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

Quasisymmetric Magnetic Fields in Asymmetric Toroidal Domains

Naoki Sato Graduate School of Frontier Sciences,
The University of Tokyo, Kashiwa, Chiba 277-8561, Japan
Email: [email protected]
Zhisong Qu Mathematical Sciences Institute, The Australian National University, Canberra, ACT 2601, Australia David Pfefferlé The University of Western Australia, 35 Stirling Highway, Crawley WA 6009, Australia Robert L. Dewar Mathematical Sciences Institute, The Australian National University, Canberra, ACT 2601, Australia
Abstract

We explore the existence of quasisymmetric magnetic fields in asymmetric toroidal domains. These vector fields can be identified with a class of magnetohydrodynamic equilibria in the presence of pressure anisotropy. First, using Clebsch potentials, we derive a system of two coupled nonlinear first order partial differential equations expressing a family of quasisymmetric magnetic fields in bounded domains. In regions where flux surfaces and surfaces of constant field strength are not tangential, this system can be further reduced to a single degenerate nonlinear second order partial differential equation with externally assigned initial data. Then, we exhibit regular quasisymmetric vector fields which correspond to local solutions of anisotropic magnetohydrodynamics in asymmetric toroidal domains such that tangential boundary conditions are fulfilled on a portion of the bounding surface. The problems of boundary shape and locality are also discussed. We find that symmetric magnetic fields can be fitted into asymmetric domains, and that the mathematical difficulty encountered in the derivation of global quasisymmetric magnetic fields lies in the topological obstruction toward global extension affecting local solutions of the governing nonlinear first order partial differential equations.

1 Introduction

In the context of ideal magnetohydrodynamics with isotropic pressure, steady magnetic confinement of a plasma is achieved by balance between Lorentz force and pressure gradient,

(×𝑩)×𝑩=P,𝑩=0inΩ,\displaystyle\left({\nabla\times\boldsymbol{B}}\right)\times\boldsymbol{B}=\nabla P,~{}~{}~{}~{}\nabla\cdot\boldsymbol{B}=0~{}~{}~{}~{}{\rm in}~{}~{}\Omega, (1a)
𝑩𝒏=0onΩ.\displaystyle\boldsymbol{B}\cdot\boldsymbol{n}=0~{}~{}~{}~{}{\rm on}~{}~{}\partial\Omega. (1b)

In the notation above, 𝑩\boldsymbol{B} is the magnetic field, PP the pressure, Ω3\Omega\subset\mathbb{R}^{3} a smooth bounded domain, and 𝒏\boldsymbol{n} the unit outward normal to the bounding surface Ω\partial\Omega. As a system of nonlinear first order partial differential equations for the Cartesian components BxB_{x}, ByB_{y}, BzB_{z} and the pressure PP, equation (1) is twice elliptic and twice hyperbolic [1]. Such mixed behavior makes (1) one of the hardest equations in mathematical physics. Indeed, while in the presence of a continuous Euclidean symmetry the hyperbolic part can be removed to obtain an elliptic nonlinear second order partial differential equation for the flux function (the Grad-Shafranov equation [2, 3, 4]), the existence of regular solutions of (1) without such isometries remains an unsolved problem in the theory of partial differential equations [5].

In the following, we shall always use the word symmetry to refer to continuous transformations of three dimensional Euclidean space that preserve the Euclidean distance between points, i.e. superpositions of translations and rotations. Asymmetry will then imply absence of such Euclidean isometries. Furthermore, a certain equilibrium will be deemed symmetric as long as the magnetic field is, without any requirements on other fields or boundary shape. It should be noted that a symmetric solution (𝑩,P)\left({\boldsymbol{B},P}\right) of (1) does not necessarily imply a symmetric boundary Ω\partial\Omega. For example, a symmetric magnetic field satisfying (1) can be embedded within an asymmetric bounded domain. A symmetric solution always implies a symmetric boundary when the boundary corresponds to an isobaric surface. On these points, see [6]. According to a conjecture due to H. Grad, well-behaved solutions of (1) with non-constant pressure in Ω\Omega and constant pressure on Ω\partial\Omega should possess a high degree of symmetry [7]. Intuitively, this is because the boundary condition for the pressure effectively forces magnetic field 𝑩\boldsymbol{B} and current ×𝑩\nabla\times\boldsymbol{B} on toroidal surfaces (hairy ball theorem [8]). In the absence of a boundary symmetry, this topological obstruction cannot be trivially overcome when one tries to lie 𝑩\boldsymbol{B} and ×𝑩\nabla\times\boldsymbol{B} consistently around the torus. A workaround for this problem consists in expanding the class of admissible solutions so that controlled discontinuities are allowed. In this construction, the domain Ω\Omega is partitioned into a given number of subdomains. Within each subdomain a constant pressure is assumed, and the corresponding magnetic field is a Beltrami field (an eigenstate of the curl operator, see [9, 10, 11] for the existence of such solutions), while pressure jumps occur at the boundaries among subdomains. This approach has proven effective for the modeling of three-dimensional magnetohydrodynamic equilibria aimed at stellarator design [12, 13, 14] (stellarators are candidate magnetic confinement devices for nuclear fusion applications that do not exhibit boundary symmetry).

Compared to an axially symmetric tokamak, a stellarator offers the advantage that the field line twist required to trap charged particles is not induced by currents within the plasma, but sustained through asymmetric coils, thus enabling improved steady state operation of the machine. This usually comes at the price of deteriorated confinement, and a nontrivial coil design that should be compatible with a quasisymmetric confining magnetic field [15], i.e. a magnetic field 𝑩\boldsymbol{B} whose strength B=𝑩𝑩B=\sqrt{\boldsymbol{B}\cdot\boldsymbol{B}} is independent of a direction 𝒖\boldsymbol{u} in space, 𝒖B=0\boldsymbol{u}\cdot\nabla B=0 (see below for a rigorous definition of quasisymmetric vector field). Here, quasisymmetry is required because it guarantees the existence of a conserved momentum arising from the invariance along 𝒖\boldsymbol{u} of the guiding center Hamiltonian, which depends on the magnetic field only through the strength BB [16]. The conserved quantity is expected to enhance particle confinement along the magnetic field.

A quasisymmetric magnetic field is mathematically characterized by the property that there exists a solenoidal vector field 𝒖\boldsymbol{u} such that both 𝑩\boldsymbol{B} and the field strength BB are Lie-invariant along 𝒖\boldsymbol{u}, i.e.

𝔏𝒖𝑩=𝟎,𝔏𝒖B=0,𝔏𝒖dV=0,\mathfrak{L}_{\boldsymbol{u}}\boldsymbol{B}=\boldsymbol{0},~{}~{}~{}~{}\mathfrak{L}_{\boldsymbol{u}}B=0,~{}~{}~{}~{}\mathfrak{L}_{\boldsymbol{u}}dV=0, (2)

where 𝔏\mathfrak{L} denotes the Lie derivative and dV=dxdydzdV=dxdydz the volume element in 3\mathbb{R}^{3}. Clearly, a symmetric magnetic field is also quasisymmetric. In this case 𝒖=𝒂+𝒃×𝒙\boldsymbol{u}=\boldsymbol{a}+\boldsymbol{b}\times\boldsymbol{x} is the generator of continuous Euclidean isometries, with 𝒂,𝒃3\boldsymbol{a},\boldsymbol{b}\in\mathbb{R}^{3} constant vectors and 𝒙\boldsymbol{x} the position vector in 3\mathbb{R}^{3}. Note that (1a) can also be expressed through the Lie derivative as 𝔏×𝑩𝑩=𝟎\mathfrak{L}_{\nabla\times{\boldsymbol{B}}}\boldsymbol{B}=\boldsymbol{0} and 𝔏𝑩dV=0\mathfrak{L}_{\boldsymbol{B}}dV=0. The quasisymmetry condition (2) is usually written in vector notation as [17],

𝑩×𝒖=ζ,𝒖B=0,𝒖=0.\boldsymbol{B}\times\boldsymbol{u}=\nabla\zeta,~{}~{}~{}~{}\boldsymbol{u}\cdot\nabla B=0,~{}~{}~{}~{}\nabla\cdot\boldsymbol{u}=0. (3)

Here, ζ\zeta is any function. Nonetheless, in stellarator applications it is customary to identify ζ\zeta with the flux function Ψ\Psi taking constant values on the bounding surface, so that both the magnetic field 𝑩\boldsymbol{B} and the quasisymmetry 𝒖\boldsymbol{u} lie on flux surfaces, and the conserved momentum is a combination of the poloidal and toroidal momenta. Unfortunately, the analysis of [18] suggests that fulfilling (1) and (3) simultaneously with a constant pressure at the bounding surface in an asymmetric configuration should not be possible (see also [19, 20, 21, 22]). This is because the Lie-invariance required for a quasisymmetric solution to exist cannot be satisfied through the available geometrical degrees of freedom. A possible strategy to overcome this difficulty is to modify the force balance equation (the first equation in (1a)) by introducing pressure anisotropy. In this regard, it has been suggested that pressure anisotropy could remove the problem of overdetermination encountered in the isotropic setting [17, 23]. In the anisotropic case, anisotropic magnetohydrodynamic equilibria are described by the boundary value problem

(×𝑩)×𝑩=Π,𝑩=0inΩ,𝑩𝒏=0onΩ,\begin{split}&\left({\nabla\times\boldsymbol{B}}\right)\times\boldsymbol{B}=\nabla\cdot\Pi,~{}~{}~{}~{}\nabla\cdot\boldsymbol{B}=0~{}~{}~{}~{}{\rm in}~{}~{}\Omega,\\ &\boldsymbol{B}\cdot\boldsymbol{n}=0~{}~{}~{}~{}{\rm on}~{}~{}\partial\Omega,\end{split} (4)

where in Cartesian coordinates (x1,x2,x3)=(x,y,z)\left({x^{1},x^{2},x^{3}}\right)=\left({x,y,z}\right) the twice contravariant symmetric (here, symmetric does not refer to continuous Euclidean isometries but to the property that the transpose tensor equals the tensor itself) pressure tensor Π\Pi has components

Πij=Pδij+(PP)B2BiBj,i,j=1,2,3.\Pi^{ij}=P_{\perp}\delta^{ij}+\frac{\left({P_{\parallel}-P_{\perp}}\right)}{B^{2}}B^{i}B^{j},~{}~{}~{}~{}i,j=1,2,3. (5)

The functions PP_{\perp} and PP_{\parallel} are pressure fields associated with the perpendicular and parallel directions to the magnetic field 𝑩\boldsymbol{B}. Indeed, setting bi=Bi/Bb^{i}=B^{i}/B, i=1,2,3i=1,2,3, one has

Π=𝒃×(P×𝒃)+(𝒃P)𝒃+(PP)(𝒃𝒃).\nabla\cdot\Pi=\boldsymbol{b}\times\left({\nabla P_{\perp}\times\boldsymbol{b}}\right)+\left({\boldsymbol{b}\cdot\nabla P_{\parallel}}\right)\boldsymbol{b}+\left({P_{\parallel}-P_{\perp}}\right)\nabla\cdot\left({\boldsymbol{b}\boldsymbol{b}}\right). (6)

Hence, the gradient of PP_{\perp} generates the pressure force in the direction perpendicular to 𝑩\boldsymbol{B}, while the gradient of PP_{\parallel} generates the pressure force in the direction parallel to 𝑩\boldsymbol{B}. Isotropic magnetohydrodynamic equilibria can be recovered when P=P=PP_{\perp}=P_{\parallel}=P, with PP the scalar pressure. We also remark that the form of the pressure tensor (5) is rooted in kinetic theory: the pressure fields PP_{\perp} and PP_{\parallel} represent the averaged squared deviations of the charged particle velocities in the guiding center approximation across and along the magnetic field with respect to the mean perpendicular and parallel flows [7, 24, 25, 26].

