Hostname: page-component-848d4c4894-8bljj Total loading time: 0 Render date: 2024-06-22T09:23:13.722Z Has data issue: false hasContentIssue false

Measures of quasisymmetry for stellarators

Published online by Cambridge University Press:  28 January 2022

E. Rodríguez*
Affiliation:
Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08543, USA Princeton Plasma Physics Laboratory, Princeton, NJ 08540, USA
E.J. Paul
Affiliation:
Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08543, USA Princeton Plasma Physics Laboratory, Princeton, NJ 08540, USA
A. Bhattacharjee
Affiliation:
Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08543, USA Princeton Plasma Physics Laboratory, Princeton, NJ 08540, USA
*
Email address for correspondence: eduardor@princeton.edu
Rights & Permissions [Opens in a new window]

Abstract

Quasisymmetric stellarators are an attractive class of optimised magnetic confinement configurations. The property of quasisymmetry (QS) is in practice limited to be approximate, and thus the construction requires measures that quantify the deviation from the exact property. In this paper we study three measure candidates used in the literature, placing the focus on their origin and a comparison of their forms. The analysis shows clearly the lack of universality in these measures. As these metrics do not directly correspond to any physical property (except when exactly quasisymmetric), optimisation should employ additional physical metrics for guidance. It is suggested that close to QS minima, one should treat QS metrics through inequality constraints so that additional physics metrics dominate optimisation. The impact of different quasisymmetric measures on optimisation is presented through an example, for which the standard metric that weights the asymmetric Fourier modes of the field magnitude appears to perform best.

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 (https://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), 2022. Published by Cambridge University Press

1 Introduction

The stellarator concept (Spitzer Reference Spitzer1958) has seen a revival in the last few decades as an option to achieve controlled thermonuclear fusion. Stellarators are three-dimensional magnetic field configurations which confine hot plasmas within. The three-dimensionality of these devices provides the necessary freedom to avoid some of the limiting features from which axisymmetric tokamaks suffer: most notably, current-driven instabilities. However, this freedom can come at a price: the three-dimensionality of the magnetic field might lead to a loss of confinement that continuous symmetry confers upon a tokamak.

Quasisymmetry (QS) (Boozer Reference Boozer1983; Nührenberg & Zille Reference Nührenberg and Zille1988; Tessarotto et al. Reference Tessarotto, Johnson, White and Zheng1996; Helander Reference Helander2014; Burby, Kallinikos & MacKay Reference Burby, Kallinikos and MacKay2020; Rodríguez, Helander & Bhattacharjee Reference Rodríguez, Helander and Bhattacharjee2020) is a property of the field that attempts to preserve in three-dimensional stellarators the good neoclassical properties (Mynick Reference Mynick2006) of the tokamak. The promise of this property has led to a search of designs that bear it. Although theoretical work suggests that in an ideal, static equilibrium with isotropic pressure one may not construct exact QS solutions (Garren & Boozer Reference Garren and Boozer1991; Landreman & Sengupta Reference Landreman and Sengupta2018; Rodríguez & Bhattacharjee Reference Rodríguez and Bhattacharjee2021b), one may design such configurations approximately. In practice there have been two main approaches to finding approximately quasisymmetric equilibria. One is to directly construct approximate solutions by employing an asymptotic expansion near the magnetic axis (Garren & Boozer Reference Garren and Boozer1991; Landreman & Sengupta Reference Landreman and Sengupta2018, Reference Landreman and Sengupta2019; Rodríguez & Bhattacharjee Reference Rodríguez and Bhattacharjee2021c) or near axisymmetry (Plunk & Helander Reference Plunk and Helander2018). The other, and most commonly employed, approach is to treat the construction of QS configurations as part of an optimisation problem (Bader et al. Reference Bader, Drevlak, Anderson, Faber, Hegna, Likin, Schmitt and Talmadge2019; Henneberg et al. Reference Henneberg, Drevlak, Nuhrenberg, Beidler, Turkin, Loizu and Helander2019).

To proceed with optimisation it is necessary to construct cost functions that penalise deviations from QS. This has been done in a number of contexts and forms (Bader et al. Reference Bader, Drevlak, Anderson, Faber, Hegna, Likin, Schmitt and Talmadge2019; Henneberg et al. Reference Henneberg, Drevlak, Nuhrenberg, Beidler, Turkin, Loizu and Helander2019; Paul et al. Reference Paul, Antonsen, Landreman and Cooper2020). However, at the present time there appears to be a lack of comparative studies of the different forms of cost functions used. The purpose of this paper is to re-evaluate carefully the origin of some of these commonly used forms and to investigate their differences, significance and content.

The paper is organised as follows. In § 2 three different measures of QS are considered. From these basics, the most apparent formal aspects of the metrics are discussed. In § 3 the measures are formally compared on the same footing. The connection between physical properties associated with QS and the form of the metrics is then explored. Section 4 presents a more practical perspective on the measures by looking at the universality of QS measures (how they compare across different near-QS designs) and the importance of the cost functions for optimisation in practice. We close the paper with some concluding remarks and suggestions for future optimisation.

2 Measures of QS

Let us start by defining the concept of QS from the fundamental perspective of single-particle dynamics. The class of magnetic fields that grants an approximate conserved momentum to the gyrocentre dynamics of charged particles is called quasi-symmetric. In particular, we focus our attention on weakly quasisymmetric magnetic fields as recently defined (Rodríguez et al. Reference Rodríguez, Helander and Bhattacharjee2020; Constantin, Drivas & Ginsberg Reference Constantin, Drivas and Ginsberg2021). Formally, a magnetic field $\boldsymbol {B}$ is weakly quasisymmetric if and only if there exists a symmetry vector field $\boldsymbol {u}$ such that (Rodríguez et al. Reference Rodríguez, Helander and Bhattacharjee2020)

(2.1)\begin{gather} \boldsymbol{u}\boldsymbol{\cdot}\boldsymbol{\nabla} B=0, \end{gather}
(2.2)\begin{gather}\boldsymbol{B}\times\boldsymbol{u}=\boldsymbol{\nabla}\varPhi, \end{gather}
(2.3)\begin{gather}\boldsymbol{\nabla}\boldsymbol{\cdot}\boldsymbol{u}=0, \end{gather}

where $\varPhi$ labels magnetic flux surfaces that we take to be nested (Rodríguez & Bhattacharjee Reference Rodríguez and Bhattacharjee2021a). We are assuming the electrostatic potential in the problem to be symmetric, in the sense that $\boldsymbol {u}\boldsymbol {\cdot }\boldsymbol {\nabla } \varLambda =0$ for $\varLambda$ the electrostatic potential. This symmetry requirement preserves the approximate conservation of the momentum characterising a QS configuration.

