circ\lambda\right] =Φθsλη=λΦθs(η)=λexp(sθ)η=𝒜[λ]+sλθη+O(s2),\displaystyle=\int_{\Phi_{\theta s}\circ\lambda}\eta=\int_{\lambda}\Phi_{\theta s}^{*}\left(\eta\right)=\int_{\lambda}\exp\left(s\mathcal{L}_{\partial_{\theta}}\right)\eta=\mathcal{A}\left[\lambda\right]+s\int_{\lambda}\mathcal{L}_{\partial_{\theta}}\eta+O\left(s^{2}\right)\,, (4)

where θ\mathcal{L}_{\partial_{\theta}} is the Lie derivative along the vector θ\partial_{\theta}. If θη=0\mathcal{L}_{\partial_{\theta}}\eta=0, the higher order term of ss will vanish and action integral 𝒜[λ]\mathcal{A}\left[\lambda\right] is preserved under transformation of the group Φθ\Phi_{\theta}. We shall say Φθ\Phi_{\theta} is a Noether symmetry on the Hamiltonian system (P,η)\left(P,\eta\right). Using the Cartan’s Magic Formula

0=θη\displaystyle 0=\mathcal{L}_{\partial_{\theta}}\eta =ιθdη+dιθη,\displaystyle=\iota_{\partial_{\theta}}d\eta+d\iota_{\partial_{\theta}}\eta\;, (5)

we get

ιθdη=dιθη=dμ,\iota_{\partial_{\theta}}d\eta=-d\iota_{\partial_{\theta}}\eta=-d\mu\,, (6)

where

μιθη,\mu\equiv\iota_{\partial_{\theta}}\eta\,, (7)

is the moment map induced by θ\partial_{\theta} [14]. Putting ιτ\iota_{\tau} on Eq. (5), yields

0=ιτιθdη=ιτdιθη=ιτdμ0=\iota_{\tau}\iota_{\partial_{\theta}}d\eta=-\iota_{\tau}d\iota_{\partial_{\theta}}\eta=-\iota_{\tau}d\mu (8)

It is easy to verify that μ\mu is also an invariant under the action of Φθ\Phi_{\theta}

θμ=0,Φθ(μ)=μ.\mathcal{L}_{\partial_{\theta}}\mu=0,\qquad\Phi_{\theta}^{*}\left(\mu\right)=\mu\,. (9)

Putting τ\mathcal{L}_{\tau} on Eq. (5), yields

τθη=τ(ιθdη+dιθη)=i[τ,θ]dηiθτdη+dτ(iθη)=0,\mathcal{L}_{\tau}\mathcal{L}_{\partial_{\theta}}\eta=\mathcal{L}_{\tau}\left(\iota_{\partial_{\theta}}d\eta+d\iota_{\partial_{\theta}}\eta\right)=i_{\left[\tau,\partial_{\theta}\right]}d\eta-\cancel{i_{\partial_{\theta}}\mathcal{L}_{\tau}d\eta}+d\mathcal{L}_{\tau}\left(i_{\partial_{\theta}}\eta\right)=0\,, (10)

where [,]\left[\cdot,\cdot\right] denotes the commutor of vector fields. If dηd\eta is not degenerate, the symmetry vector θ\partial_{\theta} should commutes with Hamiltonian vector [τ,θ]=0\left[\tau,\partial_{\theta}\right]=0.

Noether’s theorem requires the Lie derivative of Poincaré-Cartan one-form η\eta along the symmetry vector θ\partial_{\theta} vanish, which is only a local constraint on the one-form η\eta. The Noether’s theorem Eq.(6) can not tell us the global topology of the symmetry. For the same Lie algebra θ\partial_{\theta} , the orbit 𝒪\mathcal{O} of Lie group Φθ\Phi_{\theta} may be isomorphic to S1S^{1} or \mathbb{R}. The global topology of the Lie group Φθ\Phi_{\theta} depends on the nature of the extended phase space (P,η)\left(P,\eta\right). If the symmetry group Φθs\Phi_{\theta s} is a compact one-parameter group, it is isomorphic to the group S1U(1)S^{1}\cong U\left(1\right), whose orbit 𝒪\mathcal{O} is a closed curve[16]. The action integral on the closed orbit 𝒪\mathcal{O} is

μP[𝒪]=𝒪η=01Φθs(ιθη)𝑑s=μ.\mu_{P}\left[\mathcal{O}\right]=\oint_{\mathcal{O}}\eta=\int_{0}^{1}\Phi_{\theta s}^{*}\left(\iota_{\partial_{\theta}}\eta\right)ds=\mu\;. (11)

which is consistent with the moment map μ\mu.

Noether’s theorem determines the invariance of the moment map μ\mu. In many cases, there is no guarantee that the exact symmetry always exists everywhere in the phase space, and μ\mu will not be a globally valid exact invariant. If the inhomogeneity of the phase space is a small quantity in the range of the closed loop 𝒪\mathcal{O}, then the variation of μ\mu along the particle trajectory is bounded, and μ\mu is called adiabatic invariant [17]. Liouville’s theorem determines the invariance of action integral μP\mu_{P} on the closed loop 𝒪\mathcal{O}, which is an absolute invariant. Kruskal pointed out that points in the phase space of a near-periodic dynamical system form closed loops that drift along phase space trajectories, preserving their topology with only slight deformations [9]. The phase space inhomogeneity leads to the deformation of 𝒪\mathcal{O}, and the deviation between μ\mu and μP\mu_{P}. Since the deviation is bounded, we use μ\mu as the asymptotic approximation to μP\mu_{P}, which can preserve the invariance to arbitrary orders[7, 8]. We call these loops Kruskal’s ring or invariant tori in Arnold’s book [17]. The existence of Kruskal’s ring implies that the divergent Hamiltonian flow is constrained by a local compact group Φθ\Phi_{\theta}. Our aim is to decompose Kruskal’s ring 𝒪\mathcal{O} from the phase space PP to obtain a quotient manifold P/𝒪P/\mathcal{O}.

III Kruskal’s ring and guiding center

III.1 Kruskal’s ring

Consider the phase space (P,η)\left(P,\eta\right) and a local compact Lie group Φθ\Phi_{\theta} called gyro-transformation or gyro-symmetry. If the Lagrangian one-form η\eta is invariant to the action of Φθ\Phi_{\theta} throughout the phase space PP, we say the phase space PP is uniform to the gyro-transformation Φθ\Phi_{\theta}. The orbit of Φθ\Phi_{\theta} is called Kruskal’s ring 𝒪\mathcal{O}, and the points on the same orbit are called ringmates, 𝒪z={Φθ(z)}\mathcal{O}_{z}=\left\{\Phi_{\theta}\left(z\right)\right\}. The gyro-symmetry Φθ\Phi_{\theta} induce a moment map μ:P\mu:P\rightarrow\mathbb{R}, which projects Kruskal’s ring to a constant of motion along the particle trajectory. If the gyro-transformation has a fixed point Φθ(Z)=Z\Phi_{\theta}\left(Z\right)=Z, we called it the guiding center and use it as the representative of Kruskal’s ring 𝒪\mathcal{O}.

The trajectories of Kruskal’s ring constitute the quotient manifold P/𝒪P/\mathcal{O}. The guiding center trajectory is isomorphic to the Kruskal’s ring trajectory P¯P/𝒪\bar{P}\simeq P/\mathcal{O}. Then, a natural projection arises

Πρ:PP/𝒪×𝒪P¯×𝒪¯,\Pi_{-\rho}:P\rightarrow P/\mathcal{O}\times\mathcal{O}\rightarrow\bar{P}\times\mathcal{\bar{O}}\,, (12)

where the minus sign indicates the transformation is in the opposite direction of the gyro-radius, and the superscript ’bar’ indicates that it is defined at the guiding center. Let inverse projection Πρ\Pi_{\rho} be a one-parameter transformation generated by gyro-radius Πrρ=exp(rρ¯)\Pi_{r\rho}=\exp\left(r\partial_{\bar{\rho}}\right), the pullback from ringmate to guiding center is a formal power series of the Lie derivative ρ¯\mathcal{L}_{\partial_{\bar{\rho}}}

Πρ¯rα=n0rnn!ρ¯nα|Z,\Pi_{\bar{\rho}r}^{*}\alpha=\sum_{n\geq 0}\frac{r^{n}}{n!}\mathcal{L}_{\partial_{\bar{\rho}}}^{n}\left.\alpha\right|_{Z}\,, (13)

where α|Z\left.\alpha\right|_{Z} is the quantity defined at the guiding center.