Our aim in this paper is to study the boundary value problem (4) with the quasisymmetry condition (3) for the magnetic field. It will be shown that, by appropriately choosing the values of PP_{\perp} and PP_{\parallel}, it is possible to construct regular quasisymmetric magnetic fields, i.e. asymmetric solutions 𝑩\boldsymbol{B} of system (3), (4) in a neighborhood UΩU\subset\Omega within an asymmetric toroidal volume Ω\Omega such that UΩ\partial U\cap\partial\Omega\neq\emptyset. We remark that for the time being we shall not be concerned with the feasibility of the obtained equilibria in the laboratory or their stability, but simply focus on their existence.

The present paper is organized as follows. In section 2, general considerations on the geometrical implications of quasisymmetry are given. In particular, equation (3) is written in terms of a set of Clebsch potentials [27]. This form will be useful to build explicit solutions. In section 3, we discuss the representation of asymmetric tori in terms of special sets of coordinates. In section 4, we derive regular symmetric equilibria that solve system (4) within asymmetric toroidal domains. In section 5, we construct regular quasisymmetric solutions of system (4) in a neighborhood UΩU\subset\Omega of an asymmetric toroidal volume Ω\Omega that satisfy tangential boundary conditions on a portion of the boundary UΩ\partial U\cap\partial\Omega\neq\emptyset. Concluding remarks are given in section 5.

2 General remarks on quasisymmetry

Consider Cartesian coordinates (x1,x2,x3)=(x,y,z)\left({x^{1},x^{2},x^{3}}\right)=\left({x,y,z}\right) in Ω\Omega. It is convenient to write the pressure fields PP_{\perp} and PP_{\parallel} as follows:

P=𝒫12σB2,P=𝒫+12σB2.P_{\perp}=\mathcal{P}-\frac{1}{2}\sigma B^{2},~{}~{}~{}~{}P_{\parallel}=\mathcal{P}+\frac{1}{2}\sigma B^{2}. (7)

Here, the functions 𝒫\mathcal{P} and σ\sigma can be interpreted as a reference pressure field and a deviation (anisotropy) term respectively. The Cartesian components of the pressure tensor become

Πij=(𝒫12σB2)δij+σBiBj,i,j=1,2,3.\Pi^{ij}=\left({\mathcal{P}-\frac{1}{2}\sigma B^{2}}\right)\delta^{ij}+\sigma B^{i}B^{j},~{}~{}~{}~{}i,j=1,2,3. (8)

In the following, we denote with i\partial_{i}, i=1,2,3i=1,2,3, the tangent vector in the xix^{i} direction. Then, using 𝑩=0\nabla\cdot\boldsymbol{B}=0, one can verify the identity

Π=(iΠij)j=(𝒫12σB2)+𝑩(σ𝑩).\nabla\cdot\Pi=\left({\partial_{i}\Pi^{ij}}\right)\partial_{j}=\nabla\left({\mathcal{P}-\frac{1}{2}\sigma B^{2}}\right)+\boldsymbol{B}\cdot\nabla\left({\sigma\boldsymbol{B}}\right). (9)

Hence, force balance now takes the form

(1σ)(×𝑩)×𝑩=𝒫12B2σ+(𝑩σ)𝑩.\left({1-\sigma}\right)\left({\nabla\times\boldsymbol{B}}\right)\times\boldsymbol{B}=\nabla\mathcal{P}-\frac{1}{2}B^{2}\nabla\sigma+\left({\boldsymbol{B}\cdot\nabla\sigma}\right)\boldsymbol{B}. (10)

In the presence of symmetry, it is known that system (10) can be reduced to a well-posed Grad-Shafranov type equation under appropriate ellipticity conditions, among which the requirement 1σ>01-\sigma>0 [7, 26, 28]. Isotropic equilibria correspond to the case 𝒫=P\mathcal{P}=P and σ=0\sigma=0. More generally, solutions with constant σ1\sigma\neq 1 are qualitatively equivalent to isotropic configurations with P=𝒫/(1σ)P=\mathcal{P}/(1-\sigma). Next, let P0P_{0}\in\mathbb{R} be a constant. The force balance equation (10) can be trivially satisfied by setting σ=1\sigma=1 and 𝒫=P0/2\mathcal{P}=P_{0}/2 so that

P=12(P0B2),P=12(P0+B2),P_{\perp}=\frac{1}{2}\left({P_{0}-B^{2}}\right),~{}~{}~{}~{}P_{\parallel}=\frac{1}{2}\left({P_{0}+B^{2}}\right), (11)

With this choice, it follows that finding a solution of (4) under the quasisymmetry condition (3) now amounts to solving the boundary value problem

𝑩=0,𝑩×𝒖=ζ,𝒖B=0,𝒖=0inΩ,𝑩𝒏=0onΩ.\begin{split}&\nabla\cdot\boldsymbol{B}=0,~{}~{}~{}~{}\boldsymbol{B}\times\boldsymbol{u}=\nabla\zeta,~{}~{}~{}~{}\boldsymbol{u}\cdot\nabla B=0,~{}~{}~{}~{}\nabla\cdot\boldsymbol{u}=0~{}~{}~{}~{}{\rm in}~{}~{}\Omega,\\ &\boldsymbol{B}\cdot\boldsymbol{n}=0~{}~{}~{}~{}{\rm on}~{}~{}\partial\Omega.\end{split} (12)

Notice that these are just the equations satisfied by a quasisymmetric vector field tangential to Ω\partial\Omega. Furthermore, observe that in this construction the pressure fields PP_{\perp} and PP_{\parallel} are not constrained at the boundary. Instead, they are functions of the magnetic energy density. In particular, the perpendicular pressure field PP_{\perp} exactly balances the magnetic pressure since P+B2/2=P0/2P_{\perp}+B^{2}/2=P_{0}/2, the mechanical pressure along the magnetic field PP_{\parallel} satisfies P+B2/2=P0/2+B2P_{\parallel}+B^{2}/2=P_{0}/2+B^{2}, while the total energy density associated with pressure and magnetic fields is P+P+B2/2=P0+B2/2P_{\perp}+P_{\parallel}+B^{2}/2=P_{0}+B^{2}/2. If one further demands that, for example, P=0P_{\perp}=0 on Ω\partial\Omega, this results in an additional boundary condition for the magnetic field strength, 𝑩2=P0\boldsymbol{B}^{2}=P_{0} on Ω\partial\Omega. We will briefly return to this point later on, although the present study will be mainly focused on the case in which the pressure fields are not constrained at the bounding surface. We also remark that, if the quasisymmetry condition (3) is not imposed, any solenoidal magnetic field satisfying tangential boundary conditions represents a solution of anisotropic magnetohydrodynamics (4) as long as the pressure fields are chosen as in (11).

It is useful to briefly discuss the physical consistency of equation (11). First, observe that both PP_{\perp} and PP_{\parallel} can be made non-negative by taking a sufficiently large constant P0P_{0}. Now suppose that the curvature |𝒃𝒃|\left\lvert{\boldsymbol{b}\cdot\nabla\boldsymbol{b}}\right\rvert is a small quantity, and rewrite the force balance equation (×𝑩)×𝑩=Π\left({\nabla\times\boldsymbol{B}}\right)\times\boldsymbol{B}=\nabla\cdot\Pi in the following form:

12𝒃×(B2×𝒃)+B2𝒃𝒃=𝒃×(P×𝒃)+(PP)𝒃𝒃+[𝒃P+(PP)(𝒃)]𝒃.-\frac{1}{2}\boldsymbol{b}\times\left({\nabla B^{2}\times\boldsymbol{b}}\right)+B^{2}\boldsymbol{b}\cdot\nabla\boldsymbol{b}=\boldsymbol{b}\times\left({\nabla P_{\perp}\times\boldsymbol{b}}\right)+\left({P_{\parallel}-P_{\perp}}\right)\boldsymbol{b}\cdot\nabla\boldsymbol{b}+\left[\boldsymbol{b}\cdot\nabla P_{\parallel}+\left({P_{\parallel}-P_{\perp}}\right)\left({\nabla\cdot\boldsymbol{b}}\right)\right]\boldsymbol{b}. (13)

Collecting small terms involving the curvature |𝒃𝒃|\left\lvert{\boldsymbol{b}\cdot\nabla\boldsymbol{b}}\right\rvert, one thus arrives at the system

B2=PP,\displaystyle B^{2}=P_{\parallel}-P_{\perp}, (14a)
𝒃×[(P+B22)×𝒃]=𝟎,\displaystyle\boldsymbol{b}\times\left[\nabla\left({P_{\perp}+\frac{B^{2}}{2}}\right)\times\boldsymbol{b}\right]=\boldsymbol{0}, (14b)
𝒃P+(PP)(𝒃)=0.\displaystyle\boldsymbol{b}\cdot\nabla P_{\parallel}+\left({P_{\parallel}-P_{\perp}}\right)\left({\nabla\cdot\boldsymbol{b}}\right)=0. (14c)

By eliminating PP_{\parallel}, these equations reduce to

𝒃×[(P+B22)×𝒃]=𝟎,\displaystyle\boldsymbol{b}\times\left[\nabla\left({P_{\perp}+\frac{B^{2}}{2}}\right)\times\boldsymbol{b}\right]=\boldsymbol{0}, (15a)
𝒃(P+B22)=0,\displaystyle\boldsymbol{b}\cdot\nabla\left({P_{\perp}+\frac{B^{2}}{2}}\right)=0, (15b)

which imply equation (11) as desired. Hence, equation (11) corresponds to a magnetic configuration in which the magnetic field curvature |𝒃𝒃|\left\lvert{\boldsymbol{b}\cdot\nabla\boldsymbol{b}}\right\rvert represents a higher-order term in the force balance equation.

Now consider system (12). We must distinguish two cases.

  1. 1.