Although this definition has the benefit of directly emerging from single-particle considerations, it is difficult to use it practically to assess whether a field is quasisymmetric or not. Doing so would require solving for the field $\boldsymbol {u}$. Instead, we would like to have a condition that only involves magnetic field quantities directly. This can be achieved in a straightforward way by constructing $\boldsymbol {u}$ such that it satisfies (2.1) and (2.2): $\boldsymbol {u}=\boldsymbol {\nabla }\varPhi \times \boldsymbol {\nabla } B/(\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B)$. At the singular points where $\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B= 0$ this construction breaks down unless $\boldsymbol {\nabla }\varPhi \times \boldsymbol {\nabla } B\boldsymbol {\cdot }\boldsymbol {B}=0$. (Note that this follows from the conditions (2.1) and (2.2).) The vanishing of $\boldsymbol {\nabla }\varPhi \times \boldsymbol {\nabla } B\boldsymbol {\cdot }\boldsymbol {B}$ when $\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B= 0$ is referred to as pseudosymmetry (Mikhailov et al. Reference Mikhailov, Shafranov, Subbotin, Isaev, Hrenberg, Zille and Cooper2002; Skovoroda Reference Skovoroda2005), and it guarantees that all of the contours of constant $|\boldsymbol {B}|$ on a flux surface are linked to the torus in the same way, avoiding local extrema on the surfaces (and thus providing well-behaved streamlines of $\boldsymbol {u}$). Breaking pseudosymmetry leads to a diverging orbit width (Rodríguez et al. Reference Rodríguez, Helander and Bhattacharjee2020) (i.e. loss of deeply trapped and barely trapped particles) at these $|\boldsymbol {B}|$ extrema. In a QS solution, the construction of $\boldsymbol {u}$ is then well defined because QS implies pseudosymmetry.

With the given construction of $\boldsymbol {u}$, we may substitute it into (2.3) to obtain

(2.4)\begin{equation} \frac{\boldsymbol{\nabla}\varPhi\times\boldsymbol{\nabla} B\boldsymbol{\cdot}\boldsymbol{\nabla}(\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} B)}{(\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} B)^{2}}=0. \end{equation}

This yields the triple vector formulation of QS. Namely, if we define the scalar quantity

(2.5)\begin{equation} f_T=\boldsymbol{\nabla}\psi\times\boldsymbol{\nabla} B\boldsymbol{\cdot}\boldsymbol{\nabla}(\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} B), \end{equation}

the magnetic field is quasisymmetric if and only if $f_T=0$. (By continuity, this includes pseudosymmetry.)

For the purpose of this paper, all magnetic configurations have well-behaved nested flux surfaces and are non-isodynamic. By isodynamic we mean a field whose magnitude $|\boldsymbol {B}|$ is a flux function (Bernardin, Moses & Tataronis Reference Bernardin, Moses and Tataronis1986). With these provisions, $f_T$ can be evaluated for non-quasisymmetric configurations and serves as a measure of QS. We have here used $\psi$, the toroidal magnetic flux, as a label of flux surfaces rather than the arbitrary flux function $\varPhi (\psi )$. This is one of many possible choices, and one could, for instance, replace $\boldsymbol {\nabla } \psi$ with $\hat {\boldsymbol {n}}$ (the normal unit vector to the flux surface) in evaluating (2.5) on a magnetic surface within a region of continuously nested surfaces.

The measure $f_T$ is a coordinate-independent, scalar measure of the departure from QS. It is a local quantity, although it also has some global properties. For instance, $\langle f_T\rangle =0$, where $\langle \cdots \rangle$ is the flux-surface average.

We have seen how (2.5) results from the fundamental definition of QS in (2.1)–(2.3). This grants $f_T$ the special feature that its application is not tied to a particular form of equilibrium. This could be relevant for constructing cost functions in so-called one-shot optimisation schemes (Akçelik et al. Reference Akçelik, Biros, Ghattas, Hill, Keyes, van Bloemen Waanders, Heroux, Raghavan and Simon2006; Dekeyser, Reiter & Baelmans Reference Dekeyser, Reiter and Baelmans2014), i.e. a formulation in which finding equilibria is part of the optimisation problem. Although QS may be explored in the presence of forces more general than isotropic plasma pressure, $p$, it is standard to adopt the magnetohydrostatic assumption $\boldsymbol {j}\times \boldsymbol {B}=\boldsymbol {\nabla } p$, where $\boldsymbol {j} = \boldsymbol {\nabla } \times \boldsymbol {B}$ is the current density. Following standard practice, we shall do so for the remainder of this work.

A key consequence of this form of equilibrium is that $\boldsymbol {j}\boldsymbol {\cdot }\boldsymbol {\nabla }\psi =0$. Thus, there exist Boozer coordinates (Boozer Reference Boozer1981; Helander Reference Helander2014) $\{\psi,\theta,\phi \}$. In these straight-field line coordinates the covariant form of the magnetic field is $\boldsymbol {B}=I(\psi )\boldsymbol {\nabla }\theta +G(\psi )\boldsymbol {\nabla }\phi +B_\psi \boldsymbol {\nabla }\psi$ and the coordinate Jacobian $\mathcal {J}=(G+\iota I)/B^{2}$, where $\iota$ is the rotational transform. Equation (2.5) may then be written explicitly in the form

(2.6)\begin{equation} f_T= \left[\partial_\theta B\partial_\phi-\partial_\phi B\partial_\theta\right](\partial_\phi+\iota\partial_\theta)B. \end{equation}

