Hostname: page-component-848d4c4894-hfldf Total loading time: 0 Render date: 2024-05-02T21:15:01.867Z Has data issue: false hasContentIssue false

Magnetic fields with general omnigenity

Published online by Cambridge University Press:  27 February 2024

Daniel W. Dudt
Affiliation:
Princeton University, Princeton, NJ 08544, USA
Alan G. Goodman
Affiliation:
Max-Planck-Institut für Plasmaphysik, D-17491 Greifswald, Germany
Rory Conlin
Affiliation:
Princeton University, Princeton, NJ 08544, USA
Dario Panici
Affiliation:
Princeton University, Princeton, NJ 08544, USA
Egemen Kolemen*
Affiliation:
Princeton University, Princeton, NJ 08544, USA Princeton Plasma Physics Laboratory, Princeton, NJ 08543, USA
*
Email address for correspondence: ekolemen@princeton.edu

Abstract

Omnigenity is a desirable property of toroidal magnetic fields that ensures confinement of trapped particles. Confining charged particles is a basic requirement for any fusion power plant design, but it can be difficult to satisfy with the non-axisymmetric magnetic fields used by the stellarator approach. Every ideal magnetohydrodynamic equilibrium previously found to approximate omnigenity has been either axisymmetric, quasi-symmetric or has poloidally closed contours of magnetic field strength $B$. However, general omnigenous equilibria are a much larger design space than these subsets. A new model is presented and employed in the DESC stellarator optimization suite to represent and discover the full parameter space of omnigenous equilibria. Although exact omnigenity aside from quasi-symmetry is impossible, these results reveal that excellent particle confinement can be achieved in practice. Examples far from quasi-symmetry with poloidally, helically and toroidally closed $B$ contours are attained with DESC and shown to have low neoclassical collisional transport and fast particle losses.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2024. Published by Cambridge University Press

1 Introduction

Controlled nuclear fusion offers the possibility of reliable power generation without greenhouse gas emissions to satiate the growing global energy demand. The magnetic confinement approach to fusion involves containing a high-temperature plasma for long periods of time and requires a magnetic geometry that can confine these energetic charged particles. The stellarator concept (Spitzer Reference Spitzer1958) makes an attractive fusion reactor: its reliance on external magnetic fields rather than driving plasma current to confine the plasma enables it to operate at steady state and makes it less susceptible to plasma disruptions (Helander et al. Reference Helander, Beidler, Bird, Drevlak, Feng, Hatzky, Jenko, Kleiber, Proll and Turkin2012). Yet historically, stellarators were not pursued due to their relatively high levels of neoclassical transport. Unlike its axisymmetric counterpart, the tokamak, stellarators are not guaranteed to confine trapped particles in a collisionless plasma. Careful attention must be paid to design a stellarator with comparably good particle confinement, but this is becoming increasingly achievable with modern optimization tools.

Trapped particles cannot be avoided altogether in a toroidal geometry. If the radial drifts of trapped particles vanish everywhere, it is called an isodynamic magnetic field, but this is similarly unrealistic (Palumbo Reference Palumbo1968; Helander Reference Helander2014). The next best theoretical option is omnigenity, the class of magnetic fields in which the bounce-averaged radial drifts of trapped particles vanish (Hall & McNamara Reference Hall and McNamara1975; Cary & Shasharina Reference Cary and Shasharina1997b). Pseudo-symmetry is a broader category of fields in which the contours of constant magnetic field strength $B=|\boldsymbol {B}|$ on a flux surface are closed curves (Isaev et al. Reference Isaev, Mikhailov, Monticello, Mynick, Subbotin, Ku and Reiman1999). Omnigenity implies pseudo-symmetry and also that the $B_{\mathrm {max}}$ contour is straight in Boozer coordinates (and other straight field line coordinate systems with a Jacobian that only depends on $B$ and the flux surface label) (Helander Reference Helander2014). There are three classes of omnigenous magnetic fields: the $B$ contours can close either poloidally, helically or toroidally. These cases will be referred to as omnigenity with poloidal contours (OP), omnigenity with helical contours (OH) and omnigenity with toroidal contours (OT). (The term quasi-isodynamic (QI) has often been used in previous literature to mean omnigenity with poloidally closed $B$ contours. This terminology has been inconsistent, however, and is easily conflated with more general omnigenity; QI is avoided in favour of OP in this article for the sake of clarity.) Quasi-symmetry (QS) is a special subset of omnigenity where all contours of $B$ are straight in these coordinates, not only the $B_{\mathrm {max}}$ contour (Nührenberg & Zille Reference Nührenberg and Zille1988; Rodriguez, Helander & Bhattacharjee Reference Rodriguez, Helander and Bhattacharjee2020). QS magnetic fields are further subdivided by their helicity: quasi-poloidal symmetry (QP), quasi-helical symmetry (QH) and quasi-axisymmetry (QA) are the quasi-symmetric subspaces of OP, OH and OT, respectfully. The Venn diagram in figure 1 visualizes the relationship between these different classes of magnetic fields.

Figure 1. Classification of magnetic fields. The decreasing area of the inner subsets represents the smaller dimensionality of these subspaces. Note that OP refers to the same set of magnetic fields that are sometimes labelled QI.