    The vector fields 𝑩\boldsymbol{B} and 𝒖\boldsymbol{u} are linearly dependent. Then, ζ=𝟎\nabla\zeta=\boldsymbol{0}, and the magnetic field 𝑩\boldsymbol{B} is self-quasisymmetric. Indeed, system (12) reduces to the magnetic differential equation

    𝑩=0,𝑩B=0inΩ,𝑩𝒏=0onΩ.\begin{split}&\nabla\cdot\boldsymbol{B}=0,~{}~{}~{}~{}\boldsymbol{B}\cdot\nabla B=0~{}~{}~{}~{}{\rm in}~{}~{}\Omega,\\ &\boldsymbol{B}\cdot\boldsymbol{n}=0~{}~{}~{}~{}{\rm on}~{}~{}\partial\Omega.\end{split} (16)

    Note that self-quasisymmetry implies that field strength does not change along field lines. Hence, guiding center motion is not accelerated in the direction of the magnetic field, and the conserved momentum is the momentum parallel to 𝑩\boldsymbol{B}. Due to the regularity of the boundary Ω\partial\Omega, the unit outward normal 𝒏\boldsymbol{n} can be expressed as

    𝒏=Ψ|Ψ|,\boldsymbol{n}=\frac{\nabla\Psi}{\left\lvert{\nabla\Psi}\right\rvert}, (17)

    where Ψ\Psi is a single-valued function corresponding to the usual flux function. For the magnetic field 𝑩\boldsymbol{B} to possess nested flux surfaces, we demand that

    𝑩Ψ=0inΩ.\boldsymbol{B}\cdot\nabla\Psi=0~{}~{}~{}~{}{\rm in}~{}~{}\Omega. (18)

    or, denoting with Θ\Theta a possibly multivalued (angle) variable,

    𝑩=Ψ×ΘinΩ.\boldsymbol{B}=\nabla\Psi\times\nabla\Theta~{}~{}~{}~{}{\rm in}~{}~{}\Omega. (19)

    Recall that, due to the solenoidal nature of 𝑩\boldsymbol{B}, the variable Θ\Theta can always be chosen as a single-valued function in a sufficiently small region UΩU\subset\Omega (Lie-Darboux theorem [29]). In the following, we shall refer to functions associated with the representation of vector fields such as Ψ\Psi and Θ\Theta as Clebsch potentials. It should be noted that Clebsch representations of the form (19) always correspond to helicity-free configurations since the vector potential 𝑨\boldsymbol{A} associated with the magnetic field 𝑩=×𝑨\boldsymbol{B}=\nabla\times\boldsymbol{A} can be chosen as an integrable vector field, 𝑨=ΨΘ\boldsymbol{A}=\Psi\nabla\Theta. Using (19), system (16) reduces to a single nonlinear first-order partial differential equation for the variable Θ\Theta,

    |Ψ×Θ|2=E(Ψ,Θ)inΩ,\left\lvert{\nabla\Psi\times\nabla\Theta}\right\rvert^{2}=E\left({\Psi,\Theta}\right)~{}~{}~{}~{}{\rm in}~{}~{}\Omega, (20)

    where E=E(Ψ,Θ)E=E\left({\Psi,\Theta}\right) is any non-negative single-valued function of the variables Ψ\Psi and Θ\Theta. Assume a non-vanishing magnetic field, E0E\neq 0. Then, by defining the function

    Θ=dΘE(Ψ,Θ),\Theta^{\prime}=\int\frac{d\Theta}{\sqrt{E\left({\Psi,\Theta}\right)}}, (21)

    equation (20) can be written as

    |Ψ×Θ|2=1inΩ.\left\lvert{\nabla\Psi\times\nabla\Theta^{\prime}}\right\rvert^{2}=1~{}~{}~{}~{}{\rm in}~{}~{}\Omega. (22)

    Notice that equation (22) resembles the eikonal equation. Indeed, it is equivalent to

    |Θ|2=1|Ψ|2+(ΨΘ)2|Ψ|2inΩ.\left\lvert{\nabla\Theta^{\prime}}\right\rvert^{2}=\frac{1}{\left\lvert{\nabla\Psi}\right\rvert^{2}}+\frac{\left({\nabla\Psi\cdot\nabla\Theta^{\prime}}\right)^{2}}{\left\lvert{\nabla\Psi}\right\rvert^{2}}~{}~{}~{}~{}{\rm in}~{}~{}\Omega. (23)

    In general, the solution of a first-order partial differential equation such as (23) involves the integration of the associated characteristic system of ordinary differential equations. Even if such solution is obtained, it usually has a local nature. Therefore, the existence of a global self-quasisymmetric magnetic field in Ω\Omega is contingent upon the possibility of extending and/or patching local solutions consistently within Ω\Omega. We also remark that, if one does not enforce boundary conditions, constructing self-quasisymmetric fields becomes a simpler task. For example, denoting with (r,ϕ,z)\left({r,\phi,z}\right) cylindrical coordinates, the vector field

    𝑩=f(r,zrϕ)r×(zrϕ),\boldsymbol{B}=f\left({r,\frac{z}{r}-\phi}\right)\nabla r\times\nabla\left({\frac{z}{r}-\phi}\right), (24)

    with ff an arbitrary function of rr and z/rϕz/r-\phi, satisfies 𝑩=0\nabla\cdot\boldsymbol{B}=0, 𝑩B=0\boldsymbol{B}\cdot\nabla B=0, and it does not possess continuous Euclidean symmetries in general because the equation 𝔏𝒖𝑩=𝟎\mathfrak{L}_{\boldsymbol{u}}\boldsymbol{B}=\boldsymbol{0} with 𝒖=𝒂+𝒃×𝒙\boldsymbol{u}=\boldsymbol{a}+\boldsymbol{b}\times\boldsymbol{x} does not have solutions such that 𝒖𝟎\boldsymbol{u}\neq\boldsymbol{0}. In particular, the quantity z/rϕz/r-\phi does not correspond to an Euclidean helical symmetry due to the radial dependence. In addition, the vector field (24) is tangential to cylindrical surfaces (see figure 1). Finally, note that in order to validate the asymmetry of a solution in most cases it is easier to verify the violation of the condition 𝔏𝒖B2=𝒖B2=0\mathfrak{L}_{\boldsymbol{u}}B^{2}=\boldsymbol{u}\cdot\nabla B^{2}=0. Indeed, choosing a curvilinear coordinate system (y1,y2,y3)\left({y^{1},y^{2},y^{3}}\right) such that 𝒖=/y3\boldsymbol{u}=\partial/\partial y^{3} and denoting with BkB^{k} and gkg_{k\ell}, k,=1,2,3k,\ell=1,2,3, the components of magnetic field and the Euclidean metric tensor with respect to the new coordinates, it follows that 𝒖B2=2gkBkB/y3\boldsymbol{u}\cdot\nabla B^{2}=2g_{k\ell}B^{k}\partial B^{\ell}/\partial y^{3} since by hypothesis 𝒖\boldsymbol{u} is an Euclidean isometry such that gk/y3=0\partial g_{k\ell}/\partial y^{3}=0, k,=1,2,3k,\ell=1,2,3. On the other hand, 𝔏𝒖𝑩=(Bi/y3)(/yi)\mathfrak{L}_{\boldsymbol{u}}\boldsymbol{B}=\left({\partial B^{i}/\partial y^{3}}\right)\left({\partial/\partial y^{i}}\right), which implies that 𝔏𝒖𝑩\mathfrak{L}_{\boldsymbol{u}}\boldsymbol{B} cannot vanish as long as 𝔏𝒖B2\mathfrak{L}_{\boldsymbol{u}}B^{2} is different from zero.

    Refer to caption
    Figure 1: (a) Plot of the self-quasisymmetric magnetic field 𝑩=ersin(z/rϕ)r×(z/rϕ)\boldsymbol{B}=e^{-r}\sin\left({z/r-\phi}\right)\nabla r\times\nabla\left({z/r-\phi}\right) on a level set of rr. (b) Plot of the current field ×𝑩\nabla\times\boldsymbol{B}. (c) Plot of the magnetic field strength B2B^{2}. Notice that B2B^{2} does not change along 𝑩\boldsymbol{B} (compare with (a)). (d) Plot of the current field strength |×𝑩|2\left\lvert{\nabla\times\boldsymbol{B}}\right\rvert^{2}.
  2. 2.

    The vector fields 𝑩\boldsymbol{B} and 𝒖\boldsymbol{u} are linearly independent, implying that ζ𝟎\nabla\zeta\neq\boldsymbol{0}. From the second equation in (12), one sees that both 𝑩\boldsymbol{B} and 𝒖\boldsymbol{u} must be orthogonal to ζ\nabla\zeta. However, notice that on the boundary ζ\nabla\zeta is not necessarily aligned with the unit outward normal 𝒏\boldsymbol{n}. Hence, in general, ζ\zeta should not be identified with the flux function Ψ\Psi taking constant values on Ω\partial\Omega. Denoting with ϑ\vartheta and ρ\rho two possibly multivalued (angle) variables, and recalling that 𝑩\boldsymbol{B} and 𝒖\boldsymbol{u} are solenoidal vector fields, the space of solutions of system (12) can be restricted to

    𝑩=ζ×ϑ,𝒖=ζ×ρ.\boldsymbol{B}=\nabla\zeta\times\nabla\vartheta,~{}~{}~{}~{}\boldsymbol{u}=\nabla\zeta\times\nabla\rho. (25)

    Since 𝑩Ψ=0\boldsymbol{B}\cdot\nabla\Psi=0 in Ω\Omega, from (25) one has

    Ψ=Ψ(ζ,ϑ).\Psi=\Psi\left({\zeta,\vartheta}\right). (26)

    For a given flux function Ψ=Ψ(𝒙)\Psi=\Psi\left({\boldsymbol{x}}\right), we assume that equation (26) can be inverted to obtain

    ζ=ζ(ϑ,𝒙).\zeta=\zeta\left({\vartheta,\boldsymbol{x}}\right). (27)

    Then, upon substituting equations (25) and (27) into equation (12), system (12) reduces to the following coupled nonlinear first order partial differential equations for the Clebsch potentials ϑ\vartheta and ρ\rho

    ζ×ϑρ=1,|ζ×ϑ|2=F(ζ,ρ)inΩ,\nabla\zeta\times\nabla\vartheta\cdot\nabla\rho=1,~{}~{}~{}~{}\left\lvert{\nabla\zeta\times\nabla\vartheta}\right\rvert^{2}=F\left({\zeta,\rho}\right)~{}~{}~{}~{}{\rm in}~{}~{}\Omega, (28)

    where FF is any non-negative single-valued function of the Clebsch potentials ζ(ϑ,𝒙)\zeta\left({\vartheta,\boldsymbol{x}}\right) and ρ\rho. Notice that, since equation (28) is a first-order system, the main hurdle toward finding an integral is again represented by the possibility of extending a local solution to the whole Ω\Omega. If Clebsch potentials ϑ\vartheta and ρ\rho can be determined so that (28) is satisfied, the corresponding vector fields (25) define a quasisymmetric solution of anisotropic magnetohydrodynamics. Finally, observe that system (28) can be further reduced to a single nonlinear second-order partial differential equation in regions VΩV\subset\Omega where ζ×B𝟎\nabla\zeta\times\nabla B\neq\boldsymbol{0}. Indeed, the condition 𝒖B=0\boldsymbol{u}\cdot\nabla B=0 then implies that ρ=ρ(ζ,B2)\rho=\rho\left({\zeta,B^{2}}\right) with B2=|ζ×θ|2B^{2}=\left\lvert{\nabla\zeta\times\nabla\theta}\right\rvert^{2}. Therefore, the remaining equation 𝑩×𝒖=ζ\boldsymbol{B}\times\boldsymbol{u}=\nabla\zeta can be written as

    ζ×θ|ζ×θ|2=1ρB2inV.\nabla\zeta\times\nabla\theta\cdot\nabla\left\lvert{\nabla\zeta\times\nabla\theta}\right\rvert^{2}=\frac{1}{\frac{\partial\rho}{\partial B^{2}}}~{}~{}~{}~{}{\rm in}~{}~{}V. (29)