Pushing forward θ\partial_{\theta} to the guiding center, yields the rote-rate vector

θ¯Πρθ=Πρ¯θ,\partial_{\bar{\theta}}\equiv\Pi_{-\rho*}\partial_{\theta}=\Pi_{\bar{\rho}}^{*}\partial_{\theta}\,, (14)

which is named by Kruskal [9]. The orbit of rote-rate vector θ¯\partial_{\bar{\theta}} is a circle in the velocity space

Φθ¯=exp(2πsθ¯)𝒪¯TZM,\Phi_{\bar{\theta}}=\exp\left(2\pi s\partial_{\bar{\theta}}\right)\leftrightarrow\bar{\mathcal{O}}\subset T_{Z}M\,, (15)

called limiting ring. Therefore, we say the guiding center is a particle with "spin", whose magnetic moment μ\mu is equal to the action integral over the limiting ring 𝒪¯\bar{\mathcal{O}}.

III.2 Decomposition

To decompose the motion of Kruskal’s ring, we split the Hamiltonian vector τ\tau into two parts, the horizontal part τ\tau_{\parallel} and the vertical part θ\partial_{\theta},

τ\displaystyle\tau =τ+θ.\displaystyle=\tau_{\parallel}+\partial_{\theta}\;. (16)

Substituting τ\tau to Eq.(2), yields the equation of horizontal motion

ιτdη\displaystyle\iota_{\tau_{\parallel}}d\eta =ιθdη=dμ.\displaystyle=-\iota_{\partial_{\theta}}d\eta=d\mu\;. (17)

Pushing forward τ\tau_{\parallel} to the guiding center, we get the Hamiltonian vector field of guiding center

τ¯=Πρτ=Πρ¯τ.\bar{\tau}_{\parallel}=\Pi_{-\rho*}\tau_{\parallel}=\Pi_{\bar{\rho}}^{*}\tau_{\parallel}\,. (18)

It is straightforward to verify that the horizontal motion is commute with vertical motions

[θ,τ]=[θ,τ]=0,\left[\partial_{\theta},\tau\right]=\left[\partial_{\theta},\tau_{\parallel}\right]=0\;, (19)

and μ\mu is a constant of motion in both directions

ιτdμ=ιθdμ=0.\iota_{\tau_{\parallel}}d\mu=\iota_{\partial_{\theta}}d\mu=0\,. (20)

We also split the Lagrangian one-form into two parts

η=η+η,\eta=\eta_{\parallel}+\eta_{\perp}\,, (21)

where the horizontal part is orthogonal to θ\partial_{\theta}

ιθη=0,\iota_{\partial_{\theta}}\eta_{\parallel}=0\,, (22)

and the vertical part gives the moment map μ\mu

ιθη=μ.\iota_{\partial_{\theta}}\eta_{\perp}=\mu\,. (23)

From Eq. (17), the horizontal vector field is given by

τ\displaystyle*\tau_{\parallel}^{\flat} =12!volθdμηdηdη,\displaystyle=\frac{1}{2!\text{vol}_{\theta}}d\mu\wedge\eta_{\perp}\wedge d\eta\wedge d\eta\,, (24)

where

volθ13!ηdηdηdη,*\text{vol}_{\theta}\equiv\frac{1}{3!}\eta_{\perp}\wedge d\eta\wedge d\eta\wedge d\eta\,, (25)

is the phase space volume form introduced by η\eta_{\perp}. Here, the Hodge operator * maps the pp-form to (np)\left(n-p\right)-form, and the superscript \flat means lowering of indices[18]. Substituting Eq.(24) into Eq.(17), one can verify that

ιτdη=dμvolμvolθη,\iota_{\tau_{\parallel}}d\eta=d\mu-\cancel{\frac{\text{vol}_{\mu}}{\text{vol}_{\theta}}\eta_{\perp}}\,, (26)

where

volμ\displaystyle\text{vol}_{\mu} =(13!dμdηdηdη)=ιτdμ=0,\displaystyle=*\left(\frac{1}{3!}d\mu\wedge d\eta\wedge d\eta\wedge d\eta\right)=\iota_{\tau}d\mu=0\,, (27)

is the degenerate phase space volume form introduced by dμd\mu.

III.3 Perturbation

The guiding center has practical significance only if the trajectories of the ringmates are similar. In other words, the deformation of Kruskal’s ring should be limited. In the uniform phase space, Kruskal’s ring 𝒪\mathcal{O} is a circle. The deformation of Kruskal’s ring came from the inhomogeneity of phase space. Let a small quantity ερ¯\varepsilon\sim\mathcal{L}_{\partial_{\bar{\rho}}} denote the phase space inhomogeneity with respect to the gyro-radius ρ¯\partial_{\bar{\rho}}. In a slowly varying system, ε\varepsilon\ll1, the deformation is a near-identity transformation generated by a perturbation vector field GG,

ΨεG:𝒪𝒪ε,,ΨεG=exp(εG).\Psi_{\varepsilon G}:\mathcal{O}\rightarrow\mathcal{O}_{\varepsilon},\quad,\Psi_{\varepsilon G}=\exp\left(\varepsilon G\right)\,. (28)

Further, we also have

𝒪ε=ΨεG𝒪=ΨεGΠρ¯𝒪¯=Πρε𝒪¯,\mathcal{O}_{\varepsilon}=\Psi_{\varepsilon G}\mathcal{O}=\Psi_{\varepsilon G}\circ\Pi_{\bar{\rho}}\circ\bar{\mathcal{O}}=\Pi_{\rho_{\varepsilon}}\circ\bar{\mathcal{O}}\;, (29)

where ΠρεΨεGΠρ¯\Pi_{\rho_{\varepsilon}}\equiv\Psi_{\varepsilon G}\circ\Pi_{\bar{\rho}} is the perturbed guiding-center projection. The perturbed gyro-transformation Φεθ\Phi_{\varepsilon\theta} is given by

ΦεθΠερΦθ¯Περ=ΨεGΠρ¯Φθ¯Πρ¯ΨεG=AdεGAdρ¯Φθ¯.\Phi_{\varepsilon\theta}\equiv\Pi_{\varepsilon\rho}\circ\Phi_{\bar{\theta}}\circ\Pi_{-\varepsilon\rho}=\Psi_{\varepsilon G}\circ\Pi_{\bar{\rho}}\circ\Phi_{\bar{\theta}}\circ\Pi_{-\bar{\rho}}\circ\Psi_{-\varepsilon G}=\text{Ad}_{\varepsilon G}\circ\text{Ad}_{\bar{\rho}}\circ\Phi_{\bar{\theta}}\;. (30)

Then, one can verify that the perturbed gyro-transformation Φεθ\Phi_{\varepsilon\theta} preserve the Noether’s theorem