Since omnigenity is a less restrictive requirement than quasi-symmetry, it provides a larger design space with equivalent neoclassical confinement. This extra flexibility could help satisfy additional physics and engineering constraints on a stellarator reactor, such as reduced turbulent transport and simplified coil geometry. Unfortunately, Cary & Shasharina (Reference Cary and Shasharina1997a, Reference Cary and Shasharinab) proved that the only omnigenous fields that are analytic are those that are quasi-symmetric. Moreover, Garren & Boozer (Reference Garren and Boozer1991) suggested from a near-axis expansion that exact quasi-symmetry cannot be satisfied beyond a single flux surface. This is not a significant restriction in practice, however, QA and QH solutions have been found with good QS throughout a volume (Nührenberg & Zille Reference Nührenberg and Zille1988; Zarnstorff et al. Reference Zarnstorff, Berry, Brooks, Fredrickson, Fu, Hirshman, Hudson, Ku, Lazarus and Mikkelsen2001; Henneberg et al. Reference Henneberg, Drevlak, Nührenberg, Beidler, Turkin, Loizu and Helander2019; Bader et al. Reference Bader, Faber, Schmitt, Anderson, Drevlak, Duff, Frerichs, Hegna, Kruger and Landreman2020; Landreman & Paul Reference Landreman and Paul2022), and many solutions have been discovered that approach OP while being far from QS (Camacho Mata, Plunk & Jorge Reference Camacho Mata, Plunk and Jorge2022; Jorge et al. Reference Jorge, Plunk, Drevlak, Landreman, Lobsien, Camacho Mata and Helander2022; Sánchez et al. Reference Sánchez, Velasco, Calvo and Mulas2023; Goodman et al. Reference Goodman, Camacho Mata, Henneberg, Jorge, Landreman, Plunk, Smith, Mackenbach, Beidler and Helander2023). The performance of QH and OP have also been experimentally verified by the HSX and W7-X stellarators, respectively. These devices were optimized to have approximately quasi-symmetric or omnigenous designs, and both experiments have demonstrated reduced neoclassical transport (Canik et al. Reference Canik, Anderson, Anderson, Likin, Talmadge and Zhai2007; Beidler et al. Reference Beidler, Smith, Alonso, Andreeva, Baldzuhn, Beurskens, Borchardt, Bozhenkov, Brunner and Damm2021).

These recent results have been restricted to QS or OP magnetic fields, which excludes most of the full omnigenity solution space. Plunk, Landreman & Helander (Reference Plunk, Landreman and Helander2019) explain that OT and OH magnetic fields (that are not QS) cannot be achieved with a first-order expansion in the distance from the magnetic axis, but admit that these solutions could exist more generally. These classes of omnigenity have previously been neglected due to the lack of an efficient method to parametrize and explore their optimization space. Accessing this larger design space would unlock new possibilities for stellarators and is worthwhile to consider despite its increased computational complexity. This article presents an approach to achieve this goal, which is made possible through the DESC stellarator optimization code suite (Dudt & Kolemen Reference Dudt and Kolemen2020; Conlin et al. Reference Conlin, Dudt, Panici and Kolemen2023; Dudt et al. Reference Dudt, Conlin, Panici and Kolemen2023; Panici et al. Reference Panici, Conlin, Dudt and Kolemen2023; Dudt et al. Reference Dudt, Conlin, Panici, Unalmis, Kim and Kolemen2024). It is similar to the original idea of Cary & Shasharina (Reference Cary and Shasharina1997a), in the sense that a perfectly omnigenous magnetic field is constructed to have the same form on every field line in terms of some new coordinate. This omnigenous field is generalized to any helicity, as was also done by Landreman & Catto (Reference Landreman and Catto2012). The difference between the actual equilibrium magnetic field and the constructed target field is then penalized, like the technique devised by Goodman et al. (Reference Goodman, Camacho Mata, Henneberg, Jorge, Landreman, Plunk, Smith, Mackenbach, Beidler and Helander2023).

In contrast to previous approaches, this article provides a parametrization of the full phase space of all omnigenous magnetic fields, and is not restricted to a near-axis expansion formulation or to stellarator symmetry. Section 2 gives a derivation of the omnigenity model and explains how it is used as an optimization target. Numerical examples of equilibria obtained from this method are provided in § 3 for all six classes of omnigenity: OP, QP, OH, QH, OT and QA. This is the first demonstration of good neoclassical confinement throughout a volume for realistic equilibria with arbitrary omnigenous magnetic fields. The implications of these results and directions for future work are discussed in § 4.

2 Method

Let the ‘helicity’ of the omnigenous magnetic field be defined by the pair of integers $M$ and $N$ such that each $B$ contour closes on itself after traversing the torus $M$ times toroidally and $N$ times poloidally. In addition to having closed contours of magnetic field strength, the ‘bounce distance’ $\delta = \sqrt {\Delta \theta _B^2 + \Delta \zeta _B^2}$ in Boozer coordinates $(\theta _B,\zeta _B)$ along a field line between consecutive points with equal magnetic field strength must be independent of the field line label

(2.1)\begin{equation} \alpha = \frac{\theta_B - {\raise.3pt-\kern-5pt\iota} \zeta_B}{N - {\raise.3pt-\kern-5pt\iota} M}, \end{equation}

where ${\raise.3pt-\kern-5pt\iota}$ is the rotational transform. This normalization has the useful property that $\alpha$ increases by $2{\rm \pi}$ as a constant-$B$ curve is followed until it closes on itself (Landreman & Catto Reference Landreman and Catto2012).

2.1 Magnetic well

First, the "magnetic well shape" is parametrized. Assuming the magnetic field strength has the same maximum and minimum values along each field line, it can be written in the form

(2.2)\begin{equation} B=B(\rho,\eta), \end{equation}

where $\rho$ is the flux surface label and contours of constant magnetic field strength are now contours of the new coordinate $\eta$ (Cary & Shasharina Reference Cary and Shasharina1997b). The domain is chosen to be $-({{\rm \pi} }/{2}) \le \eta \le {{\rm \pi} }/{2}$ such that $B(0) = B_{\mathrm {min}}$ and $B(\pm {{\rm \pi} }/{2}) = B_{\mathrm {max}}$. The $B(\eta )$ should be monotonically increasing from $B_{\mathrm {min}}$ to $B_{\mathrm {max}}$ and satisfy the boundary conditions

(2.3)\begin{equation} \left.\frac{\partial B}{\partial\eta}\right|_{\eta=0} = \left.\frac{\partial B}{\partial\eta}\right|_{\eta={\pm}{\rm \pi}/{2}} = 0. \end{equation}

Additionally, $B$ must be an even function of $\eta$, which will be justified later. A magnetic well shape that satisfies these criteria is represented by the parameters

(2.4)\begin{equation} B_{ij} = B(\rho_i,\eta_j), \end{equation}

where the knot locations $\eta _j$ are linearly spaced in the domain $[0, {{\rm \pi} }/{2}]$ and the parameters are sorted to ensure monotonicity. The $B$ at other locations on the flux surface $\rho _i$ is evaluated by monotonic spline interpolation at $\eta =|\eta |$ to ensure it is an even function. The $C^1$ continuity of these monotonic splines is not a limitation since the omnigenous magnetic field is generally not analytic anyway. Omnigenity with multiple unique local minima and maxima of $B$ on each flux surface can exist, as shown by Parra et al. (Reference Parra, Calvo, Helander and Landreman2015), and this work could be extended to accommodate those cases.