The QS condition can be rewritten in the form $(\partial _\theta B) (\partial _\phi +\iota \partial _\theta )(\partial _\phi B/\partial _\theta B)=0$, from which it follows that $\partial _\phi B/\partial _\theta B$ must be a flux function. This implies $|\boldsymbol {B}|$ has an explicit symmetry and the conventional definition of QS then follows (Boozer Reference Boozer1983; Rodríguez, Sengupta & Bhattacharjee Reference Rodríguez, Sengupta and Bhattacharjee2021). We refer to it as the Boozer formulation of QS: a magnetic field is quasisymmetric if and only if the magnetic field magnitude can be written as a function $B=B(\psi,M\theta -N\phi )$, where $N,M\in \mathbb {Z}$. It is convenient to define the symmetry helicity $\tilde {\alpha }=N/M$ and a helical angle $\chi =\theta -\tilde {\alpha }\phi$. (We will not consider quasipoloidal symmetry, corresponding to $M=0$, although the extension should be straightforward.) There is no unique way to construct a scalar measure that quantifies how close a given magnetic field is to QS in the presented formulation. It is, however, customary to define the minimal measure

(2.7)\begin{equation} f_B={\sum_{\substack{n,m \\ n\neq\tilde{\alpha}m}}}|B_{nm}|^{2}. \end{equation}

Here $B_{nm}$ stands for the Fourier components of the magnetic field magnitude $B$ in Boozer coordinates, and the sum is over the non-symmetric components. All the mode components are weighted equally in this form (although this could be modified). In practice, only the lower-order modes significantly contribute to $f_B$, a result that follows from the smoothness of $|\boldsymbol {B}|$: an $s$ times differentiable function in $L^{1}$ on a torus will have Fourier coefficients that generally scale like $\hat {f}(m)\sim 1/|m|^{s}$ (Grafakos Reference Grafakos2008).

Unlike $f_T$, the helicity of the symmetry $\tilde {\alpha }$ is required for the evaluation of $f_B$. Such knowledge is needed to avoid the symmetric modes of $|\boldsymbol {B}|$ in the summation of (2.7). In a sense, therefore, $f_B$ measures deviations from a particular form of QS. The possibility of enforcing a particular helicity can be of practical importance when optimising for QS configurations.

This measure is also different from $f_T$ in two other important respects. First, $f_B$ is a single flux-surface scalar instead of a local measure. Second, as the Fourier mode resolution of $|\boldsymbol {B}|$ needed for (2.7) is in Boozer coordinates, an evaluation of $f_B$ requires explicit knowledge of these coordinates. This prevents application of $f_B$ to scenarios requiring $\boldsymbol {j}\boldsymbol {\cdot }\boldsymbol {\nabla }\psi \neq 0$, unlike $f_T$. The need to compute Boozer coordinates and Fourier-resolve $B$ also imposes an additional numerical burden (e.g. using existing codes such as BOOZXFORM) (Sanchez et al. Reference Sanchez, Hirshman, Ware, Berry and Spong2000; Paul et al. Reference Paul, Antonsen, Landreman and Cooper2020). While computing the Boozer coordinate transformation is typically not a major numerical bottleneck, avoiding the transformation may be convenient for some applications.

While $f_T$ does not require any particular coordinate system (in this case Boozer coordinates), it also does not allow specification of the desired helicity, $\tilde {\alpha }$. It is thus of interest to explore alternative measures of QS that incorporate both aspects. We now obtain what we call the two-term formulation of QS (Helander Reference Helander2014; Paul et al. Reference Paul, Antonsen, Landreman and Cooper2020).

We begin from (2.1)–(2.3) again. Defining the inner product $C=\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {u}/\varPhi '$ and using (2.1) and (2.2), we can write $C\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B= \boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } \psi \times \boldsymbol {\nabla } B$. This expression holds even when $\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B=0$. To involve $C$ directly in the remaining (2.3), we rewrite it in the form (Rodríguez et al. Reference Rodríguez, Sengupta and Bhattacharjee2021)

(2.8)\begin{equation} \boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla}(\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{u})=\boldsymbol{j}\boldsymbol{\cdot}\boldsymbol{\nabla}\varPhi. \end{equation}

Under the assumption of $\boldsymbol {j}\boldsymbol {\cdot }\boldsymbol {\nabla }\psi =0$, it follows that $C=C(\psi )$ must be a flux function. We can then write

(2.9)\begin{equation} f_C=\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} \psi\times \boldsymbol{\nabla} B-C(\psi) \boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} B, \end{equation}

which must vanish for some flux function $C$ for a QS field. This is the two-term formulation of QS. To avoid having to perform a search for $C$ every time (2.9) is evaluated, we need some form for the flux function $C$ in terms of more recognisable quantities. To do so, we consider the limit of a QS configuration and adopt Boozer coordinates. In that case, and for $\partial _\chi B\neq 0$ (otherwise $f_C=0$ for any value of $C$),

(2.10)\begin{equation} C=\frac{\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} \psi\times\boldsymbol{\nabla} B}{\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} B}=\frac{G+\tilde{\alpha} I}{\iota-\tilde{\alpha}}. \end{equation}

(This flux function $C$ is often given the symbol $F$ in the literature; Helander Reference Helander2014; Paul et al. Reference Paul, Antonsen, Landreman and Cooper2020). Now $C$ is expressed in terms of physical quantities: the Boozer currents $I$ and $G$ and the rotational transform $\iota$. The currents may be calculated without resorting to Boozer coordinates directly (see equation (16) in Helander (Reference Helander2014)). Any flux coordinate system will do, under the assumption of well-defined, nested flux surfaces.

It is clear that the helicity of the symmetry $\tilde {\alpha }$ needs to be specified, just as one must do to evaluate $f_B$. The function $C$ for which it is needed has important physical meaning (Rodríguez et al. Reference Rodríguez, Helander and Bhattacharjee2020), as it is a direct measure of the banana width of trapped particles, $\varLambda$, in the QS limit. In more detail, $\varLambda |\boldsymbol {\nabla }\psi |\sim \rho _\parallel C$, where $\rho _\parallel$ is the Larmor radius associated with the parallel velocity. The resonance at the surface $\iota =\tilde {\alpha }$ leads to $C\rightarrow \infty$, unless $\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla }\psi \times \boldsymbol {\nabla } B=0$ on the surface. This latter condition forces the surface to be isodynamic, which also prevents the appearance of a current singularity and the potential opening of a magnetic island (Rodríguez & Bhattacharjee Reference Rodríguez and Bhattacharjee2021a). In the present work, we avoid such surfaces altogether.

In summary, we have a measure $f_C$ that is local like $f_T$, for which the helicity of the symmetry must be prescribed as in $f_B$, but for which no Boozer coordinates are needed. This makes this formulation amenable to gradient-based optimisation techniques such as adjoint methods (Paul et al. Reference Paul, Antonsen, Landreman and Cooper2020).