𝒜[Φεθsλε]\displaystyle\mathcal{\mathcal{A}}\left[\Phi_{\varepsilon\theta s}\circ\lambda_{\varepsilon}\right] =Φεθsλεηε=λεΦεθs(ηε)\displaystyle=\int_{\Phi_{\varepsilon\theta s}\circ\lambda_{\varepsilon}}\eta_{\varepsilon}=\int_{\lambda_{\varepsilon}}\Phi_{\varepsilon\theta s}^{*}\left(\eta_{\varepsilon}\right)
=λεΠρεΦθ¯sΠρε(ηε)\displaystyle=\int_{\lambda_{\varepsilon}}\Pi_{-\rho_{\varepsilon}}^{*}\circ\Phi_{\bar{\theta}s}^{*}\circ\Pi_{\rho_{\varepsilon}}^{*}\left(\eta_{\varepsilon}\right)
=λεΠρεexp(sθ¯)Πρε(ηε)\displaystyle=\int_{\lambda_{\varepsilon}}\Pi_{-\rho_{\varepsilon}}^{*}\circ\exp\left(s\mathcal{L}_{\partial_{\bar{\theta}}}\right)\circ\Pi_{\rho_{\varepsilon}}^{*}\left(\eta_{\varepsilon}\right)
=𝒜[λε]+sΠρελεθ¯Πρε(ηε)+\displaystyle=\mathcal{A}\left[\lambda_{\varepsilon}\right]+s\int_{\Pi_{-\rho_{\varepsilon}}\circ\lambda_{\varepsilon}}\mathcal{L}_{\partial_{\bar{\theta}}}\Pi_{\rho_{\varepsilon}}^{*}\left(\eta_{\varepsilon}\right)+\cdots
=𝒜[λε]+sλ¯θ¯η¯+,\displaystyle=\mathcal{A}\left[\lambda_{\varepsilon}\right]+s\int_{\bar{\lambda}}\mathcal{L}_{\partial_{\bar{\theta}}}\bar{\eta}+\cdots\;, (31)

where the second and higher order terms will vanish at θ¯η¯=0\mathcal{L}_{\partial_{\bar{\theta}}}\bar{\eta}=0. Pullback μ¯\bar{\mu} to the perturbed trajectory, the perturbed moment map με=ιθεηε\mu_{\varepsilon}=\iota_{\partial_{\theta\varepsilon}}\eta_{\varepsilon} is an constant of motion along the perturbed trajectory

0=Πρε(ιτ¯dμ¯)=ιΠρετ¯dΠρεμ¯=ιτεdμε,0=\Pi_{-\rho_{\varepsilon}}^{*}\left(\iota_{\bar{\tau}}d\bar{\mu}\right)=\iota_{\Pi_{-\rho_{\varepsilon}}^{*}\bar{\tau}}d\Pi_{-\rho_{\varepsilon}}^{*}\bar{\mu}=\iota_{\tau_{\varepsilon}}d\mu_{\varepsilon}\;, (32)

Following the Lie perturbation method developed by Littlejohn[19, 20] and Cary [21, 22]. The perturbated moment map με\mu_{\varepsilon} may be calculated to arbitrary order.

The stable Kruskal’s ring implies that ringmates’ trajectories have symmetry Φθ𝒪\Phi_{\theta}\leftrightarrow\mathcal{O}. The deformed Kruskal’s ring correspond to the perturbated gyro-symmetry Φθε𝒪ε\Phi_{\theta\varepsilon}\leftrightarrow\mathcal{O}_{\varepsilon}. The action integral along the Kruskal’s ring μP[𝒪ε]\mu_{P}\left[\mathcal{O}_{\varepsilon}\right] is an exact invariant. As long as the symmetry group Φθ\Phi_{\theta} is compact, the exact invariant μP\mu_{P} always exist. Therefore, we say the perturbed moment map με\mu_{\varepsilon} is an asymptotical approximation of μP\mu_{P}.

III.4 Gyro-averaging

The inhomogeneities in phase space break the gyro-symmetry. The purpose of gyro-averaging is to eliminate the perturbation and to obtain the gyro-invariant unperturbed Lagrangian. The gyro-averaging is an integral along the Kruskal’s ring

α=01Φθs(α)𝑑s.\left\langle\alpha\right\rangle=\int_{0}^{1}\Phi_{\theta s}^{*}\left(\alpha\right)ds\,. (33)

Pulling it back to the limiting ring, yields

α\displaystyle\left\langle\alpha\right\rangle =01Φθ¯sΠρ¯ε(α)𝑑s=01n0εnn!Φθ¯sρ¯nΦθ¯s(α|Z)ds.\displaystyle=\int_{0}^{1}\Phi_{\bar{\theta}s}^{*}\circ\Pi_{\bar{\rho}\varepsilon}^{*}\left(\alpha\right)ds=\int_{0}^{1}\sum_{n\geq 0}\frac{\varepsilon^{n}}{n!}\mathcal{L}_{\Phi_{\bar{\theta}s}^{*}\partial_{\bar{\rho}}}^{n}\Phi_{\bar{\theta}s}^{*}\left(\left.\alpha\right|_{Z}\right)ds\;. (34)

If α|Z\left.\alpha\right|_{Z} is gyro-independent, the gyro-average α\left\langle\alpha\right\rangle only dependent on the rotation of gyro-radius Φθ¯sρ¯\Phi_{\bar{\theta}s}^{*}\partial_{\bar{\rho}}, which is a formal power series of θ¯\mathcal{L}_{\partial_{\bar{\theta}}}

Φθ¯sρ¯\displaystyle\Phi_{\bar{\theta}s}^{*}\partial_{\bar{\rho}} =exp(2πsθ¯)(ρ¯)=n0(2πs)nn!θ¯n(ρ¯).\displaystyle=\exp\left(2\pi s\mathcal{L}_{\partial_{\bar{\theta}}}\right)\left(\partial_{\bar{\rho}}\right)=\sum_{n\geq 0}\frac{\left(2\pi s\right)^{n}}{n!}\mathcal{L}_{\partial_{\bar{\theta}}}^{n}\left(\partial_{\bar{\rho}}\right)\,. (35)

Note that the form of gyro-radius ρ¯\partial_{\bar{\rho}} and rote-rate vector θ¯\partial_{\bar{\theta}} are not unique, as long as the mapping from the limiting ring 𝒪¯\bar{\mathcal{O}} to the Kruskal’s ring 𝒪\mathcal{O} holds,

Πρ¯𝒪¯𝒪θ=Πρ¯θ¯,\Pi_{\bar{\rho}}\bar{\mathcal{O}}\mapsto\mathcal{O}\leftrightarrow\partial_{\theta}=\Pi_{\bar{\rho}*}\partial_{\bar{\theta}}\,, (36)

To simplify the gyro-averaging, we let gyro-radius ρ¯\partial_{\bar{\rho}} be ’complex-like’ under the action of rote-rate vector θ¯\partial_{\bar{\theta}}

θ¯2ρ¯=ρ¯.\mathcal{L}_{\partial_{\bar{\theta}}}^{2}\partial_{\bar{\rho}}=-\partial_{\bar{\rho}}\,. (37)

Then, the rotation of gyro-radius Φθ¯sρ¯\Phi_{\bar{\theta}s}^{*}\partial_{\bar{\rho}} is a simple trigonometric polynomial

Φθ¯sρ¯\displaystyle\Phi_{\bar{\theta}s}^{*}\partial_{\bar{\rho}} =ρ+(2πs)ρˇ12(2πs)2ρ13!(2πs)3ρˇ+\displaystyle=\partial_{\rho}+\left(2\pi s\right)\partial_{\check{\rho}}-\frac{1}{2}\left(2\pi s\right)^{2}\partial_{\rho}-\frac{1}{3!}\left(2\pi s\right)^{3}\partial_{\check{\rho}}+\cdots
=n0(1)n(2πs)2n2n!ρ+n0(1)n(2πs)2n+1(2n+1)!ρˇ\displaystyle=\sum_{n\geq 0}\frac{\left(-1\right)^{n}\left(2\pi s\right)^{2n}}{2n!}\partial_{\rho}+\sum_{n\geq 0}\frac{\left(-1\right)^{n}\left(2\pi s\right)^{2n+1}}{\left(2n+1\right)!}\partial_{\check{\rho}}
=cos(2πs)ρ+sin(2πs)ρˇ\displaystyle=\cos\left(2\pi s\right)\partial_{\rho}+\sin\left(2\pi s\right)\partial_{\check{\rho}} (38)

where ρˇθ¯ρ¯\partial_{\check{\rho}}\equiv\mathcal{L}_{\partial_{\bar{\theta}}}\partial_{\bar{\rho}} is the dual vector orthogonal to ρ¯\partial_{\bar{\rho}}. And, the gyro-average α\left\langle\alpha\right\rangle is a polynomials in ρ¯\mathcal{L}_{\partial_{\bar{\rho}}} and ρˇ\mathcal{L}_{\partial_{\check{\rho}}}