2.2 Coordinate transformation

Now all that remains is to parametrize the transformation between the Boozer coordinates $(\theta _B,\zeta _B)$ and the computational coordinates $(\eta,\alpha )$. This mapping will be defined through the transformation

(2.5)\begin{equation} h : (\theta_B,\zeta_B) \leftrightarrow (\rho,\eta,\alpha). \end{equation}

In Boozer coordinates, the function $h$ is related to the helicity of the omnigenous field:

(2.6)\begin{equation} h(\theta_B,\zeta_B) = \begin{cases} N \zeta_B & \text{for }M=0, \\ -\theta_B + \dfrac{N}{M} \zeta_B & \text{for }M\ne0. \end{cases} \end{equation}

(Note that the Boozer angles are an implicit function of $\rho$, since the transformation to Boozer coordinates is unique to each flux surface.) Contours of constant $h$ are parallel to the $B_{\mathrm {max}}$ contour (which must be straight in Boozer coordinates) and in the limit of quasi-symmetry $B = B(h)$. In the computational coordinates, the function is parametrized as

(2.7)\begin{equation} h(\rho,\eta,\alpha) = 2\eta + {\rm \pi}+ \sum_{l=0}^{L_{\rho}} \sum_{m=0}^{M_{\eta}} \sum_{n={-}N_{\alpha}}^{N_{\alpha}} x_{lmn} T_{l}(2\rho-1) F_{m}(\eta) F_{n N_{\mathrm{FP}}}(\alpha), \end{equation}

where $N_{\mathrm {FP}}$ is the number of field periods, $T_{k}(\kern0.7pt y)$ are Chebyshev polynomials of the first kind defined as $T_{k}(\cos y) = \cos (k y)$ and $F_{k}(\kern0.7pt y)$ are Fourier series defined as

(2.8)\begin{equation} F_{k}(\kern0.7pt y) = \begin{cases} \cos(|k| y) & \text{for }k\ge0, \\ \sin(|k| y) & \text{for }k<0. \end{cases} \end{equation}

Chebyshev polynomials are chosen as the radial basis functions to allow for an independent well shape on each flux surface. (Any set of orthogonal polynomials would suffice. Zernike polynomials are not used since they would impose additional unphysical boundary conditions near the magnetic axis.) The coefficients $x_{lmn}$ in (2.7) parametrize the variation of the well shape on different field lines and $N_{\alpha }=0$ represents the subspace of quasi-symmetry. The boundary condition $h(\rho,\eta =-{{\rm \pi} }/{2},\alpha ) = 0$ is required to ensure $B_{\mathrm {max}}$ is a straight contour, which is chosen to be located at $h=0$ by construction. This boundary condition is enforced by a linear constraint on the even Fourier modes:

(2.9)\begin{equation} \sum_{m=0,2,4,\ldots}^{M_{\eta}} ({-}1)^{{m}/{2}}x_{lmn} = 0. \end{equation}

This constraint also satisfies the periodicity requirement that $h(\eta ={{\rm \pi} }/{2})=2{\rm \pi}$ is another $B_{\mathrm {max}}$ contour.

2.3 Optimization objective function

The optimization function penalizes the difference between the equilibrium magnetic field $B_{\mathrm {eq}}$ and the constructed omnigenous field $B$, similar to the approach used by Goodman et al. (Reference Goodman, Camacho Mata, Henneberg, Jorge, Landreman, Plunk, Smith, Mackenbach, Beidler and Helander2023). The errors are evaluated at a series of collocation points in the computational coordinates as

(2.10)\begin{equation} f_{\mathrm{om}}(\rho_i,\eta_i,\alpha_i) = \left[ B_{\mathrm{eq}}(\rho_i,\eta_i,\alpha_i) - B(\rho_i,\eta_i) \right] w(\eta_i), \end{equation}

where the weighting function $w(\eta _i) = \tfrac {3}{2} + \tfrac {1}{2} \cos (2\eta _i)$ is used to weight the minimum of the magnetic well relative to the maximum. The omnigenous target field $B$ can be readily evaluated at these collocation points from the magnetic well shape parameters $B_{ij}$ presented in (2.4). The equilibrium field $B_{\mathrm {eq}}$, however, must be evaluated at the Boozer coordinates that correspond to the desired computational coordinates, using the mapping defined by (2.5).

Simultaneously solving (2.1) and (2.6) gives the system of equations

(2.11a)\begin{gather} \begin{bmatrix} \alpha \\ h(\rho,\alpha,\eta) \end{bmatrix} = \begin{bmatrix} \dfrac{1}{N} & -\dfrac{{\raise.3pt-\kern-5pt\iota}}{N} \\ 0 & N \end{bmatrix} \begin{bmatrix} \theta_B \\ \zeta_B \end{bmatrix}\quad \text{for }M=0, N\neq0, \end{gather}
(2.11b)\begin{gather}\begin{bmatrix} \alpha \\ h(\rho,\alpha,\eta) \end{bmatrix} = \begin{bmatrix} \dfrac{-1}{M{\raise.3pt-\kern-5pt\iota}} & \dfrac{1}{M} \\ -1 & 0 \end{bmatrix} \begin{bmatrix} \theta_B \\ \zeta_B \end{bmatrix}\quad \text{for }M\neq0, N=0, \end{gather}
(2.11c)\begin{gather}\begin{bmatrix} \alpha \\ h(\rho,\alpha,\eta) \end{bmatrix} = \begin{bmatrix} \dfrac{1}{N-M{\raise.3pt-\kern-5pt\iota}} & \dfrac{-{\raise.3pt-\kern-5pt\iota}}{N-M{\raise.3pt-\kern-5pt\iota}} \\ -1 & \dfrac{N}{M} \end{bmatrix} \begin{bmatrix} \theta_B \\ \zeta_B \end{bmatrix} \quad \text{for }M\neq0, N\neq0, \end{gather}