This constitutes the derivation of the three measures of QS—$f_B$, $f_T$ and $f_C$—that will occupy us in this paper. In table 1 we summarise the main formal and construction differences of the three forms.

Table 1. Basic properties of QS measures. Summary of the basic formal features of the QS measures introduced in § 2. These include: whether the helicity needs to be specified as an input (I) or not (O), whether Boozer coordinates are required to evaluate it, whether $\boldsymbol {j}\boldsymbol {\cdot }\boldsymbol {\nabla }\psi =0$ is a necessary assumption and whether the measure is local (L) or global (G).

3 Comparison of cost functions

All three forms introduced in the previous section can be interpreted as measures of QS: they all share the basic feature that they vanish only for a configuration that is exactly quasisymmetric. However, each form treats deviations from QS differently. The focus of this section is to evaluate how these measures compare with each other formally, and to connect them to some of the physical properties usually associated with QS.

3.1 Formal relation

It is convenient to measure deviations from QS in terms of the asymmetric components of the magnetic field magnitude. Thus, to understand the forms of $f_T$ and $f_C$, it is informative to learn how the different Fourier modes of $|\boldsymbol {B}|$ are weighted in each of these metrics. Then a comparison with $f_B$ may be done on an equal footing.

Let us start with the definition of $f_C$ and write the magnitude of the magnetic field as a Fourier series:

(3.1)\begin{equation} B=\sum_{n,m}B_{nm}{\rm e}^{{\rm i}(m\theta-n\phi)}, \end{equation}

where the angular Boozer coordinates $\theta$ and $\phi$ are used for the Fourier resolution of the field. Using the construction of $C$ in (2.10) we may write, after some algebra,

(3.2)\begin{align} f_C& =\frac{B^{2}}{\tilde{\alpha}-\iota}(\partial_\phi+\tilde{\alpha}\partial_\theta)B\nonumber\\ & ={\rm i}\frac{B^{2}}{\iota-\tilde{\alpha}}\sum_{n/m\neq\tilde{\alpha}}(n-m\tilde{\alpha})B_{nm} {\rm e}^{{\rm i}(m\theta-n\phi)}. \end{align}

The first line shows that $f_C/B^{2}$, although a local measure, has some global properties like $f_T$ does. In particular, the integral along closed streamlines of $\boldsymbol {Q}=\mathcal {J}\boldsymbol {\nabla }\psi \times \boldsymbol {\nabla }\chi$ must vanish, that is, $\oint (f_C/B^{2})\,\mathrm {d}l/|\boldsymbol {Q}|=0$. The function $f_C/B^{2}$ is also clearly linear in the Fourier content of $B$, but interestingly, not all modes are weighted equally. The modes ‘furthest’ from the symmetry direction are weighted more heavily than those close to it.

Although we have successfully written $f_C$ in terms of the Fourier content of $B$, its form is still far from that of $f_B$. To make a closer comparison with $f_B$ we need to construct a single flux-surface scalar from $f_C$. We take (3.2) and write it as

(3.3)\begin{equation} B_{nm}={-}{\rm i}\frac{\iota-\tilde{\alpha}}{n-m\tilde{\alpha}}\left(\frac{f_C}{B^{2}}\right)_{nm}, \end{equation}

where $(\cdots )_{nm}$ represents the $(n,m)$ Fourier component of the function in parentheses in Boozer coordinates. Using Parseval's theorem, we write

(3.4)\begin{equation} \sum_{n,m}(n-m\tilde{\alpha})^{2}|B_{nm}|^{2}=\frac{(\iota-\tilde{\alpha})^{2}}{(2{\rm \pi})^{2}}\int\,\mathrm{d}\theta\,\mathrm{d}\phi\left(\frac{f_C}{B^{2}}\right)^{2}, \end{equation}

which can be rewritten as

(3.5)\begin{equation} \overline{\left(\frac{f_C}{B^{2}}\right)^{2}} =\sum_{n,m}\left(\frac{n-m\tilde{\alpha}}{\iota-\tilde{\alpha}}\right)^{2}|B_{nm}|^{2}, \end{equation}

where $\overline {(\cdots )}$ denotes the $(0,0)$ Fourier component of the expression in parentheses. Equation (3.5) is now in a form close to $f_B$. The scalarised version of $f_C$, in this case normalised to $B^{2}$ and averaged over the surface, has a form similar to $f_B$ in that it involves a sum over the non-symmetric Fourier modes of $B$. The main difference comes through the different weighting of the modes. The measure $f_C$ can be thought of as a modification of $f_B$ in which asymmetric components of the magnetic field magnitude furthest from QS are more heavily penalised.

The difference in weights implies that $f_B$ and $f_C$ are actually not monotonic with respect to one another. Changing the energy content of the Fourier modes may lead to $f_B$ increasing but $f_C$ decreasing (and vice versa). Having said that, there also is a class of changes for which both metrics respond in the same way. Given that all terms in the summation of (2.7) and (3.5) are positive, reducing any individual mode $|B_{nm}|$ will, for instance, lead to the reduction of both metrics. This observation has important implications for the universality of QS measures and optimisation, to be explored in the next sections.

Let us now see how $f_T$ fares in comparison. When expressed in terms of the Fourier content of the magnetic field strength, the nonlinear character of the triple product formulation is clear. We explicitly write

(3.6)\begin{equation} (\mathcal{J}^{2}f_T)_{nm}={\rm i}\sum_{k,l}(\iota l-k)(km-nl)B_{n-k,m-l}B_{k,l}, \end{equation}

where $\mathcal {J}$ is the Jacobian of Boozer coordinates and $k,l$ are integers. The nonlinearity appears in the form of a convolution that mixes modes in a non-trivial way. We also have one higher derivative of the field strength compared to $f_B$ or $f_C$. Scalarising this expression by computing an average like in the case of $f_C$ does not help in bringing this form any closer to $f_B$. The convolution couples modes together, making it unclear whether reducing a single asymmetric mode will reduce $f_T$ as it did for $f_C$ and $f_B$.