α\displaystyle\left\langle\alpha\right\rangle =n01n!01𝑑s(cos(2πs)ρ¯+sin(2πs)ρˇ)nα|Z.\displaystyle=\sum_{n\geq 0}\frac{1}{n!}\int_{0}^{1}ds\left(\cos\left(2\pi s\right)\mathcal{L}_{\partial_{\bar{\rho}}}+\sin\left(2\pi s\right)\mathcal{L}_{\partial_{\check{\rho}}}\right)^{n}\left.\alpha\right|_{Z}\,. (39)

IV Charged particle motion in a slow-varying electromagnetic field

IV.1 Poincaré-Cartan-Einstein one-form

The motion of charged particle was considered in the four-dimensional Minkowski space M=E1,3M=E^{1,3} with global Cartesian coordinates xα(x0,xi)=(ct,𝐱)x^{\alpha}\equiv\left(x^{0},x^{i}\right)=\left(ct,\mathbf{x}\right) and the metric tensors takes the form g=(1,1,1,1)g=\left(-1,1,1,1\right). The phase space PP is the cotangent bundle TMT^{*}M of MM with constraint condition,

P{(x,p)|xM,pTxM,g1(p,p)=m2c2},P\equiv\left\{\left.\left(x,p\right)\right|x\in M,p\in T_{x}^{*}M,g^{-1}\left(p,p\right)=-m^{2}c^{2}\right\}\,, (40)

where p(p0,𝐩)p\equiv\left(p^{0},\mathbf{p}\right) is the four-momentum. For a charged particle in an electromagnetic field the Lagrangian one-form (also known as the Poincaré-Cartan-Einstein one-form [23] ) is given by

η\displaystyle\eta =\displaystyle= A+p=(𝐀+𝐩)d𝐱(ϕ+p0)dt,\displaystyle A+p=\left(\mathbf{A}+\mathbf{p}\right)\cdot d\mathbf{x}-\left(\phi+p^{0}\right)dt\,, (41)

where Aα(ϕ,𝐀)A^{\alpha}\equiv\left(\phi,\mathbf{A}\right) is the four-potential. The natural unit system is adopted , let m=c=e=1m=c=e=1, and only consider the non-relativistic case,

p0=𝐯22,𝐩=𝐯.p^{0}=\frac{\mathbf{v}^{2}}{2},\;\mathbf{p}=\mathbf{v}\,. (42)

The Hamiltonian vector τ\tau is solved from Hamiltonian equation Eq.(2),

{τ𝐱τt=d𝐱dt=𝐯,τ𝐯τt=d𝐯dt=𝐯×𝐁+𝐄,\begin{cases}\frac{\tau_{\mathbf{x}}}{\tau_{t}}=\frac{d\mathbf{x}}{dt}&=\mathbf{v}\;,\\ \frac{\tau_{\mathbf{v}}}{\tau_{t}}=\frac{d\mathbf{v}}{dt}&=\mathbf{v}\times\mathbf{B}+\mathbf{E}\;,\end{cases} (43)

where the electromagnetic fields are epxressed in terms of the potentials as 𝐁=×𝐀\mathbf{B}=\nabla\times\mathbf{A} and 𝐄=ϕt𝐀\mathbf{E}=-\nabla\phi-\partial_{t}\mathbf{A}.

Define two auxiliary vector fields, one is the unit vector along the direction of the magnetic field

𝐛𝐁|B|,\mathbf{b}\equiv\frac{\mathbf{B}}{\left|B\right|}\;, (44)

and the other is the 𝐄×𝐁\mathbf{E}\times\mathbf{B} drift velocity

𝐃𝐄×𝐁𝐁2.\mathbf{D}\equiv\frac{\mathbf{E}\times\mathbf{B}}{\mathbf{B}^{2}}\;. (45)

The velocity 𝐯\mathbf{v} is decomposed into horizontal part 𝐯\mathbf{v}_{\parallel} and vertical part 𝐯\mathbf{v}_{\perp}

𝐯\displaystyle\mathbf{v}_{\parallel} 𝐯𝐛𝐛+𝐃,\displaystyle\equiv\mathbf{v}\cdot\mathbf{b}\mathbf{b}+\mathbf{D}\;, (46)
𝐯\displaystyle\mathbf{v}_{\perp} 𝐯𝐯𝐛𝐛𝐃,\displaystyle\equiv\mathbf{v}-\mathbf{v}\cdot\mathbf{b}\mathbf{b}-\mathbf{D}\;, (47)

If 𝐃0\mathbf{D}\neq 0, 𝐯\mathbf{v}_{\perp} and 𝐯\mathbf{v}_{\parallel} are not orthogonal

𝐯2+𝐯2=𝐯22𝐯𝐃.\mathbf{v}_{\parallel}^{2}+\mathbf{v}_{\perp}^{2}=\mathbf{v}^{2}-2\mathbf{v}_{\perp}\cdot\mathbf{D}\;. (48)

The decomposition of the Hamilitonian vector τ=τ+τ\mathbf{\tau}=\tau_{\parallel}+\tau_{\perp} looks like

τ\displaystyle\tau_{\parallel} =𝐯𝐱+𝐄𝐛𝐛𝐯+t,\displaystyle=\mathbf{v}_{\parallel}\cdot\partial_{\mathbf{x}}+\mathbf{E}\cdot\mathbf{b}\mathbf{b}\cdot\partial_{\mathbf{v}}+\partial_{t}\;, (49)
τ\displaystyle\tau_{\perp} =𝐯𝐱+𝐯×𝐁𝐯.\displaystyle=\mathbf{v}_{\perp}\cdot\partial_{\mathbf{x}}+\mathbf{v}_{\perp}\times\mathbf{B}\cdot\partial_{\mathbf{v}}\;. (50)

The gyro-radius in configuration space is a three-dimensional vector field

𝝆𝐛×𝐯|B|.\boldsymbol{\mathbf{\rho}}\equiv\frac{\mathbf{b}\times\mathbf{v}_{\perp}}{\left|B\right|}\;. (51)

Using 𝝆\boldsymbol{\rho}, we decompose the the four-momentum pp into vertical part

p\displaystyle p_{\perp} ι𝝆𝐱dA=𝝆×𝐁d𝐱𝐄𝝆dt\displaystyle\equiv-\iota_{\boldsymbol{\rho}\cdot\partial_{\mathbf{x}}}dA=\boldsymbol{\rho}\times\mathbf{B}\cdot d\mathbf{x}-\mathbf{E}\cdot\boldsymbol{\mathbf{\rho}}dt
=(𝐯𝐯𝐛𝐛𝐃)d𝐱(𝐯𝐃)𝐃dt\displaystyle=\left(\mathbf{v}-\mathbf{v}\cdot\mathbf{b}\mathbf{b}-\mathbf{D}\right)\cdot d\mathbf{x}-\left(\mathbf{v}-\mathbf{D}\right)\cdot\mathbf{D}dt
=𝐯(d𝐱𝐃dt),\displaystyle=\mathbf{v}_{\perp}\cdot\left(d\mathbf{x}-\mathbf{D}dt\right)\;, (52)

and horizontal part

p\displaystyle p_{\parallel} pp\displaystyle\equiv p-p_{\perp}
=(𝐯𝐛𝐛+𝐃)d𝐱(𝐯𝐃)2+𝐃22dt\displaystyle=\left(\mathbf{v}\cdot\mathbf{b}\mathbf{b}+\mathbf{D}\right)\cdot d\mathbf{x}-\frac{\left(\mathbf{v}-\mathbf{D}\right)^{2}+\mathbf{D}^{2}}{2}dt
=𝐯d𝐗𝐯2+𝐯22dt\displaystyle=\mathbf{v_{\parallel}}\cdot d\mathbf{X}-\frac{\mathbf{v}_{\parallel}^{2}+\mathbf{v}_{\perp}^{2}}{2}dt
=𝐯d𝐗𝐯22dtμ|B|dt.\displaystyle=\mathbf{v_{\parallel}}\cdot d\mathbf{X}-\frac{\mathbf{v_{\parallel}}^{2}}{2}dt-\mu\left|B\right|dt\,. (53)