which can be inverted to yield

(2.12a)\begin{gather} \begin{bmatrix} \theta_B \\ \zeta_B \end{bmatrix} = \begin{bmatrix} N & \dfrac{{\raise.3pt-\kern-5pt\iota}}{N} \\ 0 & \dfrac{1}{N} \end{bmatrix} \begin{bmatrix} \alpha \\ h(\rho,\alpha,\eta) \end{bmatrix} \quad \text{for }M=0, N\neq0, \end{gather}
(2.12b)\begin{gather}\begin{bmatrix} \theta_B \\ \zeta_B \end{bmatrix} = \begin{bmatrix} 0 & -1 \\ M & \dfrac{-1}{{\raise.3pt-\kern-5pt\iota}} \end{bmatrix} \begin{bmatrix} \alpha \\ h(\rho,\alpha,\eta) \end{bmatrix} \quad \text{for }M\neq0, N=0, \end{gather}
(2.12c)\begin{gather}\begin{bmatrix} \theta_B \\ \zeta_B \end{bmatrix} = \begin{bmatrix} N & \dfrac{M{\raise.3pt-\kern-5pt\iota}}{N-M{\raise.3pt-\kern-5pt\iota}} \\ M & \dfrac{M}{N-M{\raise.3pt-\kern-5pt\iota}} \end{bmatrix} \begin{bmatrix} \alpha \\ h(\rho,\alpha,\eta) \end{bmatrix} \quad \text{for }M\neq0, N\neq0. \end{gather}

The mapping parameters $x_{lmn}$ are used to compute $h(\rho,\eta,\alpha )$ through (2.7), and then (2.12) gives the corresponding Boozer coordinates such that $B_{\mathrm {eq}}(\rho,\eta,\alpha )$ can be evaluated in Boozer coordinates as $B_{\mathrm {eq}}(\theta _B,\zeta _B)$. These values of the equilibrium magnetic field strength are then used to compute the deviation from omnigenity in (2.10). Note that (2.12c) is undefined when ${\raise.3pt-\kern-5pt\iota} = {N}/{M}$, but this case is physically uninteresting because it corresponds to field lines being parallel to the $B_{\mathrm {max}}$ contour (for example, an axisymmetric vacuum field with no rotational transform). If $M=0$, then it must be the case that $N\neq 0$, so (2.12a) is also always well defined. See figure 2 for an illustration of how an omnigenous magnetic field is constructed and mapped to Boozer coordinates.

Figure 2. Example of an exact omnigenous magnetic field for an OP case with helicity $M=0, N=1$ and the parameters $x_{0,1,-1}=-{{\rm \pi} }/{8}, x_{0,1,0}={{\rm \pi} }/{8}, x_{0,1,+1}={{\rm \pi} }/{4}$: (a) well shape $B(\eta )$ given by a monotonic spline; (b) magnetic well along different field lines, which all have the same bounce distances $\Delta h$; (c) $B$ contours plotted in Boozer coordinates with a rotational transform of ${\raise.3pt-\kern-5pt\iota}={1}/{4}$.

2.4 Results

It can be shown that this model is guaranteed to yield a perfectly omnigenous target magnetic field. Since the straight field line coordinate system has the property $\Delta \theta _B = {\raise.3pt-\kern-5pt\iota}\Delta \zeta _B$ along a field line and ${\raise.3pt-\kern-5pt\iota}$ is constant on each flux surface, and also because $h(\theta _B,\zeta _B)$ is always a linear combination of the Boozer angles, the bounce distance is directly proportional to the change in $h$: $\delta \propto \Delta h$. The bounce points on any field line are always $\pm \eta$ by construction, since (2.2) was chosen to be an even function of $\eta$. Therefore the bounce distance can be computed as

(2.13a)\begin{align} \delta \propto \Delta h & = h(\rho,+\eta,\alpha) - h(\rho,-\eta,\alpha) \end{align}
(2.13b)\begin{align} & = 4\eta + \sum_{l=0}^{L_{\rho}} \sum_{m=0}^{M_{\eta}} \sum_{n={-}N_{\alpha}}^{N_{\alpha}} x_{lmn} \left[T_{l}(2\rho-1) F_{n N_{\mathrm{FP}}}(\alpha) \left( F_{m}(+\eta) - F_{m}(-\eta) \right) \right] \end{align}
(2.13c)\begin{align} & = 4\eta. \end{align}

In the last step, the Fourier series in $\eta$ cancel each other out because $\sum _{m\ge 0} F_{m}(\eta )$ is an even function of $\eta$. This ensures that the bounce distances set by the parameters $B_{ij}$ will be preserved across all field lines, regardless of the variation introduced by the parameters $x_{lmn}$. However, there are bounds on the allowable amplitudes of these coefficients to ensure (2.7) is a single-valued function.

Since the basis functions span the full domain, this model can approximate any omnigenous magnetic field in the limit of increasing resolution (although this work is restricted to fields with a single magnetic well). Whether a particular omnigenous target can be well approximated by an equilibrium solution is an unresolved question. The total number of coefficients used to describe an omnigenous magnetic field in (2.7) is $(L_\rho +1) (M_\eta +1) (2 N_\alpha + 1)$. Quasi-symmetry corresponds to the condition $N_\alpha =0$, so that the magnetic well shape is identical on all field lines and $B = B(h(\rho,\eta ))$. Assuming $L_\rho = M_\eta = N_\alpha$, the dimension of the full omnigenous parameter space scales as $\mathcal {O}(M_\eta ^3)$, while the quasi-symmetric subspace has a dimension of $\mathcal {O}(M_\eta ^2)$. This scaling reveals that quasi-symmetry is only a small set of the possible solutions with good particle confinement and motivates the discovery of stellarator equilibria in the much broader omnigenous design space. (This analysis is independent of the dimension of the magnetic well parametrization $B_{ij}$.) Stellarator symmetry imposes no constraints on the parameters $x_{lmn}$ due to the nonlinear nature of the coordinate mapping in (2.5). This implies that the size of the parameter space is not reduced by stellarator symmetry, but consequently no additional computational complexity is required if it is not assumed.