We may try to gain some additional understanding of how $f_T$ is affected by changes in the mode content by considering a system that lies close to QS. Consider a small deviation $\tilde {B}$ in the field magnitude from a dominantly QS field. Then a linearisation of $f_T$ takes the form

(3.7)\begin{equation} (\mathcal{J}^{2}f_T)_{nm}={\rm i}(\tilde{\alpha}m-n)(\iota m-n)\sum_llB_{\tilde{\alpha}l,l}\tilde{B}_{n-\tilde{\alpha}l,m-l}. \end{equation}

It is clear from this expression that decreasing $\tilde {B}$ decreases the magnitude of $f_T$. However, even upon linearisation, it is not obvious that reducing a particular Fourier harmonic will guarantee a lower value of $f_T$.

The measure $f_T$ is, in this form, distinct from the previous metrics. However, there is actually a natural relation between the triple vector form and $f_C$ through a simple magnetic differential equation (see appendix A). This is reasonable given the common origin of the metrics.

3.2 Physical relations

We have explored three forms for QS in this work. These constructions describe deviations from QS differently. To understand the potential physical meaning of these differences, we now explore physical phenomena often associated with QS configurations and their relation to the cost functions.

3.2.1 Single-particle dynamics

The concept of QS as presented in this work is built on single-particle dynamics. In § 2, beginning with the most fundamental definition, we constructed three measures of QS. Thus, it is natural to expect that all these metrics should have some relation to single-particle dynamics.

Let us start from the definition of the momentum: $\bar {p}=-\varPhi +v_\parallel \boldsymbol {b}\boldsymbol {\cdot }\boldsymbol {u}$, where $v_\parallel$ is the parallel velocity of the particle under consideration (Rodríguez et al. Reference Rodríguez, Helander and Bhattacharjee2020). For a QS configuration where $\boldsymbol {u}$ satisfies (2.1)–(2.3), this momentum is approximately conserved to $O(\rho /L)$, where $\rho$ is the particle Larmor radius and $L$ a characteristic length of the system. When the symmetry is broken, we may ask how this momentum evolves in time. To describe that evolution we need to define the vector field $\boldsymbol {u}$. We can always choose this field in a way that (2.1) and (2.2) are satisfied by construction. To do so we assume the existence of flux surfaces as well as pseudosymmetry (Mikhailov et al. Reference Mikhailov, Shafranov, Subbotin, Isaev, Hrenberg, Zille and Cooper2002; Skovoroda Reference Skovoroda2005). The latter guarantees that the contours of constant $|\boldsymbol {B}|$ are linked to the torus avoiding local extrema on the surfaces. These well-behaved contours are vital to the notion of a continuous symmetry, and necessary to prevent $\boldsymbol {b}\boldsymbol {\cdot }\boldsymbol {u}\rightarrow \infty$. Once $\boldsymbol {u}$ has been constructed, one evaluates the dynamics of $\bar {p}$ using the leading-order form of the equations of motion from the Littlejohn Lagrangian (Littlejohn Reference Littlejohn1983; Rodríguez et al. Reference Rodríguez, Helander and Bhattacharjee2020). We may thus write, to $O(\rho /L)$,

(3.8)\begin{equation} \frac{\mathrm{d}}{\mathrm{d}t}\bar{p} =\frac{v_\parallel^{2}}{B^{2}}\left[\boldsymbol{B} \boldsymbol{\cdot}\boldsymbol{\nabla} (\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{u}) -\boldsymbol{j}\boldsymbol{\cdot}\boldsymbol{\nabla}\psi\right] ={-}\left(\frac{v_\parallel}{\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} B}\right)^{2}f_T. \end{equation}

This shows that $f_T$ constitutes a measure of the conservation of $\bar {p}$. Of course, even if $f_T\neq 0$ the above could result in conservation of $\bar {p}$ in the bounce-averaged sense. This averaged conservation could be seen as the difference between QS and omnigeneity (Landreman & Catto Reference Landreman and Catto2012; Helander Reference Helander2014). It is also important to note from (3.8) that the cost function $f_T$ appears with a factor of $1/(\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B)^{2}$. This factor penalises deviations from QS close to the extrema of $|\boldsymbol {B}|$ along field lines. The special importance of the regions close to the extrema is reminiscent of the requirements for the confinement of barely and deeply trapped particles in omnigeneity. The presence of this resonance makes the combination $f_T/(\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B)^{2}$ numerically ill-behaved as a modified cost function. A practical implementation could perhaps be achieved by regularising the combination, as has been employed in optimisation for pseudosymmetry (Mikhailov et al. Reference Mikhailov, Shafranov, Subbotin, Isaev, Hrenberg, Zille and Cooper2002). Although attractive from the single-particle perspective, and even if the resonance is amended, the measure would still exhibit sharp gradients due to the heavy weight near extrema of $|\boldsymbol {B}|$ along field lines.

The argument related to the dynamics of $\bar {p}$ might appear to some extent artificial, as $\bar {p}$ really only gains special dynamical meaning in the limit of QS. We may alternatively look at a more physically relevant property: the net drift of magnetised particles off flux surfaces. It is convenient to express the off-surface bounce-averaged drift of particles, $\Delta \psi$, as $\partial _\alpha \mathcal {J}_\parallel$, where $\mathcal {J}_\parallel =\oint v_\parallel \,\mathrm {d}l$ is the second adiabatic invariant (calculated taking the integral along the field line and between bouncing points) and $\alpha$ is the usual magnetic field-line label (Helander Reference Helander2014). To evaluate this derivative of $\mathcal {J}_\parallel$, consider two nearby lying magnetic field lines on the same magnetic surface at $\alpha$ and $\alpha +\delta \alpha$ (see figure 1). Take the turning points that define the integral for $\mathcal {J}_\parallel$ to be at a field magnitude $|\boldsymbol {B}|=B_t$. We assume that these bounce points exist on both of these field lines (as we are looking at the $\mathcal {J}_\parallel$ associated with a class of particles defined by its bounce points). Then taking the line integral along the contour in figure 1, we compute

(3.9)\begin{equation} \left(\int_{C_\alpha}+\int_{C_{\delta\alpha}}+\int_{C_{B+}}+\int_{C_{B-}}\right)v_\parallel\boldsymbol{b}\boldsymbol{\cdot}\mathrm{d}\boldsymbol{l}=\delta\alpha\int \boldsymbol{\nabla}\psi\boldsymbol{\cdot}\boldsymbol{\nabla}\times(v_\parallel \boldsymbol{b})\frac{\mathrm{d}l}{B}. \end{equation}