Appending a closed form d(𝝆𝐀)-d\left(\boldsymbol{\rho}\cdot\mathbf{A}\right) to the Poincaré-Cartan-Einstein one-form Eq.(41), yields

η\displaystyle\eta =A+p𝝆𝐱A,\displaystyle=A+p_{\parallel}-\mathcal{L}_{\boldsymbol{\rho}\cdot\partial_{\mathbf{x}}}A\,, (54)

which will simplify our subsequent derivation.

IV.2 Gyro-transformation

First consider the case that the electromagnetic field is homogeneous, ε=0\varepsilon=0. The gyro-transformation is a rotation of the gyro-radius 𝝆\boldsymbol{\rho}

Φθ(z)=(t,𝐱+𝝆(cosθ1)+𝝆×𝐛sinθ,𝐯+|B|𝝆×𝐛(cosθ1)|B|𝝆sinθ),\Phi_{\theta}\left(z\right)=\left(t,\mathbf{x}+\boldsymbol{\rho}\left(\cos\theta-1\right)+\boldsymbol{\rho}\times\mathbf{b}\sin\theta,\mathbf{v}+\left|B\right|\boldsymbol{\rho}\times\mathbf{b}\left(\cos\theta-1\right)-\left|B\right|\boldsymbol{\rho}\sin\theta\right)\,, (55)

whose generator is obtained from vertical Hamiltonian vector field τ\tau_{\perp} (Eq.(50))

θddsΦθs|s=0=𝝆×𝐛𝐱|B|𝝆𝐯=τ|B|.\partial_{\theta}\equiv\frac{d}{ds}\left.\Phi_{\theta s}\right|_{s=0}=\boldsymbol{\rho}\times\mathbf{b}\cdot\partial_{\mathbf{x}}-\left|B\right|\boldsymbol{\rho}\cdot\partial_{\mathbf{v}}=\frac{\tau_{\perp}}{\left|B\right|}\;. (56)

The action integral along the Kruskal’s ring 𝒪\mathcal{O} is

μP[𝒪]\displaystyle\mu_{P}\left[\mathcal{O}\right] =12π𝒪(A+p)\displaystyle=\frac{1}{2\pi}\oint_{\mathcal{O}}\left(A+p\right)
=12πD𝒪𝐁𝑑𝐒+12π02πΦθ(ιθp)𝑑θ\displaystyle=-\frac{1}{2\pi}\int_{D_{\mathcal{O}}}\mathbf{B}\cdot d\mathbf{S}+\frac{1}{2\pi}\int_{0}^{2\pi}\Phi_{\theta}^{*}\left(\iota_{\partial_{\theta}}p\right)d\theta
=|B|ρ22+𝐯2|B|=|B|ρ22=𝐯22|B|,\displaystyle=-\frac{\left|B\right|\rho^{2}}{2}+\frac{\mathbf{v}_{\perp}^{2}}{\left|B\right|}=\frac{\left|B\right|\rho^{2}}{2}=\frac{\mathbf{v}_{\perp}^{2}}{2\left|B\right|}\;, (57)

where D𝒪D_{\mathcal{O}} is the area enclosed by 𝒪\mathcal{O}, and the Stokes’ theorem is applied. The Eq.(57) is the original definition of the magnetic moment [24].

The calculation of the moment map μ=ιθη\mu=\iota_{\partial_{\theta}}\eta is tricky because the potential A(t,𝐱)A\left(t,\mathbf{x}\right) of uniform electromagnetic field are not constant. The traditional solution is to expand η\eta around the guiding center Z=(t,𝐗,𝐕)Z=\left(t,\mathbf{X},\mathbf{V}\right) with respect the gyro-radius 𝝆\boldsymbol{\rho}[9]. Expanding Eq. (54) with respect 𝝆𝐱\boldsymbol{\rho}\cdot\partial_{\mathbf{x}}, yields

η¯=η|X\displaystyle\bar{\eta}=\left.\eta\right|_{X} =p+A+𝝆𝐱A+ε12𝝆𝐱2A\displaystyle=p_{\parallel}+A+\mathcal{L}_{\boldsymbol{\rho}\cdot\partial_{\mathbf{x}}}A+\varepsilon\frac{1}{2}\mathcal{L}_{\boldsymbol{\rho}\cdot\partial_{\mathbf{x}}}^{2}A
𝝆𝐱Aε𝝆𝐱2A+O(ε2)\displaystyle-\mathcal{L}_{\boldsymbol{\rho}\cdot\partial_{\mathbf{x}}}A-\varepsilon\mathcal{L}_{\boldsymbol{\rho}\cdot\partial_{\mathbf{x}}}^{2}A+O\left(\varepsilon^{2}\right)
=A+p12𝝆d𝐕+O(ε).\displaystyle=A+p_{\parallel}-\frac{1}{2}\boldsymbol{\rho}\cdot d\mathbf{V}+O\left(\varepsilon\right)\,. (58)

Pushing θ\partial_{\theta} forward to the guiding center, yields the limiting rote-rate vector θ¯\partial_{\bar{\theta}} as follows

θ¯=Π(𝝆𝐱)θ=|B|𝝆𝐕+O(ε).\partial_{\bar{\theta}}=\Pi_{\left(-\boldsymbol{\rho}\cdot\partial_{\mathbf{x}}\right)*}\partial_{\theta}=-\left|B\right|\boldsymbol{\rho}\cdot\partial_{\mathbf{V}}+O\left(\varepsilon\right)\,. (59)

It is straightforward to verify that the Lagrangian one-form η¯\bar{\eta} (Eq. (58)) is gyro-invariant in a uniform electromagnetic field

θ¯η¯\displaystyle\mathcal{L}_{\partial_{\bar{\theta}}}\bar{\eta} =0,\displaystyle=0\,, (60)
ιθ¯dη¯\displaystyle\iota_{\partial_{\bar{\theta}}}d\bar{\eta} =dιθ¯η¯=dμ¯.\displaystyle=-d\iota_{\partial_{\bar{\theta}}}\bar{\eta}=-d\bar{\mu}\,. (61)

and the magnetic moment μ\mu is given by

μ¯=ιθ¯η¯=|B|ρ22,\bar{\mu}=\iota_{\partial_{\bar{\theta}}}\bar{\eta}=\frac{\left|B\right|\rho^{2}}{2}\;, (62)

which is equal to the action integral along the Kruskal’s ring μP[𝒪]\mu_{P}\left[\mathcal{O}\right].