3 Examples

Numerical examples from each class of omnigenity are presented: the three types of quasi-symmetry (QP, QH and QA) and their more general omnigenous counterparts (OP, OH and OT, respectfully). These six solutions were obtained in DESC using the following optimization procedure. First, an initial guess was generated from the near-axis expansion code pyQIC (Jorge Reference Jorge2024) (for the OP and QP cases) or pyQSC (Landreman & Sengupta Reference Landreman and Sengupta2019; Landreman, Sengupta & Plunk Reference Landreman, Sengupta and Plunk2019) (for the other cases). It was evaluated at an aspect ratio of $R_0/a \approx 20$, since the near-axis expansions are most accurate at high aspect ratios. This configuration was re-solved in DESC while preserving the $\mathcal {O}(\rho )$ behaviour of the near-axis expansion solution. The equilibrium was then optimized using a multi-objective cost function of the form

(3.1)\begin{equation} \boldsymbol{f} = \begin{bmatrix} f_{\mathrm{om}}(\rho,\eta,\alpha) \\ \sigma f_{\mathrm{eq}}(\rho,\theta,\zeta) \end{bmatrix}, \end{equation}

where $f_{\mathrm {om}}(\rho,\eta,\alpha )$ are the omnigenity errors from (2.10), $f_{\mathrm {eq}}(\rho,\theta,\zeta )$ are the equilibrium force balance errors as described by Dudt & Kolemen (Reference Dudt and Kolemen2020) and $\sigma$ is a relative weighting between the two objective functions. This penalty method was used instead of the traditional optimization approach of re-solving the fixed-boundary equilibrium at each iteration (Dudt et al. Reference Dudt, Conlin, Panici and Kolemen2023) because it can better handle the near-axis constraints and is equivalent in the limit $\sigma \to \infty$. Omnigenity was targeted on five flux surfaces $\rho = 0.2, 0.4, 0.6, 0.8, 1.0$ and the shape of the magnetic well $B(\eta )$ on each surface was parametrized by a monotonic spline with eight knots. The force balance errors were evaluated throughout the plasma volume. The residuals in (3.1) were minimized with a least-squares trust region algorithm. The total magnetic flux, equilibrium profiles and $\mathcal {O}(\rho )$ behaviour from the near-axis expansion solution were constrained during the optimization. This process was repeated three times, with the numerical resolutions and $\sigma$ increasing at each iteration. Finally, the optimal configuration underwent a fixed-boundary solve to ensure its volume-averaged normalized force balance error was ${<}1\,\%$ (Panici et al. Reference Panici, Conlin, Dudt and Kolemen2023).

Each optimization took less than thirty minutes to run on a single NVIDIA A100 GPU. All of the examples have a major radius of $1\,\mathrm {m}$ and an average magnetic field strength on axis of $1\,\mathrm {T}$ by construction. The results are stellarator symmetric, but this was a design choice and not a limitation of the method. Figure 3 displays the magnetic field strength in Boozer coordinates on the $\rho =1$ surface for these equilibria. These plots clearly reveal the desired helicity of the (straight) $B_{\mathrm {max}}$ contours for each case, and they help picture how the bounce distances are roughly equal on every field line (satisfying the condition ${\partial \delta }/{\partial \alpha } = 0$). A scalar omnigenity error is defined as the mean of the local relative errors:

(3.2)\begin{equation} \frac{1}{K} \sum_{i=1}^K \frac{\left| B_{\mathrm{eq}}(\rho_i,\eta_i,\alpha_i) - B(\rho_i,\eta_i) \right|}{B(\rho_i,\eta_i)} \end{equation}

and the normalized quasi-symmetry error in Boozer coordinates is defined as (Rodriguez, Paul & Bhattacharjee Reference Rodriguez, Paul and Bhattacharjee2022)

(3.3)\begin{equation} \frac{\sqrt{\sum_{{n}/{m}\neq{N}/{M}} B^2_{mn}}}{\sqrt{\sum_{m,n}B^2_{mn}}}. \end{equation}

These errors along with the aspect ratio and maximum elongation of each solution are provided in table 1. The aspect ratio was indirectly constrained during the optimization by fixing the magnetic axis shape, magnetic field strength on axis, and total magnetic flux; the elongation was not constrained.

Figure 3. Contours of constant magnetic field magnitude in Boozer coordinates at the boundary surface $\rho =1$ for a single field period. The magnetic field line at $\alpha =0$ is shown as a black dashed line in panels (af). The insets show a three-dimensional view of the field strength on each surface.

Table 1. Comparison between the six omnigenous solutions of geometric parameters and magnetic field errors at the $\rho =1$ surface. The aspect ratios are calculated using the same formula as the VMEC equilibrium code (Hirshman & Whitson Reference Hirshman and Whitson1983). The omnigenity errors were computed with different collocation points than those used in the optimization.

As another measure of the levels of omnigenity achieved, the effective ripple $\epsilon _{\mathrm {eff}}^{3/2}$ – the magnitude of the neoclassical collisional transport in the ${1}/{\nu }$ regime, where $\nu$ is the collision frequency – is computed by NEO (Nemov et al. Reference Nemov, Kasilov, Kernbichler and Heyn1999) for each solution. As another measure of confinement, the time history of fusion-born alpha particles lost in a reactor scale device is also computed by SIMPLE (Albert, Kasilov & Kernbichler Reference Albert, Kasilov and Kernbichler2020). In a perfectly omnigenous magnetic field, $\epsilon _{\mathrm {eff}}^{3/2}$ should be zero and no particles should be lost besides those with banana orbits that are wide enough to leave the plasma volume. Both of these confinement measures are plotted in figure 4. The numerical examples shown here were only optimized for omnigenity and equilibrium quality, and other factors that may be important for a fusion reactor design were not considered. More details about each optimized solution are provided in the following subsections.

Figure 4. Measures of confinement. The quasi-symmetric solutions are plotted as dotted lines, while the more general omnigenous solutions are plotted as solid lines. Solutions with the same type of helicity are plotted in the same colour. The Wendelstein 7-X configuration with $\beta =4\,\%$ is plotted in a black dashed line for comparison (Landreman Reference Landreman2021). (a) Neoclassical collisional transport magnitude, computed by NEO. (b) Collisionless losses of fusion-born alpha particles initialized at $\rho =0.5$, computed by SIMPLE. All configurations were scaled to the minor radius and magnetic field strength of ARIES-CS.