    For a given ρ=ρ(ζ,B2)\rho=\rho\left({\zeta,B^{2}}\right), a quasisymmetric solution in VV can thus be obtained by solving the equation above for the unknown variable θ\theta. We remark that equation (29) is degenerate because it is invariant under the transformation θθ+ξ\theta\rightarrow\theta+\xi, with ξ=ξ(ζ)\xi=\xi\left({\zeta}\right) an arbitrary function of ζ\zeta. Unfortunately, equation (29) cannot hold in the whole Ω\Omega (that is we never have V=ΩV=\Omega). To see this, consider again system (12) and suppose that ζ×B𝟎\nabla\zeta\times\nabla B\neq\boldsymbol{0} in Ω\Omega. Since both 𝑩\boldsymbol{B} and 𝒖\boldsymbol{u} lie on surfaces of constant ζ\zeta, the conditions 𝒖=0\nabla\cdot\boldsymbol{u}=0 and 𝒖B=0\boldsymbol{u}\cdot\nabla B=0 imply that the quasisymmetry must satisfy

    𝒖=σ(ζ,B)ζ×B,\boldsymbol{u}=\sigma\left({\zeta,B}\right)\nabla\zeta\times\nabla B, (30)

    where σ=σ(ζ,B)\sigma=\sigma\left({\zeta,B}\right) is some function of ζ\zeta and BB. Using the property 𝑩ζ=0\boldsymbol{B}\cdot\nabla\zeta=0, the equation 𝑩×𝒖=ζ\boldsymbol{B}\times\boldsymbol{u}=\nabla\zeta therefore reduces to

    σ𝑩B=1.\sigma\boldsymbol{B}\cdot\nabla B=1. (31)

    Due to the regularity of the magnetic field and its derivatives, we must have σ0\sigma\neq 0 for (31) to hold. Using 𝑩=0\nabla\cdot\boldsymbol{B}=0 and 𝑩𝒏=0\boldsymbol{B}\cdot\boldsymbol{n}=0 on Ω\partial\Omega, the equation above implies that

    0=Ω(B𝑩)𝑑V=ΩdVσ.0=\int_{\Omega}\nabla\cdot\left({B\boldsymbol{B}}\right)dV=\int_{\Omega}\frac{dV}{\sigma}. (32)

    For the quasisymmetry 𝒖\boldsymbol{u} to be a continuous function, the fraction 1/σ1/\sigma must be continuous. Hence, equation (32) can be satisfied only if there exists at least one point 𝒙Ω\boldsymbol{x}^{\ast}\in\Omega such that 1/σ(𝒙)=01/\sigma\left({\boldsymbol{x}^{\ast}}\right)=0, implying |𝒖(𝒙)|=\left\lvert{\boldsymbol{u}\left({\boldsymbol{x}^{\ast}}\right)}\right\rvert=\infty. This contradicts the regularity of 𝒖\boldsymbol{u}. Therefore, a global regular quasisymmetric configuration cannot exist in the case ζ×B𝟎\nabla\zeta\times\nabla B\neq\boldsymbol{0}. In other words, if a global regular quasisymmetric configuration exists, there must be regions/points where ζ×B=𝟎\nabla\zeta\times\nabla B=\boldsymbol{0}. Evidently, the property ρ=ρ(ζ,B2)\rho=\rho\left({\zeta,B^{2}}\right) breaks down at those points where ζ×B=𝟎\nabla\zeta\times\nabla B=\boldsymbol{0}, and equation (29) does not hold there.

3 Harmonic orthogonal coordinates and asymmetric tori

In this section, we are concerned with the representation of asymmetric tori through special sets of coordinates. Such representation will be used in the following sections to derive symmetric and quasisymmetric anisotropic magnetohydrodynamic equilibria in asymmetric tori. Let (r,ϕ,z)\left({r,\phi,z}\right) denote cylindrical coordinates. First, consider an axially symmetric torus. Such surface is defined as a level set of the function

Ψax=12[(rr0)2+z2],\Psi_{\rm ax}=\frac{1}{2}\left[\left({r-r_{0}}\right)^{2}+z^{2}\right], (33)

with r0r_{0} a positive real constant representing the major radius of the torus (the distance from the origin to the center of the torus). The axial symmetry is described by the vector field 𝒖=ϕ=r2ϕ\boldsymbol{u}=\partial_{\phi}=r^{2}\nabla\phi. Indeed, 𝔏ϕΨax=0\mathfrak{L}_{\partial_{\phi}}\Psi_{\rm ax}=0. The axially symmetric torus (33) can be thought as a special case of a more general family of toroidal surfaces in 3\mathbb{R}^{3}, which correspond to level sets of the function

ΨT=12[(μμ0)2+(zh)2].\Psi_{\rm T}=\frac{1}{2}\left[\left({\mu-\mu_{0}}\right)^{2}+\left({z-h}\right)^{2}\right]. (34)

Here, μ=μ(x,y)\mu=\mu\left({x,y}\right) is a single-valued function in the (x,y)\left({x,y}\right) plane measuring the distance from the origin in 2\mathbb{R}^{2} and μ0\mu_{0} a positive real constant corresponding to the major radius of the torus (more generally μ0\mu_{0} can be a single-valued function, although this case is not considered here). The more the level sets of μ\mu differ from circles, the more the torus ΨT\Psi_{\rm T} departs from axial symmetry in the (x,y)\left({x,y}\right) plane. Similarly, the single-valued function h=h(𝒙)h=h\left({\boldsymbol{x}}\right) expresses the deviation of the toroidal axis from the (x,y)\left({x,y}\right) plane. Figure 2 shows an axially symmetric torus (33) and three asymmetric tori (34) obtained from different choices of μ\mu and hh. The asymmetry of the tori (b), (c), and (d) in figure 1 can be verified by observing that these surfaces are not Lie-invariant with respect to the generator of continuous Euclidean isometries, that is the equation 𝔏𝒖ΨT=0\mathfrak{L}_{\boldsymbol{u}}\Psi_{\rm T}=0 with 𝒖=𝒂+𝒃×𝒙\boldsymbol{u}=\boldsymbol{a}+\boldsymbol{b}\times\boldsymbol{x} does not have solution for any nontrivial choice of the constant vectors 𝒂,𝒃3\boldsymbol{a},\boldsymbol{b}\in\mathbb{R}^{3}.

Refer to caption
Figure 2: (a) Axially symmetric torus Ψax=0.1\Psi_{\rm ax}=0.1 with r0=1r_{0}=1. (b) Asymmetric torus ΨT=0.1\Psi_{\rm T}=0.1 with μ=x4+y4\mu=\sqrt{x^{4}+y^{4}}, μ0=1\mu_{0}=1, and h=0h=0. (c) Asymmetric torus ΨT=0.1\Psi_{\rm T}=0.1 with μ=r\mu=r, μ0=1\mu_{0}=1, and h=rsin2(3ϕ)h=r\sin^{2}\left({3\phi}\right). (d) Asymmetric torus ΨT=0.1\Psi_{\rm T}=0.1 with μ=x4+y4\mu=\sqrt{x^{4}+y^{4}}, μ0=1\mu_{0}=1, and h=rsin2(3ϕ)h=r\sin^{2}\left({3\phi}\right).

If quasisymmetry is not imposed, solutions of (4) in asymmetric tori (34) can be easily obtained by enforcing (11) and setting

𝑩=ΨT×Φ,\boldsymbol{B}=\nabla\Psi_{\rm T}\times\nabla\Phi, (35)

with Φ\Phi an arbitrary function whose gradient field is linearly independent of ΨT\nabla\Psi_{\rm T}. In this case, a sufficient condition for the perpendicular pressure P=(P0B2)/2P_{\perp}=\left({P_{0}-B^{2}}\right)/2 to be constant on Ω\partial\Omega (which corresponds to a level set of ΨT\Psi_{\rm T} by construction) is that Φ\Phi is chosen by requiring that ΨT×P=𝟎\nabla\Psi_{\rm T}\times\nabla P_{\perp}=\boldsymbol{0}, or, substituting the expression for PP_{\perp},

|ΨT×Φ|2=E(ΨT)inΩ,\left\lvert{\nabla\Psi_{\rm T}\times\nabla\Phi}\right\rvert^{2}=E\left({\Psi_{\rm{}_{T}}}\right)~{}~{}~{}~{}{\rm in}~{}~{}\Omega, (36)

where E=E(ΨT)E=E\left({\Psi_{\rm T}}\right) is a non-negative function of ΨT\Psi_{\rm T}. Of course, if PP_{\perp} is constant on Ω\partial\Omega, so is PP_{\parallel} due to (11). Regarding the solvability of (36), considerations analogous to those pertaining to equation (20) apply. Furthermore, by comparison with equation (20), it is clear that a magnetic field fulfilling (36) is also self-quasisymmetric since 𝑩B=0\boldsymbol{B}\cdot\nabla B=0 (recall equation (16)). Solutions of (36) can be obtained easily in axially symmetric tori (33). Indeed, first observe that the vector fields

𝑩=λϕ,\boldsymbol{B}=\lambda\nabla\phi, (37)

with λ=λ(r,z)\lambda=\lambda\left({r,z}\right) are self-quasisymmetric because they satisfy (16) when Ω\Omega is an axially symmetric torus (33). Then, equation (36) holds provided that λ2=r2E(Ψax)\lambda^{2}=r^{2}E\left({\Psi_{\rm ax}}\right). Of course, the self-quasisymmetry of (37) corresponds to the usual rotational symmetry because 𝔏ϕ𝑩=𝟎\mathfrak{L}_{\partial_{\phi}}\boldsymbol{B}=\boldsymbol{0}. However, the construction leading to (37) can be generalized to a certain class of asymmetric tori and, in particular, to derive symmetric solutions in asymmetric toroidal domains. To see this, notice that the orthogonal coordinates (logr,ϕ,z)\left({\log{r},\phi,z}\right) represent a special case of a larger family of orthogonal harmonic coordinates (x1,x2,x3)=(μ,ν,z)\left({x^{1},x^{2},x^{3}}\right)=\left({\mu,\nu,z}\right) satisfying the following properties in Ω\Omega:

xixj=δij,Δxi=0,|x1|=|x2|,|x3|=1,i,j=1,2,3.\nabla x^{i}\cdot\nabla x^{j}=\delta^{ij},~{}~{}~{}~{}\Delta x^{i}=0,~{}~{}~{}~{}\left\lvert{\nabla x^{1}}\right\rvert=\left\lvert{\nabla x^{2}}\right\rvert,~{}~{}~{}~{}\left\lvert{\nabla x^{3}}\right\rvert=1,~{}~{}~{}~{}i,j=1,2,3. (38)

The coordinates μ\mu and ν\nu can be constructed as follows. Let Γ1\Gamma_{1} and Γ2\Gamma_{2} denote two smooth non-intersecting Jordan (simple and closed) curves in 2\mathbb{R}^{2}. Let Σ1\Sigma_{1} and Σ2\Sigma_{2} be the areas enclosed by Γ1\Gamma_{1} and Γ2\Gamma_{2} respectively. Assume that Σ2Σ1\Sigma_{2}\subset\Sigma_{1} and define Γ=Γ1Γ2\Gamma=\Gamma_{1}\cup\Gamma_{2} and Σ=Σ1Σ¯2\Sigma=\Sigma_{1}-\mkern 1.5mu\overline{\mkern-1.5mu\Sigma\mkern-1.5mu}\mkern 1.5mu_{2}, where the bar denotes the closure of a set. Then, the coordinate μ=μ(x,y)\mu=\mu\left({x,y}\right) is obtained as solution of the following Dirichlet boundary value problem for the Laplace equation,

Δμ=0inΣ,μ=c1onΓ1,μ=c2onΓ2,c1,c2,c1c2.\Delta\mu=0~{}~{}~{}~{}{\rm in}~{}~{}\Sigma,~{}~{}~{}~{}\mu=c_{1}~{}~{}~{}~{}{\rm on}~{}~{}\Gamma_{1},~{}~{}~{}~{}\mu=c_{2}~{}~{}~{}~{}{\rm on}~{}~{}\Gamma_{2},~{}~{}~{}~{}c_{1},c_{2}\in\mathbb{R},~{}~{}~{}~{}c_{1}\neq c_{2}. (39)

Observe that μ\mu can be thought of as a measure of the distance from the origin in 2\mathbb{R}^{2}. Next, the coordinate ν=ν(x,y)\nu=\nu\left({x,y}\right) is chosen as the harmonic conjugate of μ\mu. In particular, denoting with 𝒏~\tilde{\boldsymbol{n}} the unit outward normal to the boundary Γ\Gamma, the coordinate ν\nu is a multivalued (angle) variable whose gradient ν\nabla\nu is a harmonic vector field in Σ\Sigma, i.e.