IV.3 Gyro-average

For the inhomogeneous case ε0\varepsilon\neq 0, we need pull the Lagrangian one-form η\eta back to the guiding center. Substituting Eq.(54) to Eq.(13), yields

ηε\displaystyle\eta_{\varepsilon} =n0εnn!ρn(1εA+pρA)=A+p+n>0εnn!ηn,\displaystyle=\sum_{n\geq 0}\frac{\varepsilon^{n}}{n!}\mathcal{L}_{\partial_{\rho}}^{n}\left(\frac{1}{\varepsilon}A+p_{\parallel}-\mathcal{L}_{\partial_{\rho}}A\right)=A+p_{\parallel}+\sum_{n>0}\frac{\varepsilon^{n}}{n!}\eta_{n}\,, (63)

where

ηn\displaystyle\eta_{n} ρnpnn+1ρn+1A.\displaystyle\equiv\mathcal{L}_{\partial_{\rho}}^{n}p_{\parallel}-\frac{n}{n+1}\mathcal{L}_{\partial_{\rho}}^{n+1}A\;. (64)

The head order does not depend on the gyro-radius ρ\partial_{\rho}

η=A+p,\eta_{\parallel}=A+p_{\parallel}\,, (65)

which is the horizontal Lagrangian one-form. The other orders are polynomials of Lie derivative ρ\mathcal{L}_{\partial_{\rho}}

η=n>0εnn!ηn,\eta_{\perp}=\sum_{n>0}\frac{\varepsilon^{n}}{n!}\eta_{n}\,, (66)

which form the vertical Lagrangian one-form.

Since pp_{\parallel} and A(t,𝐗)A\left(t,\mathbf{X}\right) are gyro-independent, the gyro-transformation of ηε\eta_{\varepsilon} only depend on the rotation of gyro-radius ρ¯\partial_{\bar{\rho}}

Φθ¯sηn\displaystyle\Phi_{\bar{\theta}s}^{*}\eta_{n} =Φθ¯sρ¯npnn+1Φθ¯sρ¯n+1A.\displaystyle=\mathcal{L}_{\Phi_{\bar{\theta}s}^{*}\partial_{\bar{\rho}}}^{n}p_{\parallel}-\frac{n}{n+1}\mathcal{L}_{\Phi_{\bar{\theta}s}^{*}\partial_{\bar{\rho}}}^{n+1}A\,. (67)

Let the gyro-radius ρ¯\partial_{\bar{\rho}} satisfy the complex-like condition Eq.(37), then the gyro-averaged Lagrangian one-form ηn\left\langle\eta_{n}\right\rangle are polynomials of the Lie derivative ρ¯\mathcal{L}_{\partial_{\bar{\rho}}} (see Eq.(39)),

ηn\displaystyle\left\langle\eta_{n}\right\rangle =01𝑑s(cos(2πs)ρ¯+sin(2πs)ρˇ)np\displaystyle=\int_{0}^{1}ds\left(\cos\left(2\pi s\right)\mathcal{L}_{\partial_{\bar{\rho}}}+\sin\left(2\pi s\right)\mathcal{L}_{\partial_{\check{\rho}}}\right)^{n}p_{\parallel}
nn+101𝑑s(cos(2πs)ρ¯+sin(2πs)ρˇ)n+1A.\displaystyle-\frac{n}{n+1}\int_{0}^{1}ds\left(\cos\left(2\pi s\right)\mathcal{L}_{\partial_{\bar{\rho}}}+\sin\left(2\pi s\right)\mathcal{L}_{\partial_{\check{\rho}}}\right)^{n+1}A\,. (68)

Let the rote-rate vector be

θ¯|B|𝝆𝐕.\partial_{\bar{\theta}}\equiv-\left|B\right|\boldsymbol{\rho}\cdot\partial_{\mathbf{V}}\;. (69)

The gyro-radius is obtaineded from the complex-like condition Eq.(37)

ρ¯\displaystyle\partial_{\bar{\rho}} 𝝆𝐗+ε(𝝆𝐛×𝐛×(𝐕𝐃)+𝝆𝐃)𝐕,\displaystyle\equiv\boldsymbol{\rho}\cdot\partial_{\mathbf{X}}+\varepsilon\left(-\boldsymbol{\rho}\cdot\nabla\mathbf{b}\mathbf{\times}\mathbf{b}\mathbf{\times}\left(\mathbf{V}-\mathbf{D}\right)+\boldsymbol{\rho}\cdot\nabla\mathbf{D}\right)\cdot\partial_{\mathbf{V}}\,, (70)

whose second term come from the perturbation caused by the inhomogeneity. There are different equivalent forms of ρ¯\partial_{\bar{\rho}} and θ¯\partial_{\bar{\theta}} that correspond to different decomposition of gyro-motion [12]. How to find a suitable decomposition for the gyro-motion is a problem worthy of in-depth discussion.

IV.4 The guiding-center dynamic

In order to obtain the dynamics of the guiding center, we first perform the gyro-average on the Lagrangian one-form η¯=ηε\bar{\eta}=\left\langle\eta_{\varepsilon}\right\rangle. The vertical part is given by Eq.(68),

η¯\displaystyle\bar{\eta}_{\perp} =η=14ε(ρ¯2A+ρˇ2A)+O(ε2).\displaystyle=\left\langle\eta_{\perp}\right\rangle=-\frac{1}{4}\varepsilon\left(\mathcal{L}_{\partial_{\bar{\rho}}}^{2}A+\mathcal{L}_{\partial_{\check{\rho}}}^{2}A\right)+O\left(\varepsilon^{2}\right)\,. (71)

Substituting Eq.(70), yields

η¯\displaystyle\bar{\eta}_{\perp} =12𝝆d𝐕+μ¯2(𝐛×𝐛)𝐛d𝐗+μ¯2(𝐛×𝐃)dt,\displaystyle=-\frac{1}{2}\boldsymbol{\rho}\cdot d\mathbf{V}_{\perp}+\frac{\bar{\mu}}{2}\left(\mathbf{b}\cdot\nabla\times\mathbf{b}\right)\mathbf{b}\cdot d\mathbf{X}+\frac{\bar{\mu}}{2}\left(\mathbf{b}\cdot\nabla\times\mathbf{D}\right)dt\,, (72)

where

μ¯\displaystyle\bar{\mu} ιθ¯η¯=ιθ¯η¯=𝐕22|B|,\displaystyle\equiv\iota_{\partial_{\bar{\theta}}}\bar{\eta}=\iota_{\partial_{\bar{\theta}}}\bar{\eta}_{\perp}=\frac{\mathbf{V}_{\perp}^{2}}{2\left|B\right|}\;, (73)

is the magnetic moment.

Extracting the factor μ\mu from η¯\bar{\eta}_{\perp}, yields a dimensionless one-form

ημ=σ\displaystyle\frac{\eta_{\perp}}{\mu}=\sigma d𝐜𝐚+12(𝐛×𝐛)𝐛d𝐗12𝐛×𝐃dt,\displaystyle\equiv-d\mathbf{c}\cdot\mathbf{a}+\frac{1}{2}\left(\mathbf{b}\cdot\nabla\times\mathbf{b}\right)\mathbf{b}\cdot d\mathbf{X}-\frac{1}{2}\mathbf{b}\cdot\nabla\times\mathbf{D}dt\,, (74)

where

𝐜\displaystyle\mathbf{c} 𝐕|𝐕|=𝐚×𝐛,\displaystyle\equiv\frac{\mathbf{V}_{\perp}}{\left|\mathbf{V}_{\perp}\right|}=\mathbf{a}\times\mathbf{b}\,, (75)
𝐚\displaystyle\mathbf{a} θ¯𝐜=𝐛×𝐜=𝐛×𝐕|𝐕|,\displaystyle\equiv-\mathcal{L}_{\partial_{\bar{\theta}}}\mathbf{c}=\mathbf{b}\times\mathbf{c}=\frac{\mathbf{b}\times\mathbf{V}_{\perp}}{\left|\mathbf{V}_{\perp}\right|}\,, (76)