3.1 Poloidal contours

Omnigenity with poloidally closed $B$ contours is characterized by a helicity of $M=0$. These examples of OP and QP are vacuum equilibria ($\langle \beta \rangle =0$) that were based on initial guesses generated by pyQIC. The OP case is a two field period configuration with a mirror ratio of $\varDelta =0.13$ and the QP case has a single field period with $\varDelta =0.17$, where the mirror ratio is defined as

(3.4)\begin{equation} \varDelta = \frac{B_\mathrm{max}-B_\mathrm{min}}{B_\mathrm{max}+B_\mathrm{min}} \end{equation}

on the magnetic axis. In addition to the flux surface geometry, the omnigenity parameters $B_{ij}$ and $x_{lmn}$ were also considered free variables during the optimization. A resolution of $L_\rho = M_\eta = N_\alpha = 4$ was used for the OP case, while the QP case used a resolution of $L_\rho = M_\eta = 4$, $N_\alpha = 0$ to enforce quasi-symmetry. The OP solution has worse QS error than the QP example, as expected, but achieves an order of magnitude lower omnigenity error and effective ripple, and two orders of magnitude lower fast particle losses (see table 1 and figure 4).

3.2 Helical contours

The typical case of omnigenity with helically closed $B$ contours is $M=1$ and $N=N_{\mathrm {FP}}$. These examples of OH and QH are five field period vacuum equilibria ($\langle \beta \rangle =0$) using the same initial guess from pyQSC. As in the previous poloidal examples, the parameters $B_{ij}$ describing the magnetic well shapes on each surface were considered free optimization variables. However, in these runs, the coordinate mapping of (2.7) was prescribed: $x_{lmn}=0\, \forall l,m,n$ in the QH case, and $x_{0,1,+1}=-{{\rm \pi} }/{6}$ was the only non-zero parameter in the OH case. This parameter value for the OH example was chosen such that the resulting field, shown in figure 3(c), qualitatively resembles the $M=1$, $N=4$ omnigenous field example constructed by Landreman & Catto (Reference Landreman and Catto2012). In both cases, all of the $x_{lmn}$ variables were held fixed during the optimizations to ensure that they converged to a quasi-symmetric and general omnigenous magnetic field, respectively. The QS error of the OH solution is an order of magnitude higher than that of the QH result, verifying that it is far from quasi-symmetry (see table 1).

3.3 Toroidal contours

Here, $N=0$ describes omnigenity with toroidally closed $B$ contours. As explained by Landreman & Catto (Reference Landreman and Catto2012), "it is difficult to construct $N=0$ omnigenous fields that depart strongly from quasi-symmetry $\ldots$ because the $B$ contours in a $N=0$ quasi-symmetric field are already nearly parallel to the field lines when ${\raise.3pt-\kern-5pt\iota}/N_{\mathrm {FP}}<1$". These examples of OT and QA are single field period equilibria based on the same initial guess from pyQSC, with a finite toroidal current profile and a volume averaged normalized plasma pressure of $\langle \beta \rangle =1.2\,\%$. The plasma current provides additional rotational transform, which allows for an OT case that is far from quasi-symmetry (as seen in figure 3b and by the relatively high QS error in table 1). These toroidal solutions were found using the same optimization process as the previous helical examples, except that the coordinate mapping for the target OT field had $x_{0,1,-1}={{\rm \pi} }/{6}$ as the only non-zero parameter. As in the helical examples, all of the $x_{lmn}$ variables were constrained to their target values while all of the $B_{ij}$ were free variables during the optimization.

4 Discussion

This is the first demonstration of equilibria solutions that closely approximate every class of omnigenous magnetic fields. All of the $B$ contour plots in figure 3 have their expected helicities and it can be seen that the distances between equal contours are roughly equivalent along any field line. The average relative omnigenity errors listed in table 1 are all $\lessapprox 1\,\%$, indicating the optimized equilibria closely resemble their omnigenous targets. The good confinement properties of these omnigenous solutions at moderate aspect ratio are verified by the neoclassical collisional transport magnitudes $\epsilon _{\mathrm {eff}}^{3/2}$ and the loss fractions displayed in figure 4. These relatively low values of both confinement metrics are all comparable in magnitude to previous optimized configurations (Landreman & Paul Reference Landreman and Paul2022). In particular, the OT solution has very low effective ripple, and the OP and OH solutions have almost no alpha particle losses – on par with the QS results. This indicates the existence of stellarator equilibria other than OP that can approach omnigenity without being quasi-symmetric.

The OH and OT results do have higher omnigenity errors and worse confinement than their respective quasi-symmetric counterparts, but the target parameters for these examples were chosen arbitrarily and other omnigenous fields with better confinement properties might exist. Omnigenous fields that are not quasi-symmetric are impossible to achieve exactly, and more examples are needed to understand if they can generally be approximated well. The OT result also has surprisingly high alpha particle losses considering its very low effective ripple, exemplifying the complicated relationship between these two confinement metrics. This could be explained by wide banana orbits that extend beyond the plasma volume, but more investigation is required to determine the cause. The outlier of all the results is the QP solution, which cannot exist near the magnetic axis (Plunk et al. Reference Plunk, Landreman and Helander2019) and is known to be difficult to achieve far from the axis without highly elongated surfaces. This QP example may have unrealistic geometry for a reactor, but it is still encouraging to find a quasi-poloidally symmetric configuration with similar levels of confinement as the Wendelstein 7-X experiment.