Δν=0inΣ,ν𝒏~=0onΓ.\Delta\nu=0~{}~{}~{}~{}{\rm in}~{}~{}\Sigma,~{}~{}~{}~{}\nabla\nu\cdot\tilde{\boldsymbol{n}}=0~{}~{}~{}~{}{\rm on}~{}~{}\Gamma. (40)

Finally, equation (38) is satisfied in virtue of the Cauchy-Riemann equations.

As an example, suppose that Γ1\Gamma_{1} and Γ2\Gamma_{2} are the boundaries of the horizontal sections of two confocal elliptic cylinders centered at the origin in 3\mathbb{R}^{3}. Then, the orthogonal harmonic coordinates (μ,ν,z)\left({\mu,\nu,z}\right) correspond to the so-called elliptic cylindrical coordinates, which are related to the Cartesian coordinates (x,y,z)\left({x,y,z}\right) by the transformation

x=acoshμcosν,y=asinhμsinν,x=a\cosh\mu\cos\nu,~{}~{}~{}~{}y=a\sinh\mu\sin\nu, (41)

with aa a positive real constant such that the foci of the elliptic sections of the cylinders are located at 𝒙=(a,0,z)\boldsymbol{x}=\left({a,0,z}\right) and 𝒙=(a,0,z)\boldsymbol{x}=\left({-a,0,z}\right). Introducing the quantity

δ=1a2|μ|2=1a2|ν|2=sin2ν+sinh2μ=(1x2+y2a2)2+4y2a2,\delta=\frac{1}{a^{2}\left\lvert{\nabla\mu}\right\rvert^{2}}=\frac{1}{a^{2}\left\lvert{\nabla\nu}\right\rvert^{2}}=\sin^{2}\nu+\sinh^{2}\mu=\sqrt{\left({1-\frac{x^{2}+y^{2}}{a^{2}}}\right)^{2}+\frac{4y^{2}}{a^{2}}}, (42)

one can derive the following expressions for the quadrant x,y0x,y\geq 0,

μ=arcsinh[121+x2+y2a2+δ],\displaystyle\mu=\operatorname{arcsinh}\left[\frac{1}{\sqrt{2}}\sqrt{-1+\frac{x^{2}+y^{2}}{a^{2}}+\delta}\right], (43a)
ν=arcsin[121x2+y2a2+δ],\displaystyle\nu=\arcsin\left[\frac{1}{\sqrt{2}}\sqrt{1-\frac{x^{2}+y^{2}}{a^{2}}+\delta}\right], (43b)

and also

μ=sinhμcosνx+coshμsinνya(sin2ν+sinh2μ)=sinhμcoshμxx+coshμsinhμyya2(sin2ν+sinh2μ),\displaystyle\nabla\mu=\frac{\sinh\mu\cos\nu\nabla x+\cosh\mu\sin\nu\nabla y}{a\left({\sin^{2}\nu+\sinh^{2}\mu}\right)}=\frac{\frac{\sinh\mu}{\cosh\mu}x\nabla x+\frac{\cosh\mu}{\sinh\mu}y\nabla y}{a^{2}\left({\sin^{2}\nu+\sinh^{2}\mu}\right)}, (44a)
ν=sinhμcosνycoshμsinνxa(sin2ν+sinh2μ)=sinhμcoshμxycoshμsinhμyxa2(sin2ν+sinh2μ).\displaystyle\nabla\nu=\frac{\sinh\mu\cos\nu\nabla y-\cosh\mu\sin\nu\nabla x}{a\left({\sin^{2}\nu+\sinh^{2}\mu}\right)}=\frac{\frac{\sinh\mu}{\cosh\mu}x\nabla y-\frac{\cosh\mu}{\sinh\mu}y\nabla x}{a^{2}\left({\sin^{2}\nu+\sinh^{2}\mu}\right)}. (44b)

In the other quadrants, the coordinate ν\nu can be defined so that ν[0,2π)\nu\in[0,2\pi) when moving around a μ\mu contour such as Γ1\Gamma_{1} or Γ2\Gamma_{2}. In particular, πν\pi-\nu for x<0x<0 and y0y\geq 0, 2πν2\pi-\nu for x0x\geq 0 and y<0y<0, and π+ν\pi+\nu for x,y<0x,y<0 with ν\nu given by (43b). In figure 3, level sets of cylindrical coordinates logr\log{r}, ϕ\phi and elliptic cylindrical coordinates μ\mu, ν\nu as defined in (43) are shown together with the respective gradient fields.

Refer to caption
Figure 3: (a) Contours of ϕ\phi and gradient field ϕ\nabla\phi. (b) Contours of logr\log{r} and gradient field logr\nabla\log{r}. (c) Contours of ν\nu as defined in (43b) and gradient field ν\nabla\nu. (d) Contours of μ\mu as defined in (43a) and gradient field μ\nabla\mu. In these plots, a=2a=2.

By substituting (43a) into the expression (34), one obtains elliptic toroidal surfaces. In the following sections, we will see how to embed symmetric and quasisymmetric magnetic fields within such domains.

4 Symmetric and self-quasisymmetric magnetic fields in asymmetric tori

Let Ω\Omega denote the volume enclosed by the elliptic torus

Ψel=12[(μμ0)2+z2],\Psi_{\rm el}=\frac{1}{2}\left[\left({\mu-\mu_{0}}\right)^{2}+z^{2}\right], (45)

with μ\mu given by (43a). Again, the surface Ψel\Psi_{\rm el} does not possess continuous Euclidean symmetries because the equation 𝔏𝒖Ψel=0\mathfrak{L}_{\boldsymbol{u}}\Psi_{\rm el}=0 with 𝒖=𝒂+𝒃×𝒙\boldsymbol{u}=\boldsymbol{a}+\boldsymbol{b}\times\boldsymbol{x} does not have nontrivial solutions 𝒖𝟎\boldsymbol{u}\neq\boldsymbol{0}. In particular, note that horizontal cuts of Ψel\Psi_{\rm el} are ellipsoidal. As discussed in the previous section, in the absence of quasisymmetry and boundary conditions on the pressure fields, anisotropic magnetohydrodynamic equilibria within Ω\Omega can be obtained by enforcing (11) and by setting 𝑩=Ψel×Φ\boldsymbol{B}=\nabla\Psi_{\rm el}\times\nabla\Phi with Φ\Phi an arbitrary function whose gradient field is linearly independent of Ψel\nabla\Psi_{\rm el}. Observing that Ψelν=0\nabla\Psi_{\rm el}\cdot\nabla\nu=0, a family of solutions with a more familiar form is

𝑩=Ψel×ν+λν,\boldsymbol{B}=\nabla\Psi_{\rm el}\times\nabla\nu+\lambda\nabla\nu, (46)

with λ=λ(μ,z)\lambda=\lambda\left({\mu,z}\right). Next, we seek for a translationally symmetric solution such that the direction of symmetry is 𝒖=z\boldsymbol{u}=\nabla z. For simplicity, assume that 𝑩=λν\boldsymbol{B}=\lambda\nabla\nu. Evidently, 𝑩=𝒖=0\nabla\cdot\boldsymbol{B}=\nabla\cdot\boldsymbol{u}=0. Hence, equation (12) is satisfied provided that

λν×z=ζ,zλ2|ν|2=0.\lambda\nabla\nu\times\nabla z=\nabla\zeta,~{}~{}~{}~{}\frac{\partial}{\partial z}\lambda^{2}\left\lvert{\nabla\nu}\right\rvert^{2}=0. (47)

Since by construction ν×z=μ\nabla\nu\times\nabla z=\nabla\mu and |ν|/z=0\partial\left\lvert{\nabla\nu}\right\rvert/\partial z=0, these conditions are identically satisfied for any λ=λ(μ)\lambda=\lambda\left({\mu}\right). Then, ζ=λ𝑑μ\zeta=\int\lambda d\mu. This example shows that a symmetric solution can be fitted within an asymmetric domain. In figure 4 a plot of a symmetric magnetic field constructed as described above is given.

Refer to caption
Figure 4: (a) Plot of the magnetic field 𝑩=eμν\boldsymbol{B}=-e^{-\mu}\nabla\nu with translational symmetry 𝒖=z\boldsymbol{u}=\nabla z over the surface Ψel=0.1\Psi_{\rm el}=0.1. Here, μ\mu and ν\nu are elliptic cylindrical coordinates as given in (43). This magnetic field is a symmetric solution of anisotropic magnetohydrodynamics (12) with pressure fields (11) in an asymmetric domain whose boundary is a level set of Ψel\Psi_{\rm el}. (b) Plot of the corresponding current ×𝑩=eμ|μ|2z\nabla\times\boldsymbol{B}=e^{-\mu}\left\lvert{\nabla\mu}\right\rvert^{2}\nabla z. In these plots, μ0=1\mu_{0}=1 and a=2a=2.