To obtain the equality, we apply the Stokes theorem and take the limit of small $\delta \alpha$, taking the latter integral with respect to length along a field line, $l$. Assuming that $\boldsymbol {j}\boldsymbol {\cdot }\boldsymbol {\nabla }\psi =0$ and noting that the integrals along $C_{B+}$ and $C_{B-}$ vanish as $v_\parallel =0$ wherever $|\boldsymbol {B}|=B_t$, we may rewrite

(3.10)\begin{equation} \partial_\alpha \mathcal{J}_\parallel{=}{-}\int\frac{\boldsymbol{\nabla}\psi\times\boldsymbol{\nabla} B\boldsymbol{\cdot}\boldsymbol{B}}{\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} B}\frac{1}{B^{2}}\left(\frac{E}{v_\parallel}+\frac{v_\parallel}{2}\right)\,\mathrm{d}B, \end{equation}

where $B$ is being used to parameterise the integral along the field line. To be precise, we should split the integral along the field line into concatenated pieces joint at $|\boldsymbol {B}|$ extrema. The vanishing of $\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B$ at the stitching points (and the consequent divergence of the integrand) is generally not problematic as one can consider switching back to integrating with respect to $l$ arbitrarily close to them, with that small piece of the integral constituting a vanishingly small contribution. We drop this distinction for simplicity.

Figure 1. Schematic of the loop integral for $\partial _\alpha \mathcal {J}_\parallel$. Schematic of the path integral followed to compute the field line dependence of the second adiabatic invariant $\mathcal {J}_\parallel$. The contours match field lines and contours of constant $|\boldsymbol {B}|=B_t$.

One may write the kinematic expression $(E/v_\parallel +v_\parallel /2)/B^{2}=\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla }(v_\parallel /B)/(\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B)$. As a result, any flux function multiplying this kinematic expression will lead to a vanishing integral. Thus, the first fraction in the integrand can be written as $f_C/(\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B)$. This shows that $f_C$ has a direct relation to the field-line dependence of the second adiabatic invariant. The complicated nature of the kinematic factor, however, obscures this relation. Proceeding with integration by parts and using (A2):

(3.11)\begin{equation} \partial_\alpha \mathcal{J}_\parallel{=}\left.\frac{\boldsymbol{\nabla}\psi\times\boldsymbol{\nabla} B\boldsymbol{\cdot}\boldsymbol{B}}{\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} B}\frac{v_\parallel}{B}\right|_{v_\parallel{=}0}+\int\frac{f_T}{(\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} B)^{2}}v_\parallel\,\mathrm{d}l. \end{equation}

Thus, we have constructed a simple relation between the bounce-averaged, off-surface drift $\Delta \psi$ and the QS measure $f_T$. The measure comes once again normalised by a factor $(\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B)^{2}$. However, this $f_T$ term is not the only term in (3.11). For the expression in (3.11) to be finite (or vanish as we want for QS), the boundary term in (3.11) should not diverge, which requires $\boldsymbol {\nabla }\psi \times \boldsymbol {\nabla } B\boldsymbol {\cdot }\boldsymbol {B}=0$ wherever $\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B=0$. This condition is precisely the condition of pseudosymmetry introduced earlier (Mikhailov et al. Reference Mikhailov, Shafranov, Subbotin, Isaev, Hrenberg, Zille and Cooper2002; Skovoroda Reference Skovoroda2005). Note that this pseudosymmetry condition appears separate from the $f_T$ integral in (3.11). Thus, generally, even when $f_T\neq 0$ as in ominigeneous configurations, the condition should be satisfied if $\partial _\alpha \mathcal {J}_\parallel$ vanishes (and deeply trapped and barely trapped particles are confined). As discussed in § 2, requiring $f_T=0$ everywhere includes this condition. Guaranteeing that the boundary term vanishes also ensures the existence of bouncing points $B_t$ on nearby field lines, in agreement with the assumptions made in evaluating $\partial _\alpha \mathcal {J}_\parallel$.

The discussion above shows that both $f_T$ and $f_C$ can be regarded as natural measures of QS in the single-particle perspective. However, when the configuration fails to be pseudosymmetric, the interpretation of $f_T$ is obscured.

The measure $f_C$ offers some additional particle dynamics insight. As we mentioned when constructing $f_C$ in § 2, the function $C$ in $f_C$ is a measure of the banana width of bouncing particles in the QS limit. The expression in (3.2) inherits the scaling $\propto 1/\bar {\iota }$, where $\bar {\iota }=\iota -\tilde {\alpha }$ from $C$. Hence, $f_C$ also has scaling related to the orbit width and the bootstrap current, which is proportional to the banana width (Helander Reference Helander2014).

Although the metrics bear some important physical content in their relation to the behaviour of single-particle dynamics, the mapping is not one-to-one. More sophisticated measures are required for an accurate description of the dynamics of single particles. Especially important are alpha particles, for which complicated metrics (see Bader et al. Reference Bader, Drevlak, Anderson, Faber, Hegna, Likin, Schmitt and Talmadge2019) or expensive simulations are needed.

3.2.2 Neoclassical transport

Let us now consider the important physics of neoclassical transport (also related to enhanced neoclassical viscosity). One of the distinguishing features of QS is to grant a low level of neoclassical transport to the magnetic configuration (Boozer Reference Boozer1983). In particular, it prevents the detrimental neoclassical $1/\nu$ transport (we do not consider turbulent transport here, which, however, has been seen to be a dominant form of transport in many optimised stellarators once neoclassical transport has been sufficiently reduced) that tends to spoil confinement in stellarators. Of course, as we deviate from QS, this property will degrade. We would like to understand how this degradation is captured (if at all) by the different metrics of QS presented in the paper.

To study this, we introduce first the classical stellarator concept of helical ripple. We define helical ripple as $\epsilon _h\sim B_{nm}/B$, a measure of the symmetry-breaking magnetic field magnitude. This quantity is clearly linked to $f_B$ in (2.7). When a single, small asymmetric mode on top of a QS field is considered, in the $1/\nu$ and $\sqrt {\nu }$ regimes, this ripple leads to an enhancement of neoclassical losses that scales roughly as $D\sim \epsilon _h^{3/2}$ (Ho & Kulsrud Reference Ho and Kulsrud1987; Nemov et al. Reference Nemov, Kasilov, Kernbichler and Heyn1999; Helander & Simakov Reference Helander and Simakov2008). Therefore, there is a direct relationship between $\epsilon _h$ and transport. In this single-mode classical stellarator scenario $f_B$ is the ripple (unlike in the more general case). It is then clear that $f_B$ does incorporate some direct information about transport.