The purpose of the examples in § 3 is to establish solutions that are representative of each type of omnigenity that can exist. The model presented in § 2 and the optimization capabilities of DESC provide flexibility for a variety of applications. The parameters $B_{ij}$ and $x_{lmn}$ can either be constrained to target a specific location in the omnigenity design space (as in the OH, OT, QH and QA examples), or they can be free optimization variables (as in the OP and QP examples). Providing more degrees of freedom would likely result in the discovery of configurations with even better confinement properties or that satisfy additional optimization objectives. Eliminating the assumption of stellarator symmetry is a logical step in this direction, as it doubles the size of the omnigenity solution space that can be explored. As another example, the boundary condition of (2.9) could be relaxed and only satisfied approximately; this would violate the already impossible goal of exact omnigenity, but might improve the solution quality overall. Alternatively, constraining most of the parameters would enable a systematic scan of the full omnigenity design space. Mapping this space, like the surveys of quasi-symmetry made by Landreman (Reference Landreman2022) and Rodríguez, Sengupta & Bhattacharjee (Reference Rodríguez, Sengupta and Bhattacharjee2023), is an exciting direction for future research. Other ideas for future work include investigating how the larger omnigenity phase space (in contrast to QS) could simplify stellarator coil geometry and yield more promising candidates for fusion power plants.

Acknowledgements

The code and data that support the findings of this study are openly available in the DESC repository at https://github.com/PlasmaControl/DESC. J.L. Velasco & collaborators have presented a similar method to parametrize a subset of omnigenous magnetic fields with poloidal contours, and they are thanked for their contributions.

Editor P. Helander thanks the referees for their advice in evaluating this article.

Funding

This work was supported by the U.S. Department of Energy under contract numbers DE-AC02-09CH11466, DE-SC0022005, and Field Work Proposal No. 1019. The United States Government retains a non-exclusive, paid-up, irrevocable, world-wide license to publish or reproduce the published form of this manuscript, or allow others to do so, for United States Government purposes.

Declaration of interests

The authors report no conflict of interest.

References

Albert, C.G., Kasilov, S.V. & Kernbichler, W. 2020 Accelerated methods for direct computation of fusion alpha particle losses within, stellarator optimization. J. Plasma Phys. 86, 815860201.CrossRefGoogle Scholar
Bader, A., Faber, B.J., Schmitt, J.C., Anderson, D.T., Drevlak, M., Duff, J.M., Frerichs, H., Hegna, C.C., Kruger, T.G., Landreman, M., et al. 2020 Advancing the physics basis for quasi-helically symmetric stellarators. J. Plasma Phys. 86, 905860506.CrossRefGoogle Scholar
Beidler, C.D., Smith, H.M., Alonso, A., Andreeva, T., Baldzuhn, J., Beurskens, M.N.A., Borchardt, M., Bozhenkov, S.A., Brunner, K.J., Damm, H., et al. 2021 Demonstration of reduced neoclassical energy transport in Wendelstein 7-X. Nature 596, 221226.CrossRefGoogle ScholarPubMed
Camacho Mata, K., Plunk, G.G. & Jorge, R. 2022 Direct construction of stellarator-symmetric quasi-isodynamic magnetic configurations. J. Plasma Phys. 88, 905880503.CrossRefGoogle Scholar
Canik, J.M., Anderson, D.T., Anderson, F.S.B., Likin, K.M., Talmadge, J.N. & Zhai, K. 2007 Experimental demonstration of improved neoclassical transport with quasihelical symmetry. Phys. Rev. Lett. 98, 085002.CrossRefGoogle ScholarPubMed
Cary, J.R. & Shasharina, S.G. 1997 a Helical plasma confinement devices with good confinement properties. Phys. Rev. Lett. 78, 674677.CrossRefGoogle Scholar
Cary, J.R. & Shasharina, S.G. 1997 b Omnigenity and quasihelicity in helical plasma confinement systems. Phys. Plasmas 4, 3323.CrossRefGoogle Scholar
Conlin, R., Dudt, D.W., Panici, D. & Kolemen, E. 2023 The desc stellarator code suite. Part 2. Perturbation and continuation methods. J. Plasma Phys. 89, 955890305.CrossRefGoogle Scholar
Dudt, D.W., Conlin, R., Panici, D. & Kolemen, E. 2023 The desc stellarator code suite. Part 3. Quasi-symmetry optimization. J. Plasma Phys. 89, 955890201.CrossRefGoogle Scholar
Dudt, D.W., Conlin, W., Panici, D., Unalmis, K., Kim, P. & Kolemen, E. 2024 DESC. https://github.com/PlasmaControl/DESC.Google Scholar
Dudt, D.W. & Kolemen, E. 2020 DESC: a stellarator equilibrium solver. Phys. Plasmas 27, 102513.CrossRefGoogle Scholar
Garren, D.A. & Boozer, A.H. 1991 Existence of quasihelically symmetric stellarators. Phys. Fluids B 3, 2822.CrossRefGoogle Scholar
Goodman, A.G., Camacho Mata, K., Henneberg, S.A., Jorge, R., Landreman, M., Plunk, G.G., Smith, H.M., Mackenbach, R.J.J., Beidler, C.D. & Helander, P. 2023 Constructing precisely quasi-isodynamic magnetic fields. J. Plasma Phys. 89 (5), 905890504.CrossRefGoogle Scholar
Hall, L.S. & McNamara, B. 1975 Three-dimensional equilibrium of the anisotropic, finite-pressure guiding-center plasma: theory of the magnetic plasma. Phys. Fluids 18, 552565.CrossRefGoogle Scholar
Helander, P. 2014 Theory of plasma confinement in non-axisymmetric magnetic fields. Rep. Prog. Phys. 77, 087001.CrossRefGoogle ScholarPubMed
Helander, P., Beidler, C.D., Bird, T.M., Drevlak, M., Feng, Y., Hatzky, R., Jenko, F., Kleiber, R., Proll, J.H.E., Turkin, Y., et al. 2012 Stellarator and tokamak plasmas: a comparison. Plasma Phys. Control. Fusion 54, 124009.CrossRefGoogle Scholar
Henneberg, S., Drevlak, M., Nührenberg, C., Beidler, C.D., Turkin, Y., Loizu, J. & Helander, P. 2019 Properties of a new quasi-axisymmetric configuration. Nucl. Fusion 59, 026014.CrossRefGoogle Scholar
Hirshman, S.P. & Whitson, J.C. 1983 Steepest-descent moment method for three-dimensional magnetohydrodynamic equilibria. Phys. Fluids 26, 35533568.CrossRefGoogle Scholar
Isaev, M.Y., Mikhailov, M.I., Monticello, D.A., Mynick, H.E., Subbotin, A.A., Ku, L.P. & Reiman, A.H. 1999 The pseudo-symmetric optimization of the national compact stellarator experiment. Phys. Plasmas 6, 31743179.CrossRefGoogle Scholar
Jorge, R., Plunk, G.G., Drevlak, M., Landreman, M., Lobsien, J.-F., Camacho Mata, K. & Helander, P. 2022 A single-field-period quasi-isodynamic stellarator. J. Plasma Phys. 88, 175880504.CrossRefGoogle Scholar
Landreman, M. 2021 Data for the paper “magnetic fields with precise quasisymmetry”.CrossRefGoogle Scholar
Landreman, M. 2022 Mapping the space of quasisymmetric stellarators using optimized near-axis expansion. J. Plasma Phys. 88, 905880616.CrossRefGoogle Scholar
Landreman, M. & Catto, P.J. 2012 Omnigenity as generalized quasisymmetry. Phys. Plasmas 19, 056103.CrossRefGoogle Scholar
Landreman, M. & Paul, E. 2022 Magnetic fields with precise quasisymmetry for plasma confinement. Phys. Rev. Lett. 128, 035001.CrossRefGoogle ScholarPubMed
Landreman, M. & Sengupta, W. 2019 Constructing stellarators with quasisymmetry to high order. J. Plasma Phys. 85, 815850601.CrossRefGoogle Scholar
Landreman, M., Sengupta, W. & Plunk, G.G. 2019 Direct construction of optimized stellarator shapes. Part 2. Numerical quasisymmetric solutions. J. Plasma Phys. 85, 905850103.CrossRefGoogle Scholar
Nemov, V.V., Kasilov, S.V., Kernbichler, W. & Heyn, M.F. 1999 Evaluation of $1/\nu$ neoclassical transport in stellarators. Phys. Plasmas 6, 46224632.CrossRefGoogle Scholar
Nührenberg, J. & Zille, R. 1988 Quasi-helically symmetric toroidal stellarators. Phys. Lett. A 129, 113117.CrossRefGoogle Scholar
Palumbo, D. 1968 Some considerations on closed configurations of magnetohydrostatic equilibrium. Nuovo Cimento B 53, 507511.CrossRefGoogle Scholar
Panici, D., Conlin, R., Dudt, D.W. & Kolemen, E. 2023 The desc stellarator code suite. Part 1. Quick and accurate equilibria computations. J. Plasma Phys. 89, 955890303.CrossRefGoogle Scholar
Parra, F.I., Calvo, I., Helander, P. & Landreman, M. 2015 Less constrained omnigeneous stellarators. Nucl. Fusion 55, 033005.CrossRefGoogle Scholar
Plunk, G.G., Landreman, M. & Helander, P. 2019 Direct construction of optimized stellarator shapes. Part 3. Omnigenity near the magnetic axis. J. Plasma Phys. 85, 905850602.CrossRefGoogle Scholar
Rodríguez, E., Sengupta, W. & Bhattacharjee, A. 2023 Constructing the space of quasisymmetric stellarators through near-axis expansion. Plasma Phys. Control. Fusion 65, 095004.Google Scholar
Rodriguez, E., Helander, P. & Bhattacharjee, A. 2020 Necessary and sufficient conditions for quasisymmetry. Phys. Plasmas 27, 062501.CrossRefGoogle Scholar
Rodriguez, E., Paul, E.J. & Bhattacharjee, A. 2022 Measures of quasisymmetry for stellarators. J. Plasma Phys. 88, 905880109.CrossRefGoogle Scholar
Sánchez, E., Velasco, J.L., Calvo, I. & Mulas, S. 2023 A quasi-isodynamic configuration with good confinement of fast ions at low plasma.CrossRefGoogle Scholar
Spitzer, L. 1958 The stellarator concept. Phys. Fluids 1, 253-264.CrossRefGoogle Scholar
Zarnstorff, M.C., Berry, L.A., Brooks, A., Fredrickson, E., Fu, G.-Y., Hirshman, S., Hudson, S., Ku, L.-P., Lazarus, E., Mikkelsen, D., et al. 2001 Physics of the compact advanced stellarator NCSX. Plasma Phys. Control. Fusion 43, A237A249.CrossRefGoogle Scholar
Figure 0