Let us now consider the existence of self-quasisymmetric magnetic fields (16) within Ω\Omega such that the perpendicular pressure PP_{\perp} is constant on the boundary. Such solution can be obtained by solving (36) with ΨT=Ψel\Psi_{\rm T}=\Psi_{\rm el} for the Clebsch potential Φ\Phi. Assuming that B0B\neq 0, it is convenient to introduce the quantity Φ=Φ/E(Ψel)\Phi^{\prime}=\Phi/\sqrt{E\left({\Psi_{\rm el}}\right)} so that the equation to be solved becomes

|Ψel×Φ|2=1inΩ.\left\lvert{\nabla\Psi_{\rm el}\times\nabla\Phi^{\prime}}\right\rvert^{2}=1~{}~{}~{}~{}{\rm in}~{}~{}\Omega. (48)

Substituting the expression for Ψel\Psi_{\rm el}, one arrives at the following nonlinear first-order partial differential equation

(Φν)2[z2+(μμ0)2a2δ]+[zΦμ(μμ0)Φz]2=a2δinΩ.\left({\frac{\partial\Phi^{\prime}}{\partial\nu}}\right)^{2}\left[z^{2}+\frac{\left({\mu-\mu_{0}}\right)^{2}}{a^{2}\delta}\right]+\left[z\frac{\partial\Phi^{\prime}}{\partial\mu}-\left({\mu-\mu_{0}}\right)\frac{\partial\Phi^{\prime}}{\partial z}\right]^{2}=a^{2}\delta~{}~{}~{}~{}{\rm in}~{}~{}\Omega. (49)

To further simplify the equation, it is convenient to introduce the change of variables (μ,z)(ρ,θ)\left({\mu,z}\right)\rightarrow\left({\rho,\theta}\right) defined by

μμ0=ρcosθ,z=ρsinθ.\mu-\mu_{0}=\rho\cos\theta,~{}~{}~{}~{}z=\rho\sin\theta. (50)

Note that ρ2=2Ψel\rho^{2}=2\Psi_{\rm el}. Then, equation (49) reads as

(Φν)2ρ2(sin2θ+cos2θa2δ)+(Φθ)2=a2δinΩ,\left({\frac{\partial\Phi^{\prime}}{\partial\nu}}\right)^{2}\rho^{2}\left(\sin^{2}\theta+\frac{\cos^{2}\theta}{a^{2}\delta}\right)+\left({\frac{\partial\Phi^{\prime}}{\partial\theta}}\right)^{2}=a^{2}\delta~{}~{}~{}~{}{\rm in}~{}~{}\Omega, (51)

where δ=sin2ν+sinh2(μ0+ρcosθ)\delta=\sin^{2}\nu+\sinh^{2}\left(\mu_{0}+\rho\cos\theta\right). For given ρ\rho, (51) is a nonlinear first-order partial differential equation with two independent variables quadratic in the derivatives on a flux surface Ψel\Psi_{\rm el}. Indeed, it has the form

F(p,q,ν,θ)=fp2+q2g=0,F\left({p,q,\nu,\theta}\right)=fp^{2}+q^{2}-g=0, (52)

where

p=Φν,q=Φθ,f=ρ2(sin2θ+cos2θa2δ),g=a2δ.p=\frac{\partial\Phi^{\prime}}{\partial\nu},~{}~{}~{}~{}q=\frac{\partial\Phi^{\prime}}{\partial\theta},~{}~{}~{}~{}f=\rho^{2}\left(\sin^{2}\theta+\frac{\cos^{2}\theta}{a^{2}\delta}\right),~{}~{}~{}~{}g=a^{2}\delta. (53)

A local solution of (51) can be found by considering the Cauchy problem for the characteristic system of ordinary differential equations associated with (51) (see e.g. [30]). To this end, denote with ξ\xi a parameter and assign initial conditions as below:

ν=ν0(ξ),θ=θ0(ξ),Φ=Φ0(ξ),p=p0(ξ),q=q0(ξ),\nu=\nu_{0}\left({\xi}\right),~{}~{}~{}~{}\theta=\theta_{0}\left({\xi}\right),~{}~{}~{}~{}\Phi^{\prime}=\Phi^{\prime}_{0}\left({\xi}\right),~{}~{}~{}~{}p=p_{0}\left({\xi}\right),~{}~{}~{}~{}q=q_{0}\left({\xi}\right), (54)

where the functions ν0\nu_{0}, θ0\theta_{0}, Φ0\Phi^{\prime}_{0}, p0p_{0}, and q0q_{0} are determined so that the following non-degeneracy, compatibility, and transversality conditions are fulfilled:

(Fp)2+(Fq)20,\displaystyle\left({\frac{\partial F}{\partial p}}\right)^{2}+\left({\frac{\partial F}{\partial q}}\right)^{2}\neq 0, (55a)
(dν0dξ)2+(dθ0dξ)20,\displaystyle\left({\frac{d\nu_{0}}{d\xi}}\right)^{2}+\left({\frac{d\theta_{0}}{d\xi}}\right)^{2}\neq 0, (55b)
F(p0,q0,ν0,θ0)=0,\displaystyle F\left({p_{0},q_{0},\nu_{0},\theta_{0}}\right)=0, (55c)
dΦ0dξ=p0dν0dξ+q0dθ0dξ,\displaystyle\frac{d\Phi^{\prime}_{0}}{d\xi}=p_{0}\frac{d\nu_{0}}{d\xi}+q_{0}\frac{d\theta_{0}}{d\xi}, (55d)
Fpdθ0dξFqdν0dξ0.\displaystyle\frac{\partial F}{\partial p}\frac{d\theta_{0}}{d\xi}-\frac{\partial F}{\partial q}\frac{d\nu_{0}}{d\xi}\neq 0. (55e)

Equation (55a) is identically satisfied as long as ff, pp, and/or qq are different from zero. Indeed,

(Fp)2+(Fq)2=4f2p2+4q20.\left({\frac{\partial F}{\partial p}}\right)^{2}+\left({\frac{\partial F}{\partial q}}\right)^{2}=4f^{2}p^{2}+4q^{2}\neq 0. (56)

Equation (55c) can be solved for q0q_{0}. We have

q02=g0f0p02.q_{0}^{2}=g_{0}-f_{0}p^{2}_{0}. (57)

Here, g0=g(ν0,θ0)g_{0}=g\left({\nu_{0},\theta_{0}}\right) and f0=f(ν0,θ0)f_{0}=f\left({\nu_{0},\theta_{0}}\right). The other equations (55b), (55d), and (55e) can be satisfied, for example, by demanding that ν0=ξ\nu_{0}=\xi, θ0=cθ\theta_{0}=c_{\theta}\in\mathbb{R}, and p0=0p_{0}=0. In this case (55b), (55d), and (55e) become

(dν0dξ)2+(dθ0dξ)2=10,\displaystyle\left({\frac{d\nu_{0}}{d\xi}}\right)^{2}+\left({\frac{d\theta_{0}}{d\xi}}\right)^{2}=1\neq 0, (58a)
dΦ0dξ=p0dν0dξ+q0dθ0dξ=0,\displaystyle\frac{d\Phi^{\prime}_{0}}{d\xi}=p_{0}\frac{d\nu_{0}}{d\xi}+q_{0}\frac{d\theta_{0}}{d\xi}=0, (58b)
Fpdθ0dξFqdν0dξ=2q00.\displaystyle\frac{\partial F}{\partial p}\frac{d\theta_{0}}{d\xi}-\frac{\partial F}{\partial q}\frac{d\nu_{0}}{d\xi}=-2q_{0}\neq 0. (58c)

Hence, a set of consistent initial conditions is

ν0=ξ,θ0=cθ,Φ0=cΦ,p0=0,q0=g0.\nu_{0}=\xi,~{}~{}~{}~{}\theta_{0}=c_{\theta},~{}~{}~{}~{}\Phi^{\prime}_{0}=c_{\Phi^{\prime}},~{}~{}~{}~{}p_{0}=0,~{}~{}~{}~{}q_{0}=\sqrt{g_{0}}. (59)

with cΦc_{\Phi^{\prime}}\in\mathbb{R}. Then, solving the characteristic Lagrange-Charpit system

dνFp=dθFq=dΦpFp+qFq=dpFν=dqFθ=dτ,\frac{d\nu}{\frac{\partial F}{\partial p}}=\frac{d\theta}{\frac{\partial F}{\partial q}}=\frac{d\Phi^{\prime}}{p\frac{\partial F}{\partial p}+q\frac{\partial F}{\partial q}}=-\frac{dp}{\frac{\partial F}{\partial\nu}}=-\frac{dq}{\frac{\partial F}{\partial\theta}}=d\tau, (60)

one obtains a parametric solution ν=ν(ξ,τ)\nu=\nu\left({\xi,\tau}\right), θ(ξ,τ)\theta\left({\xi,\tau}\right), and Φ(ξ,τ)\Phi^{\prime}\left({\xi,\tau}\right) in a neighborhood of the line defined by the initial conditions (59). This result implies that, for a given function E(Ψel)>0E\left({\Psi_{\rm el}}\right)>0, a self-quasisymmetric solution 𝑩=E(Ψel)Ψel×Φ\boldsymbol{B}=\sqrt{E\left({\Psi_{\rm el}}\right)}\nabla\Psi_{\rm el}\times\nabla\Phi^{\prime} of anisotropic magnetohydrodynamics with constant perpendicular pressure on the boundary can be constructed in a neighborhood UΩU\subset\Omega such that the boundary condition 𝑩𝒏=0\boldsymbol{B}\cdot\boldsymbol{n}=0 is satisfied on UΩ\partial U\cap\partial\Omega\neq\emptyset. As shown by the example above, in the context of anisotropic magnetohydrodynamics the existence of quasisymmetric solutions ultimately translates into the global consistency of locally obtained solutions, i.e. whether the solution defined in UU can be prolonged to cover the whole Ω\Omega. We therefore conjecture that rigorous results concerning the existence of quasisymmetric vector fields cannot be obtained unless the issue of global extension of solutions of first-order nonlinear partial differential equations is carefully addressed. Finally, we observe that the results discussed in the present section apply to asymmetric tori generated trough general orthogonal harmonic coordinates, i.e. they are not restricted to elliptic geometry.

5 Quasisymmetric magnetic fields in asymmetric tori

The purpose of the present section is to derive quasisymmetric magnetic fields. We will see that these vector fields can be interpreted as local solutions of anisotropic magnetohydrodynamics (12) in an asymmetric torus with perpendicular and parallel pressure fields given by (11). In particular, these solutions hold in a neighborhood UΩU\subset\Omega with UΩ\partial U\cap\partial\Omega\neq\emptyset, they do not have continuous Euclidean symmetries, no boundary conditions are imposed on the pressure fields, and 𝑩×𝒖=ζ𝟎\boldsymbol{B}\times\boldsymbol{u}=\nabla\zeta\neq\boldsymbol{0} (the fields are not self-quasisymmetric). To achieve this goal, we solve equation (28) in a neighborhood UU of a toroidal domain Ω\Omega whose boundary Ω\partial\Omega corresponds to a level set of a function ΨT\Psi_{\rm T}.