are two unit vector fields perpendicular to the direction of magnetic field 𝐛\mathbf{b}. The dimensionless one-form σ\sigma only depends on the spatial-temporal inhomogeneity of the electromagnetic field. The interior product of θ¯\partial_{\bar{\theta}} and σ\sigma is unit one, ιθ¯σ=1\iota_{\partial_{\bar{\theta}}}\sigma=1. A natural question arises. Is σ\sigma the covector of gyrophase dθd\theta? Or, can σ\sigma define the global gyrophase θ\theta? To answer this question, we check the exterior derivative of σ\sigma

dσ=\displaystyle d\sigma= 12×𝐑×d𝐗d𝐗(R+t𝐑)d𝐗dt,\displaystyle\frac{1}{2}\nabla\times\mathbf{R}\times d\mathbf{X}\wedge d\mathbf{X}-\left(\nabla R+\partial_{t}\mathbf{R}\right)\cdot d\mathbf{X}\wedge dt\>, (77)

where

𝐑\displaystyle\mathbf{R} 𝐜𝐚12(𝐛×𝐛)𝐛,\displaystyle\equiv\nabla\mathbf{c}\cdot\mathbf{a}-\frac{1}{2}\left(\mathbf{b}\cdot\nabla\times\mathbf{b}\right)\mathbf{b}\,, (78)
R\displaystyle R t𝐜𝐚12𝐛×𝐃,\displaystyle\equiv-\partial_{t}\mathbf{c}\cdot\mathbf{a}-\frac{1}{2}\mathbf{b}\cdot\nabla\times\mathbf{D}\,, (79)

is a four-vector (𝐑,R)\left(\mathbf{R},R\right) in the configuration space. Since dσ0d\sigma\neq 0, the dimensionless one-form σ\sigma is not an exact form σdθ\sigma\neq d\theta, and we cannot define the global gyrophase θ\theta from σ\sigma.

The traditional guiding center theory defines the gyrophase on a predefined local orthogonal coordinate frame, which may not exist globally over a non-trivial field topology. This issue was raised and discussed by Sugiyama[3, 2] and Krommes[4] in 2009. Soon after, Burdy and Qin[5] identified the obstruction to the global existence of gyrophase is the vector field ×(𝐜𝐚)\nabla\times\left(\nabla\mathbf{c}\cdot\mathbf{a}\right). Bohosian [25] gave similar results in his earlier work. However, the global gyro-phase is not a necessary condition for gyro-symmetry. The gyro-symmetry depends only on the homogeneity of the phase space within the range of Kruskal’s ring. The classical approach obscures the geometric meaning of the gyro-symmetry. A coordinate-free geometric representation is a more suitable alternative. The Eq.(77) shows that the existence of gyrophase depends on the integrability of σ\sigma, or requires two-form dσ=0d\sigma=0 to vanish everywhere. This is a straightforward conclusion of the geometric method.

Continuing the derivation of guiding center dynamics, the horizontal Lagrangian one-form is gyro-independent

η¯=η=A+p=(𝐀+𝐕)d𝐗(ϕ+𝐕22μ|B|)dt,\bar{\eta}_{\parallel}=\left\langle\eta_{\parallel}\right\rangle=A+p_{\parallel}=\left(\mathbf{A}+\mathbf{V_{\parallel}}\right)\cdot d\mathbf{X}-\left(\phi+\frac{\mathbf{V_{\parallel}}^{2}}{2}-\mu\left|B\right|\right)dt\,, (80)

which no additional calculations are required. Using Eq.(77), the Lagrangian two-form is rewritten as

dη¯=dη¯+dη¯=Ω+dμσ,d\bar{\eta}=d\bar{\eta}_{\parallel}+d\bar{\eta}_{\perp}=\Omega^{\dagger}+d\mu\wedge\sigma\,, (81)

where

Ω\displaystyle\Omega^{\dagger} dη+μdσ\displaystyle\equiv d\eta_{\parallel}+\mu d\sigma
=12𝐁×d𝐗d𝐗+𝐄d𝐗dt\displaystyle=\frac{1}{2}\mathbf{B}^{\dagger}\times d\mathbf{X}\wedge d\mathbf{X}+\mathbf{E}^{\dagger}\cdot d\mathbf{X}\wedge dt
+𝐛d𝐕𝐛d𝐗(𝐕𝐃)d𝐕dt,\displaystyle+\mathbf{b}\cdot d\mathbf{V}\wedge\mathbf{b}\cdot d\mathbf{X}-\left(\mathbf{V}-\mathbf{D}\right)\cdot d\mathbf{V}\wedge dt\,, (82)

is the gyro-independent part of Lagrangian two-form, and

𝐁\displaystyle\mathbf{B}^{\dagger} 𝐁+×𝐕+μ×𝐑,\displaystyle\equiv\mathbf{B}+\nabla\times\mathbf{V_{\parallel}}+\mu\nabla\times\mathbf{R}\,, (83)
𝐄\displaystyle\mathbf{E}^{\dagger} 𝐄t𝐕(𝐕22+μ|B|)μ(R+t𝐑),\displaystyle\equiv\mathbf{E}-\partial_{t}\mathbf{V_{\parallel}}-\nabla\left(\frac{\mathbf{V_{\parallel}}^{2}}{2}+\mu\left|B\right|\right)-\mu\left(\nabla R+\partial_{t}\mathbf{R}\right)\;, (84)

are effective electromagnetic fields. Then, the Eq.(17) is simplified as

τ=3!2!dμσΩΩσΩΩΩ=dμσΩΩ2𝐁𝐛.*\tau_{\parallel}^{\flat}=\frac{3!}{2!}\frac{d\mu\wedge\sigma\wedge\Omega^{\dagger}\wedge\Omega^{\dagger}}{\sigma\wedge\Omega^{\dagger}\wedge\Omega^{\dagger}\wedge\Omega^{\dagger}}=\frac{d\mu\wedge\sigma\wedge\Omega^{\dagger}\wedge\Omega^{\dagger}}{2\mathbf{B}^{\dagger}\cdot\mathbf{b}}\,. (85)

Substituting Eq. (82), yields the Hamiltonian vector field of guiding center

τt\displaystyle\tau_{\parallel t} =𝐁𝐛|B|,\displaystyle=\frac{\mathbf{B}^{\dagger}\cdot\mathbf{b}}{\left|B\right|}\,, (86)
τ𝐗\displaystyle\tau_{\parallel\mathbf{X}} =𝐕𝐛𝐁+𝐄×𝐛|B|𝐛×μ,\displaystyle=\frac{\mathbf{V}\cdot\mathbf{b}\mathbf{B}^{\dagger}+\mathbf{E}^{\dagger}\times\mathbf{b}}{\left|B\right|}-\mathbf{b}\times\nabla\mu\,, (87)
τ𝐕\displaystyle\tau_{\parallel\mathbf{V}} =(𝐁𝐄|B|+𝐁μ)𝐛\displaystyle=\left(\frac{\mathbf{B}^{\dagger}\cdot\mathbf{E}^{\dagger}}{\left|B\right|}+\mathbf{B}^{\dagger}\cdot\nabla\mu\right)\mathbf{b}
+|𝐕|(τ𝐗𝐑τtR)𝐚\displaystyle+\left|\mathbf{V}_{\perp}\right|\left(\tau_{\parallel\mathbf{X}}\cdot\mathbf{R}-\tau_{\parallel t}R\right)\mathbf{a}
12|𝐕|(τ𝐗lnμ+τttlnμ)𝐜.\displaystyle-\frac{1}{2}\left|\mathbf{V}_{\perp}\right|\left(\tau_{\parallel\mathbf{X}}\cdot\nabla\ln\mu+\tau_{\parallel t}\partial_{t}\ln\mu\right)\mathbf{c}\,. (88)

where τ𝐗\tau_{\mathbf{X}} is the drift motion of the guide center, and τ𝐕\tau_{\mathbf{V}} is the acceleration of Kruskal’s ring. In Eq. (88), the first term is the acceleration of the guiding center in the direction of the magnetic field, the second term represents the expansion of Kruskal’s ring, and the third term represents the rotational acceleration of Kruskal’s ring. One can simply verify that τ\tau_{\parallel} is gyro-independent and orthogonal to the gyro-motion θ¯τ=ιτη=0\mathcal{L}_{\partial_{\bar{\theta}}}\tau_{\parallel}=\iota_{\tau_{\parallel}}\eta_{\perp}=0. The magnetic moment μ\mu is a constant for guiding center motion ιτdμ=0\iota_{\tau_{\parallel}}d\mu=0.