This classical view on transport has the drawback of treating all ‘ripples’ on the same footing. However, a more detailed approach shows that different modes contribute distinctly (Calvo et al. Reference Calvo, Parra, Velasco and Alonso2014). Transport in the $1/\nu$ regime is driven by bouncing particles near minima of $|\boldsymbol {B}|$ along field lines. The enhancement of transport can then come from the introduction of significant additional wells in which new bouncing particles will live. However, this is not the only mechanism that can enhance transport. As particles precess following constant $\mathcal {J}_\parallel$ surfaces, a significant difference in the $|\boldsymbol {B}|$ profile between field lines on the same flux surface will lead to an enhancement in the drifting of bouncing particles off the surface. (Additional effects such as collisional detrapping may also occur.) These phenomena are quantified in Calvo et al. (Reference Calvo, Parra, Velasco and Alonso2014). Small deviations from QS with helical ripple $\epsilon _h$ that do not affect the $|\boldsymbol {B}|$ profile significantly lead to a transport enhancement $\langle \boldsymbol {\varGamma }\boldsymbol {\cdot }\boldsymbol {\nabla }\psi \rangle \sim \epsilon _h^{2}$. (The same scaling is anticipated for flow damping and generally deviations from intrinsic ambipolarity.) Now, if the deviation is due to a mode with a large gradient along $\boldsymbol {B}$, then the changes lead to an enhancement by a factor $O(\epsilon _h)$. When the deviations lead to significant differences between field lines, then the transport changes significantly, not amenable to a perturbative treatment. (Aspects of this scenario can be understood from an analysis of the second adiabatic invariant, studied in the previous section.)

In this more detailed picture it is clear that different modes of $|\boldsymbol {B}|$ contribute differently to transport. The enhancement is significant when the gradients along the field line are large in the sense $|\boldsymbol {b}_0\boldsymbol {\cdot }\boldsymbol {\nabla } B_1|\sim |\boldsymbol {b}_0\boldsymbol {\cdot }\boldsymbol {\nabla } B_0|$, with $B_1$ the asymmetric field strength. In a more illuminating form, $|\boldsymbol {b}_0\boldsymbol {\cdot }\boldsymbol {\nabla } B_1|/|\boldsymbol {b}_0\boldsymbol {\cdot }\boldsymbol {\nabla } B_0|\sim [(n-\iota m)/(M(\tilde {\alpha }-\iota ))](\epsilon _h /\epsilon _\mathrm {QS})$, where we have taken $\epsilon _{\text {QS}}$ to represent the helical ripple associated with the symmetric part of $B$ with symmetric poloidal mode number $M$. The weight in the numerator indicates the predominance of modes with large gradients along the field line, while the denominator serves as a comparative measure for the the gradient along the field line due to the symmetric field.

This difference between modes is clearly missed by the construction of $f_B$. However, we have shown that the angle-averaged $(f_C/B^{2})^{2}$ provides a weighted version of $f_B$ in which higher-order modes are increasingly penalised $\propto (n-m\tilde {\alpha })^{2}$. Although this unequal weight provides a distinction between modes, the weight is not of the form $(n-\iota m)$ as in the transport criterion. It instead weights higher-order modes according to their similarity with respect to the symmetry direction. Thus, $f_C$ only qualitatively matches the high-order mode weighting of the transport scaling. The weight that appears in the definition of $f_C$ can be interpreted as a consequence of the smoothness of the function. The quasisymmetric modes are excluded from the sum through a weight that can be written as a differential operator: the weight must be zero at $n/m=\tilde {\alpha }$ and smoothly increasing away from it. It thus appears natural that the weight in $f_C$ is centred about the symmetry direction. This requirement is avoided in $f_B$, where the symmetric discrete modes are avoided. To more accurately treat the impact on transport, one could imagine a modified version of $f_B$ in which the mode weights are given by $(n-\iota m)$ and the symmetric components are excluded from the summation. It would, however, still require Boozer coordinates (unlike $f_C$). Interestingly, the weight $(m\iota - n)$ appears in the linearised version of $f_T$ in § 3.1.

So far we have focused on small, isolated deviations from QS. In general, however, transport calculations are significantly more complex. The presence of multiple ripples will lead to intricate changes in the trapping and passing particle behaviour. There exist more sophisticated metrics that attempt to describe these changes in more detail. A commonly used example is the $1/\nu$ transport regime effective ripple, $\epsilon _{\mathrm {eff}}^{3/2}$ (Nemov et al. Reference Nemov, Kasilov, Kernbichler and Heyn1999; Nemov, Kasilov & Kernbichler Reference Nemov, Kasilov and Kernbichler2014). The complexity of this measure is not reflected in either $f_B$ or $f_C$ (see the following sections). Thus, although $f_B$ and $f_C$ do contain relevant information regarding transport, they cannot be taken as accurate measures of it.

3.2.3 Current singularities and islands

Although flux surfaces are assumed throughout our analysis, we may also relate the QS metrics to the potential presence of current singularities and magnetic islands (Tessarotto, Johnson & Zheng Reference Tessarotto, Johnson and Zheng1995; Rodríguez & Bhattacharjee Reference Rodríguez and Bhattacharjee2021a). Fourier modes of $|\boldsymbol {B}|$ that resonate with the finite rotational transform can lead to Pfirsch–Schlüter (PS) current singularities. These finite-$\beta$ currents are associated with the potential opening of magnetic islands at the location of the singularities (Reiman & Boozer Reference Reiman and Boozer1984; Rodríguez & Bhattacharjee Reference Rodríguez and Bhattacharjee2021a). The PS current singularity arises due to the periodicity constraint on the magnetic differential equation,

(3.12)\begin{equation} \boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla}\left(\frac{j_{\|}}{B}+\frac{p'C}{B^{2}}\right)={-}2p' \frac{f_C}{B^{3}}, \end{equation}

for the parallel current $j_{\|}$. The periodicity condition required for solvability is

(3.13)\begin{equation} p' \oint \frac{f_C}{B^{4}} \, \mathrm{d}l = 0, \end{equation}

where $l$ measures length along a field line. We note that $f_C$ appears explicitly. This condition will be satisfied in QS, thus precluding the existence of PS singularities (except possibly at the conflictive $\iota =\tilde {\alpha }$ surface; see § 2). Note that solvability may still be obtained away from QS if $p' = 0$ or the line integral vanishes.