First, we prescribe the toroidal domain Ω\Omega. Instead of breaking axial symmetry by modifying the distance function rr with a new variable μ\mu, it is convenient to deform the torus by introducing a non-zero hh in ΨT{\Psi_{\rm T}} of (34). To simplify calculations, we postulate that h=h(r,ϕ)h=h\left({r,\phi}\right). We thus have

ΨT=12[(rr0)2+(zh)2].\Psi_{\rm T}=\frac{1}{2}\left[\left({r-r_{0}}\right)^{2}+\left({z-h}\right)^{2}\right]. (61)

For the magnetic field to possess nested flux surfaces, we further demand that

𝑩=f(r,zh)r×(zh).\boldsymbol{B}=f\left({r,z-h}\right)\nabla r\times\nabla\left({z-h}\right). (62)

Here, ff is a function of rr and zhz-h to be determined together with hh by imposing the quasisymmetry condition. Notice that 𝑩ΨT=0\boldsymbol{B}\cdot\nabla\Psi_{\rm T}=0 implying 𝑩𝒏=0\boldsymbol{B}\cdot\boldsymbol{n}=0 on Ω\partial\Omega. The magnetic field also satisfies 𝑩=0\nabla\cdot\boldsymbol{B}=0. Hence, the remaining equations in (12) are

𝑩×𝒖=dgdζζ,𝒖B=0,𝒖=0inΩ.\boldsymbol{B}\times\boldsymbol{u}=\frac{dg}{d\zeta}\nabla\zeta,~{}~{}~{}~{}\boldsymbol{u}\cdot\nabla B=0,~{}~{}~{}~{}\nabla\cdot\boldsymbol{u}=0~{}~{}~{}~{}{\rm in}~{}~{}\Omega. (63)

In the equation above we have replaced ζ\zeta with a function of ζ\zeta, g=g(ζ)g=g\left({\zeta}\right), for later convenience. Assuming the Clebsch representation 𝒖=ζ×ρ\boldsymbol{u}=\nabla\zeta\times\nabla\rho with ζ=zh\zeta=z-h, from (63) we thus arrive at (28), which now reads as a system of equations that must be solved for the Clebsch potential ρ\rho and the displacement hh,

r×(zh)ρ=1,\displaystyle\nabla r\times\nabla\left({z-h}\right)\cdot\nabla\rho=1, (64a)
(zh)×ρ[1r2(hϕ)2]=0.\displaystyle\nabla\left({z-h}\right)\times\nabla\rho\cdot\nabla\left[\frac{1}{r^{2}}\left({\frac{\partial h}{\partial\phi}}\right)^{2}\right]=0. (64b)

Observe that in deriving (64a) and (64b) we set dg/dζ=fdg/d\zeta=f and eliminated the dependence of ff on rr. Next, define the quantity

η=1rhϕ.\eta=\frac{1}{r}\frac{\partial h}{\partial\phi}. (65)

Then, assuming that (zh)\nabla\left({z-h}\right) and (r1h/ϕ)2\nabla\left({r^{-1}\partial h/\partial\phi}\right)^{2} are linearly independent, equation (64b) implies that

ρ=ρ(zh,η).\rho=\rho\left({z-h,\eta}\right). (66)

Substituting (66) into (64a) we arrive at

ρηr×(zh)η=1,\frac{\partial\rho}{\partial\eta}\nabla r\times\nabla\left({z-h}\right)\cdot\nabla\eta=1, (67)

which, after some manipulations, becomes

ρη1r22hϕ2=1,-\frac{\partial\rho}{\partial\eta}\frac{1}{r^{2}}\frac{\partial^{2}h}{\partial\phi^{2}}=1, (68)

Since hh does not depend on zz, we put ρ=ρ(η)\rho=\rho\left({\eta}\right). Then, for given ρ(η)\rho\left({\eta}\right), equation (68) must be solved for hh. Now recall that h/ϕ=rη\partial h/\partial\phi=r\eta. Therefore, equation (67) is equivalent to

ρϕ=r.\frac{\partial\rho}{\partial\phi}=-r. (69)

Assuming that the function ρ=ρ(η)\rho=\rho\left({\eta}\right) can be inverted with inverse ρ1\rho^{-1} and integrating with respect to ϕ\phi gives

hϕ=rρ1(αrϕ),\frac{\partial h}{\partial\phi}=r\rho^{-1}\left({\alpha-r\phi}\right), (70)

where α=α(r)\alpha=\alpha\left({r}\right) is an arbitrary integration factor. Once ρ1\rho^{-1} is assigned, a further integration with respect to ϕ\phi gives the desired solution. However, due to the radial dependence of αrϕ\alpha-r\phi, the quantity h/ϕ\partial h/\partial\phi is not periodic in ϕ\phi. Therefore, the components of the corresponding quasisymmetric magnetic field (62) and the flux function (61) become multivalued functions, and the obtained quasisymmetric solution is well-defined only locally. In particular, the solution is defined in some region UΩU\subset\Omega, and the irregularities of the bounding surface Ω\partial\Omega arising from the multivalued nature of ΨT\Psi_{\rm T} can be removed by repairing (smoothing) the flux function ΨT\Psi_{\rm T} outside UU. To see this, set ρ(η)=k1η\rho\left({\eta}\right)=k^{-1}\eta, with k>0k>0 a real constant, and choose f=sin(zh)f=\sin\left({z-h}\right). Then, we have

hϕ=kr(αrϕ),h=krϕ(α12rϕ)+β,\frac{\partial h}{\partial\phi}=kr\left({\alpha-r\phi}\right),~{}~{}~{}~{}h=kr\phi\left({\alpha-\frac{1}{2}r\phi}\right)+\beta, (71)

with β=β(r)\beta=\beta\left({r}\right) an arbitrary integration factor, and also

𝑩=sin[zβkrϕ(α12rϕ)][rϕ+k(αrϕ)z],\displaystyle\boldsymbol{B}=-\sin\left[z-\beta-kr\phi\left({\alpha-\frac{1}{2}r\phi}\right)\right]\left[r\nabla\phi+k\left({\alpha-r\phi}\right)\nabla z\right], (72a)
𝒖=r(αrϕ)ϕ+r+r(kα22+β)z.\displaystyle\boldsymbol{u}=r\left({\frac{\partial\alpha}{\partial r}-\phi}\right)\nabla\phi+\nabla r+\frac{\partial}{\partial r}\left(k\frac{\alpha^{2}}{2}+\beta\right)\nabla z. (72b)

For the component of the magnetic field to be single-valued, one must determine α\alpha and β\beta so that

𝑩|ϕ=0=𝑩|ϕ=2π,\boldsymbol{B}\rvert_{\phi=0}=\boldsymbol{B}\rvert_{\phi=2\pi}, (73)

which implies

zβ=zβ2πkr(απr)+2πm,m\displaystyle z-\beta=z-\beta-2\pi kr\left({\alpha-\pi r}\right)+2\pi m,~{}~{}~{}~{}m\in\mathbb{Z} (74a)
kα=k(α2πr).\displaystyle k\alpha=k\left({\alpha-2\pi r}\right). (74b)

However, excluding the case k=0k=0 which corresponds to axial symmetry, these conditions cannot be satisfied. Hence, the local nature of the solution.

As an example of quasisymmetric configuration, consider the case α=β=0\alpha=\beta=0. Then, one can verify the quasisymmetry of the derived solution as below:

𝑩=sin(z+12kr2ϕ2)(rϕkrϕz),\displaystyle\boldsymbol{B}=-\sin\left(z+\frac{1}{2}kr^{2}\phi^{2}\right)\left(r\nabla\phi-kr\phi\nabla z\right), (75a)
𝑩=0,\displaystyle\nabla\cdot\boldsymbol{B}=0, (75b)
B2=sin2(z+12kr2ϕ2)(1+k2r2ϕ2),\displaystyle B^{2}=\sin^{2}\left({z+\frac{1}{2}kr^{2}\phi^{2}}\right)\left({1+k^{2}r^{2}\phi^{2}}\right), (75c)
𝒖=rrϕϕ,\displaystyle\boldsymbol{u}=\nabla r-r\phi\nabla\phi, (75d)
𝒖=0,\displaystyle\nabla\cdot\boldsymbol{u}=0, (75e)
𝑩×𝒖=cos(z+12kr2ϕ2),\displaystyle\boldsymbol{B}\times\boldsymbol{u}=-\nabla\cos\left({z+\frac{1}{2}kr^{2}\phi^{2}}\right), (75f)
𝒖B2=0,\displaystyle\boldsymbol{u}\cdot\nabla B^{2}=0, (75g)
𝑩ΨT=0.\displaystyle\boldsymbol{B}\cdot\nabla\Psi_{\rm T}=0. (75h)

We also remark that this solution does not have continuous Euclidean symmetries because the equation 𝔏𝒖𝑩=𝟎\mathfrak{L}_{\boldsymbol{u}}\boldsymbol{B}=\boldsymbol{0} with 𝒖=𝒂+𝒃×𝒙\boldsymbol{u}=\boldsymbol{a}+\boldsymbol{b}\times\boldsymbol{x} does not have solutions with 𝒂,𝒃𝟎\boldsymbol{a},\boldsymbol{b}\neq\boldsymbol{0}. A plot of the obtained magnetic field and related quantities is given in figure 5.

Refer to caption
Figure 5: (a) Plot of the quasisymmetric magnetic field 𝑩\boldsymbol{B} of (75a) on a level set of (61). (b) Plot of the current ×𝑩\nabla\times\boldsymbol{B}. (c) Plot of the quasisymmetry 𝒖\boldsymbol{u}. (d) Plot of the field stregth B2B^{2}. Observe that 𝑩\boldsymbol{B}, ×𝑩\nabla\times\boldsymbol{B}, 𝒖\boldsymbol{u}, and ΨT\Psi_{\rm T} exhibit a discontinuity at ϕ=0\phi=0 due to the multivalued nature of ϕ\phi. In these plots, k=0.18k=0.18.

The quasisymmetric magnetic field of equation (75) is such that the quasisymmetry 𝒖\boldsymbol{u} is not tangential to flux surfaces ΨT\Psi_{\rm T} since 𝒖ΨT0\boldsymbol{u}\cdot\nabla\Psi_{\rm T}\neq 0. However, in the design of a stellarator it is customary to demand that 𝒖\boldsymbol{u} lies on flux surfaces to ensure that the conserved momentum associated with guiding center motion is a combination of the poloidal and toroidal momenta. This requirement amounts to identifying ζ\zeta with the flux function ΨT\Psi_{\rm T} in the quasisymmetry equations (3). Such configuration can be achieved by repeating the procedure discussed above while enforcing the Clebsch representations

𝑩=f(ΨT)r×(zh),𝒖=ΨT×ρ,\boldsymbol{B}=f\left({\Psi_{\rm T}}\right)\nabla r\times\nabla\left({z-h}\right),~{}~{}~{}~{}\boldsymbol{u}=\nabla\Psi_{\rm T}\times\nabla\rho, (76)

where ff is an arbitrary function of ΨT\Psi_{\rm T}. In this case, system (28) reduces to

r×(zh)ρ=1,\displaystyle\nabla r\times\nabla\left({z-h}\right)\cdot\nabla\rho=1, (77a)
ΨT×ρ[1r2(hϕ)2]=0.\displaystyle\nabla\Psi_{\rm T}\times\nabla\rho\cdot\nabla\left[\frac{1}{r^{2}}\left({\frac{\partial h}{\partial\phi}}\right)^{2}\right]=0. (77b)

Here, we set dg/dζ=dg/dΨT=fdg/d\zeta=dg/d\Psi_{\rm T}=f. Again, defining η=r1h/ϕ\eta=r^{-1}\partial h/\partial\phi and requiring that ρ=ρ(η)\rho=\rho\left({\eta}\right), one arrives at equation (68) with solution (70). For example, if ρ=k1η\rho=k^{-1}\eta and α=β=0\alpha=\beta=0 one obtains the family of quasisymmetry magnetic fields

𝑩=f(rϕkrϕz),\displaystyle\boldsymbol{B}=-f\left(r\nabla\phi-kr\phi\nabla z\right), (78a)
𝑩=0,\displaystyle\nabla\cdot\boldsymbol{B}=0, (78b)
B2=f2(1+k2r2ϕ2),\displaystyle B^{2}=f^{2}\left({1+k^{2}r^{2}\phi^{2}}\right), (78c)
𝒖=(rr0)z+(z+12kr2ϕ2)(rrϕϕ),\displaystyle\boldsymbol{u}=-\left({r-r_{0}}\right)\nabla z+\left({z+\frac{1}{2}kr^{2}\phi^{2}}\right)\left({\nabla r-r\phi\nabla\phi}\right), (78d)
𝒖=0,\displaystyle\nabla\cdot\boldsymbol{u}=0, (78e)
𝑩×𝒖=fΨT=g,\displaystyle\boldsymbol{B}\times\boldsymbol{u}=f\nabla\Psi_{\rm T}=\nabla g, (78f)
𝒖B2=0,\displaystyle\boldsymbol{u}\cdot\nabla B^{2}=0, (78g)
𝑩ΨT=0,\displaystyle\boldsymbol{B}\cdot\nabla\Psi_{\rm T}=0, (78h)
𝒖ΨT=0.\displaystyle\boldsymbol{u}\cdot\nabla\Psi_{\rm T}=0. (78i)

Observe that both 𝑩\boldsymbol{B} and 𝒖\boldsymbol{u} are tangential to flux surfaces.