V Summary and discussion

To summarize, we discuss the dynamics of a charged particle in a time-dependent, slowly varying electromagnetic field, whose Lagrangian is a one-form η\eta on the seven-dimensional contact manifold PP. The orbit of gyro-symmetry Φθ\Phi_{\theta} is a closed ring in the phase space, called Kruskal’s ring 𝒪\mathcal{O}. The guiding center is the center of Kruskal’s ring, the fixed point of gyro-symmetry Φθ\Phi_{\theta}. By properly defining the rote-rate vector θ¯\partial_{\bar{\theta}} (Eq.(69)) and the gyro-radius ρ¯\partial_{\bar{\rho}} (Eq.(70)), we give a general expression for the gyro-averaging (Eq. (68) ). Further, the geometric decomposition of the gyro-motion is given and verified (Eq. (72) and (80)). As a result, we obtain the coordinate-free expression of the non-relativistic guiding-center dynamics in the time-dependent slow-varying electromagnetic field. (Eq. (87) and (88)).

We understand the gyro-symmetry as the similarity of the trajectory of ringmates on the same Kruskal’s ring, which depends only on the local homogeneity of the electromagnetic field. The gyro-averaging is integral along the Kruskal’s ring. The guiding center is a particle with "spin", whose magnetic moment μ\mu is equal to the action integral over the Kruskal’s ring. The guiding center dynamics is the dynamics of Kruskal’s ring. The purpose of guiding center theory is to decompose particle motion into vertical gyro-motion and horizontal drift motion. In the classical guiding center theory, the vertical part of Lagrangian is expressed as μdθ\mu d\theta. We now recognize that dθd\theta may not be globally defined on the non-trivial magnetic field. Even if we ignore the global validity and consider only the local domain, dθd\theta is still an ambiguous expression. Appling the local gyro-phase coordinate frame to the dimensionless one-form ϑ\vartheta, yields ϑ=dθ+d𝐗𝐜𝐚+dtt𝐜𝐚+𝒪(ε)\vartheta=d\theta+d\mathbf{X}\cdot\nabla\mathbf{c}\cdot\mathbf{a}+dt\cdot\partial_{t}\mathbf{c}\cdot\mathbf{a}+\mathcal{O}\left(\varepsilon\right), which means dθd\theta only make scenes when the electromagnetic field is homogeneous. The geometric decomposition method can avoid the confusion caused by the local gyro-phase coordinates frame. The expansion (Eq.(63)) and averaging method (Eq.(68)) ensure that each order of the expansion is gyro-independent.

In the subsequent work, the geometric methods established in this paper will be applied to Lie perturbations, gyro-kinetics theory, and others. Thus, the rich geometric of gyro-dynamics should be further revealed.

Acknowledgements.
The author sincerely thanks Prof. Chang-Xuan Yu, Prof. Wan-Dong Liu, Prof. Jin-Lin Xie, and the unmentioned supervisors. The author is grateful for their support during his early academic career. The author also would like to thank Dr. Jian-Yuan Xiao and Dr. Pei-Feng Fan for the fruitful discussions on geometry and numerical algorithm. This work was supported by the National MCF Energy R&D Program under Contract No.2018YFE0304102.

References

  • [1] John R. Cary and Alain J. Brizard. Hamiltonian theory of guiding-center motion. Reviews of Modern Physics, 81(2):693–738, May 2009.
  • [2] Linda E. LE Sugiyama. Response to Comment on "Guiding center plasma models in three dimensions". Physics of Plasmas, 16(8):1–4, August 2009.
  • [3] Linda E. Sugiyama. Guiding center plasma models in three dimensions. Physics of Plasmas, 15(9):1–13, September 2008.
  • [4] John A. Krommes. Comment on "Guiding center plasma models in three dimensions" [Phys. Plasmas 15, 092112 (2008)]. Physics of Plasmas, 16(8):084701, August 2009.
  • [5] J. W. Burby and H. Qin. Gyrosymmetry: Global considerations. Physics of Plasmas, 19(5):052106, 2012.
  • [6] J. W. Burby. Guiding center dynamics as motion on a formal slow manifold in loop space. Journal of Mathematical Physics, 61(1):012703, January 2020.
  • [7] J. W. Burby and J. Squire. General formulas for adiabatic invariants in nearly periodic Hamiltonian systems. Journal of Plasma Physics, 86(6):835860601, December 2020.
  • [8] J. W. Burby and E. Hirvijoki. Normal stability of slow manifolds in nearly periodic Hamiltonian systems. Journal of Mathematical Physics, 62(9):093506, September 2021.
  • [9] Martin Kruskal. Asymptotic theory of Hamiltonian and other systems with all solutions nearly periodic. Journal of Mathematical Physics, 3(4):806, July 1962.
  • [10] H Qin. A short introduction to general gyrokinetic theory. In Topics in kinetic theory. PPPL, 2005.
  • [11] S. M. Omohundro. Geometric perturbation theory and plasma physics. Thesis/Dissertation, Lawrence Berkeley Lab., CA (USA), April 1985.
  • [12] F. I. Parra, I. Calvo, J. W. Burby, J. Squire, and H. Qin. Equivalence of two independent calculations of the higher order guiding center Lagrangian. Physics of Plasmas, 21(10):104506, October 2014.
  • [13] V. I. Arnold. Appendix 4 Contact structures. In Mathematical Methods of Classical Mechanics, volume 60 of Graduate Texts in Mathematics. Springer New York, New York, NY, 1989.
  • [14] Jerrold E Marsden and Tudor S Ratiu. Introduction to Mechanics and Symmetry, volume 17 of Texts in Applied Mathematics. Springer New York, New York, NY, 1999.
  • [15] V. I. Arnold. 44 The integral invariang of Poincare-Cartan. In Mathematical Methods of Classical Mechanics, volume 60 of Graduate Texts in Mathematics. Springer New York, New York, NY, 1989.
  • [16] Brian C. Hall. Lie Groups, Lie Algebras, and Representations: An Elementary Introduction, volume 222 of Graduate Texts in Mathematics. Springer International Publishing, Cham, 2015.
  • [17] Vladimir Igorevich Arnol’d. Theory of perturbations of conditionally periodic motion. In Mathematical methods of classical mechanics, page 405. Springer-Verlag, 1989.
  • [18] Mari{\’a}n Fecko. Differential Geometry and Lie Groups for Physicists. Cambridge University Press, Cambridge, 2006.
  • [19] Robert G. Littlejohn. Hamiltonian perturbation theory in noncanonical coordinates. Journal of Mathematical Physics, 1982.
  • [20] Robert G. Littlejohn. A guiding center Hamiltonian: A new approach. Journal of Mathematical Physics, 1979.
  • [21] John R Cary and Robert G Littlejohn. Noncanonical Hamiltonian mechanics and its application to magnetic field line flow. Annals of Physics, 151(1):1–34, November 1983.
  • [22] J Cary. Lie transform perturbation theory for Hamiltonian systems. Physics Reports, 79(2):129–159, December 1981.
  • [23] H Qin and W M Tang. Pullback transformations in gyrokinetic theory. Physics of Plasmas, 11(3), 2004.
  • [24] John David Jackson. Classical electrodynamics. Wiley, New York, 3rd ed edition, 1999.
  • [25] BM M Boghosian. Covariant Lagrangian methods of relativistic plasma theory. PhD thesis, UNIVERSITY OF CALIFORNIA DAVIS, 1987.