Figure 1. Classification of magnetic fields. The decreasing area of the inner subsets represents the smaller dimensionality of these subspaces. Note that OP refers to the same set of magnetic fields that are sometimes labelled QI.

Figure 1

Figure 2. Example of an exact omnigenous magnetic field for an OP case with helicity $M=0, N=1$ and the parameters $x_{0,1,-1}=-{{\rm \pi} }/{8}, x_{0,1,0}={{\rm \pi} }/{8}, x_{0,1,+1}={{\rm \pi} }/{4}$: (a) well shape $B(\eta )$ given by a monotonic spline; (b) magnetic well along different field lines, which all have the same bounce distances $\Delta h$; (c) $B$ contours plotted in Boozer coordinates with a rotational transform of ${\raise.3pt-\kern-5pt\iota}={1}/{4}$.

Figure 2

Figure 3. Contours of constant magnetic field magnitude in Boozer coordinates at the boundary surface $\rho =1$ for a single field period. The magnetic field line at $\alpha =0$ is shown as a black dashed line in panels (af). The insets show a three-dimensional view of the field strength on each surface.

Figure 3

Table 1. Comparison between the six omnigenous solutions of geometric parameters and magnetic field errors at the $\rho =1$ surface. The aspect ratios are calculated using the same formula as the VMEC equilibrium code (Hirshman & Whitson 1983). The omnigenity errors were computed with different collocation points than those used in the optimization.

Figure 4

Figure 4. Measures of confinement. The quasi-symmetric solutions are plotted as dotted lines, while the more general omnigenous solutions are plotted as solid lines. Solutions with the same type of helicity are plotted in the same colour. The Wendelstein 7-X configuration with $\beta =4\,\%$ is plotted in a black dashed line for comparison (Landreman 2021). (a) Neoclassical collisional transport magnitude, computed by NEO. (b) Collisionless losses of fusion-born alpha particles initialized at $\rho =0.5$, computed by SIMPLE. All configurations were scaled to the minor radius and magnetic field strength of ARIES-CS.