The construction of quasisymmetric magnetic fields presented above can be generalized to the case of surfaces defined through harmonic orthogonal coordinates (μ,ν,z)\left({\mu,\nu,z}\right). Setting h=h(μ,ν)h=h\left({\mu,\nu}\right), the relevant Clebsch parametrization is

ΨT=12[(μμ0)2+(zh)2],𝑩=f(ΨT)μ×(zh),𝒖=ΨT×ρ.\Psi_{\rm T}=\frac{1}{2}\left[\left({\mu-\mu_{0}}\right)^{2}+\left({z-h}\right)^{2}\right],~{}~{}~{}~{}\boldsymbol{B}=f\left({\Psi_{\rm T}}\right)\nabla\mu\times\nabla\left({z-h}\right),~{}~{}~{}~{}\boldsymbol{u}=\nabla\Psi_{\rm T}\times\nabla\rho. (79)

Then, putting dg/dΨT=fdg/d\Psi_{\rm T}=f, the quasisymmetry equations (28) now read

ΨT×ρ{|μ|2[1+|μ|2(hν)2]}=0,\displaystyle\nabla\Psi_{\rm T}\times\nabla\rho\cdot\nabla\left\{\left\lvert{\nabla\mu}\right\rvert^{2}\left[1+\left\lvert{\nabla\mu}\right\rvert^{2}\left({\frac{\partial h}{\partial\nu}}\right)^{2}\right]\right\}=0, (80a)
μ×(zh)ρ=1.\displaystyle\nabla\mu\times\nabla\left({z-h}\right)\cdot\nabla\rho=1. (80b)

Again, requiring that ρ=ρ(η)\rho=\rho\left({\eta}\right) with

η=|μ|2[1+|μ|2(hν)2],\eta=\left\lvert{\nabla\mu}\right\rvert^{2}\left[1+\left\lvert{\nabla\mu}\right\rvert^{2}\left({\frac{\partial h}{\partial\nu}}\right)^{2}\right], (81)

we arrive at one equation which determines the displacement hh:

dρdηην|μ|2=1.-\frac{d\rho}{d\eta}\frac{\partial\eta}{\partial\nu}\left\lvert{\nabla\mu}\right\rvert^{2}=1. (82)

For a given ρ\rho, the corresponding solution hh of the equation above assigns a family of quasisymmetric magnetic fields parametrized by the function ff as defined in equation (79). Nonetheless, as in the previous cases, the globality of the solution is not guaranteed.

6 Concluding remarks

In this work, we studied the existence of quasisymmetric magnetic fields in asymmetric toroidal domains. If the perpendicular and parallel pressure fields are chosen as in equation (11), this problem is equivalent to finding quasisymmetric magnetohydrodynamic equilibria within the framework of anisotropic magentohydrodynamics. In particular, constructing a quasisymmetric configuration amounts to solving system (12) for the vector fields 𝑩\boldsymbol{B} and 𝒖\boldsymbol{u}. This system can be written in terms of two coupled nonlinear first-order partial differential equations (28), or a single nonlinear second-order partial differential equation (29). By using harmonic orthogonal coordinates, we first devised a method to obtain symmetric solutions of (28) in asymmetric toroidal domains. Then, we constructed regular local self-quasisymmetric and quasisymmetric magnetic fields in asymmetric tori by solving (28) through a Clebsch parametrization of the solution tailored on the flux function associated with the toroidal boundary. These solutions are local in the sense that they solve (28) in a region UΩU\subset\Omega and satisfy boundary conditions on a portion of the boundary, UΩ\partial U\cap\partial\Omega\neq\emptyset.

The obtained results highlight two aspects that characterize quasisymmetric vector fields. On one hand, the symmetry of the boundary should be treated separately from the symmetry of the solution, since we have shown that a symmetric magnetic field can be fitted within an asymmetric domain. On the other hand, due to the first-order nature of the governing equations, the mathematical challenge posed by quasisymmetry can be ascribed to the local nature of the solutions obtained by integrating the characteristic system of ordinary differential equations. Finally, we note that the possibility that the obstruction encountered in the derivation of global solutions is intrinsic to three-dimensional equilibria cannot be ruled out, in the sense that the absence of Euclidean isometry may prevent the existence of such fields. In particular, it appears that the existence of global quasisymmetric fields is contingent upon the availability of curvilinear coordinates whose metric coefficients exhibit invariance properties analogous to those satisfied by cylindrical coordinates, and specifically by the toroidal angle ϕ\phi which obeys |ϕ|/ϕ=0\partial\left\lvert{\nabla\phi}\right\rvert/\partial\phi=0.

Acknowledgment

The research of NS was supported by JSPS KAKENHI Grant No. 21K13851 and 17H01177.

Data Availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  • [1] Z. Yoshida and H. Yamada, Structurally-unstable electrostatic potentials in plasmas, Prog. Theor. Phys. 84, 2 (1990).
  • [2] H. Grad and H. Rubin, Hydromagnetic equilibria and force-free fields, Proc. 2nd United Nations Int. Conf. on the Peaceful Uses of Atomic Energy 31, pp. 190-197 (1958).
  • [3] J. W. Edenstrasser, Unified treatment of symmetric MHD equilibria, J. Plasma Phys. 24, pp. 299-313 (1980).
  • [4] J. W. Edenstrasser, The only three classes of symmetric MHD equilibria, J. Plasma Phys. 24, pp. 515–518 (1980).
  • [5] N. Sato, Existence of ideal magnetofluid equilibria without continuous Euclidean symmetries, Plasma Phys. Control. Fusion 61, 125014 (2019).
  • [6] N. Sato, Symmetric ideal magnetofluidostatic equilibria with nonvanishing pressure gradients in asymmetric confinement vessels, Phys. Plasmas 27, 122507 (2020).
  • [7] H. Grad, Toroidal containment of a plasma, Phys. Fluids 10, 1 (1967).
  • [8] M. Eisenberg and R. Guy, A proof of the hairy ball theorem, The Americal Mathematical Monthly 86, 7, pp. 571-574 (1979).
  • [9] Z. Yoshida and Y. Giga, Remarks on spectra of operator rot, Math. Z. 204, pp. 235-245 (1990).
  • [10] O. P. Bruno, Existence of three-dimensional toroidal MHD equilibria with nonconstant pressure, Comm. Pure Appl. Math. 49, pp. 717-764 (1996).
  • [11] A. Enciso and D. Peralta-Salas, Beltrami fields with a nonconstant proportionality factor are rare, Arch. Rational Mech. Anal. 220, pp. 243-260 (2016).
  • [12] R.L. Dewar, Z. Yoshida, A. Bhattacharjee, S.R. Hudson, Variational formulation of relaxed and multi-region relaxed magnetohydrodynamics, J. Plasma Phys. 81, 515810604 (2015).
  • [13] S. R. Hudson, R. L. Dewar, G. Dennis, M. J. Hole, M. McGann, G. von Nessi, and S. Lazerson, Computation of multi-region relaxed magnetohydrodynamic equilibria Phys. Plasmas 19, 112502 (2012).
  • [14] Z. S. Qu, D Pfefferlé, S. R. Hudson, A. Baillod, A. Kumar, R. L. Dewar, and M. J. Hole, Stepped pressure equilibrium with relaxed flow and applications in reversed-field pinch plasmas, Plasma Phys. Control. Fusion 62, 054002 (2020).
  • [15] P. Helander, Theory of plasma confinement in non-axisymmetric magnetic fields, Rep. Prog. Phys. 77, 087001 (2014).
  • [16] J. R. Cary and A. J. Brizard, Hamiltonian theory of guiding-center motion, Rev. Mod. Phys. 81, 2 (2009).
  • [17] E. Rodriguez, P. Helander, and A. Bhattacharjee, Necessary and sufficient conditions for quasisymmetry, Phys. Plasmas 27, 062501 (2020).
  • [18] D. A. Garren, and A. H. Boozer, Existence of quasihelically symmetric stellarators, Phys. Fluids B: Plasma Phys. 3, 2822 (1991).
  • [19] W. Sengupta, E. J. Paul, H. Weitzner, and A. Bhattacharjee, Vacuum magnetic fields with exact quasisymmetry near a flux surface. Part 1. Solutions near an axisymmetric surface, J. Plasma Phys. 87, 2 (2021).
  • [20] G. G. Plunk and P. Helander, Quasi-axisymmetric magnetic fields: weakly non-axisymmetric case in a vacuum, J. Plasma Phys. 84, 2 (2018).
  • [21] P. Constantin, T. D. Drivas, and D. Ginsberg, On quasisymmetric plasma equilibria sustained by small force, J. Plasma Phys. 87, 1 (2021).
  • [22] P. Constantin, T. Drivas, and D. Ginsberg, Flexibility and Rigidity in Steady Fluid Motion, Comm. Math. Phys. (2021).
  • [23] E. Rodriguez and A. Bhattacharjee, Solving the problem of overdetermination of quasisymmetric equilibrium solutions by near-axis expansions. II. Circular axis stellarator solutions, Phys. Plasmas 28, 012509 (2021).
  • [24] H. Grad, The guiding center plasma, Proc. Sympos. Appl. Math. 18, pp. 162-248 Amer. Math. Soc., Providence, R.I. (1967).
  • [25] D. Dobrott and J. M. Greene, Steady Flow in the Axially Symmetric Torus Using the Guiding-Center Equations, The Physics of Fluids 13, 2391 (1970).
  • [26] R. Iacono, A. Bondeson, F. Troyon, and R. Gruber, Axisymmetric toroidal equilibrium with flow and anisotropic pressure, Physics of Fluids B: Plasma Physics 2, 1794 (1990).
  • [27] Z. Yoshida, Clebsch parametrization: basic properties and remarks on its applications J. Math. Phys 50, 113101 (2009).
  • [28] B. Layden, M. J. Hole, and R. Ridden-Harper, High-beta equilibria in tokamaks with pressure anisotropy and toroidal flow, Phys. Plamas 22, 122513 (2015).
  • [29] M. de Léon, Methods of Differential Geometry in Analytical Mechanics, Elsevier, New York, pp. 250-253 (1989).
  • [30] A. D. Polyanin, V. F. Zaitsev, and A. Moussiaux, Handbook of first order partial differential equations, Taylor & Francis, New York, pp. 371-372 (2002).