One may generally show that in magnetohydrostatics, the resonant radial magnetic field due to the singularities is $B_{\rho nm}\propto (1/B^{2})_{nm}$ (in simplified geometry), and thus $W_i\propto \sqrt {(1/B^{2})_{nm}}$ (Reiman & Boozer Reference Reiman and Boozer1984; Bhattacharjee et al. Reference Bhattacharjee, Hayashi, Hegna, Nakajima and Sato1995; Rodríguez & Bhattacharjee Reference Rodríguez and Bhattacharjee2021a). A non-QS field will potentially have contributions at many rational surfaces, as driven by the asymmetric modes. In that sense, any sum over the asymmetric modes of $B$, such as $f_B$ or $f_C$, will include some of this information. Resonances only occur when the rotational transform matching the asymmetric mode helicity exists within the plasma, and thus a more faithful metric would only account for the resonant asymmetric modes. Due to the generally asymmetric geometry of configurations, resonant fields arising from PS currents may appear at other rational surfaces through toroidal coupling (Rodríguez & Bhattacharjee Reference Rodríguez and Bhattacharjee2021a). This is of course only a partial account of the relevant island physics, focusing on the potentially deleterious seeding of islands. The appearance of islands will in practice depend on a plethora of additional effects beyond the simple magnetohydrostatic equilibria considerations, including, for instance, the healing effects of flows or additional kinetic effects (Fitzpatrick Reference Fitzpatrick1993; Hegna Reference Hegna2011, Reference Hegna2012).

4 Quasisymmetry in practice: universality and cost functions

The discussion so far has explored the origin of the three different measures of QS—$f_B$, $f_C$ and $f_T$—and discussed their formal properties and related physical content. In this section we explore some more practical aspects of these measures. First, we address the question ‘Which configuration is more quasisymmetric?’ and discuss the universality of QS. At a second stage, we explore the implications of these different formulations for the optimisation of stellarators through numerical examples.

4.1 Universality of QS

The goal of the metrics $f_B$, $f_C$ and $f_T$ is to quantify the deviation from QS. In this sense, all these forms would appear to be equally valid candidates. Given the variety of choice, we would like to understand to what extent it is meaningful to state that a certain field configuration is more quasisymmetric than another as measured by these metrics.

As demonstrated in the comparison of the previous section, QS metrics are not generally monotonic with respect to each other. For example, decreasing $f_B$ does not imply decreasing $f_C$. One could, for instance, exchange the energies of a larger low-order mode to a high-order one without changing $f_B$ but modifying $f_C$. The nonlinearity of $f_T$ makes the lack of monotonicity somewhat more marked. This lack of monotonicity makes it difficult to determine magnetic configurations which are ‘closer’ to QS, or for that matter, if the question is meaningful. As different measures of QS yield different answers, any comparison must be defined with respect to a particular standard. The notion thus lacks universality: the quantification of closeness to QS is in some sense arbitrary, and does not necessarily have a clear physical meaning attached to it.

Let us consider some practical examples. In order to use the metrics of QS quantitatively, we first need to normalise and scalarise them. Only then will we be able to make fair comparisons across different devices. Perhaps unsurprisingly, we will see that this procedure is not unique, emphasising the lack of universality.

Let us start with the metric $f_B$, which measures the Fourier content of $B$ in Boozer coordinates. It is customary to normalise it to the total magnetic field energy. So we define

(4.1)\begin{equation} \hat{f}_B=\sqrt{\frac{1}{\overline{B^{2}}}f_B}. \end{equation}

Doing so ensures $\hat {f}_B\in [0,1)$. This finite range of values makes $f_B$ convenient, although the requirement $B>0$ excludes the value of 1. Although this normalisation seems the most natural, it is not unique. For instance, we could decide to normalise $f_B$ to (only) the symmetric part of the field excluding the flux-averaged part, or we could normalise by the flux-surface average of $B^{2}$, $\langle B^{2}\rangle$. For the purpose of this discussion, we take the natural and more conventional form (4.1).

The choice of a ‘natural’ normalisation is unclear for $f_C$ and $f_T$. Fundamentally, our normalisation should give a dimensionless measure. This principle is, of course, not specific enough. To constrain the construction of the measure further, we try making their Fourier forms look closest to $f_B$. This is a convenient choice. Let us start with $f_C$, which dimensionally scales as $f_C\sim B^{3}$. In addition to this dimensional factor, motivated by the form of (3.2), we also choose to normalise by $1/(\tilde {\alpha }-\iota )$ to bring the measure closer to the form of $f_B$. One could argue for adding additional dimensionless factors such as the inverse aspect ratio, $\epsilon$, which characterises the magnitude of $C\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B$ for $\chi = \theta$, but for simplicity we do not do so here. In view of the normalisation of $f_B$, we normalise and scalarise $f_C$ as follows:

(4.2)\begin{equation} \hat{f}_C=\sqrt{\frac{\langle f_C^{2}\rangle}{\langle B^{2}\rangle^{3}}|\tilde{\alpha}-\iota|^{2}}. \end{equation}

This leads to a form that is close to but not identical to (3.2).

The normalisation of $f_T$ is inherently more complex. The triple vector product introduces explicit length scales which must be non-dimensionalised. Once again, we look at the form of $f_T$ in (3.5). It is then natural to normalise and scalarise $f_T$ using

(4.3)\begin{equation} \hat{f}_T=\sqrt{\langle f_T^{2}\rangle\frac{R^{4}}{\langle{B^{2}}\rangle^{4}}}. \end{equation}

In obtaining this normalisation we have used the scaling $\mathcal {J}\sim R/B$, where $R$ is the major radius. Once again, one could argue that the choice is, to some extent, arbitrary. It is interesting to note that the combination $f_T/(\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B)^{2}$, recurrent in the description of single-particle dynamics in the previous section, is by construction dimensionless. Thus, it appears to be a natural choice. However, we do not consider it here due to the added complication of the resonant denominators and the associated steep gradients.

Now that we have chosen specific normalisations for the QS metrics, we may compare their values in stellarator equilibria. We consider two scenarios. First, we focus on the comparison of the metrics in configurations that have been designed to be approximately quasisymmetric. Then we study how these measures differ when the system has not necessarily been optimised to be quasisymmetric.

We present in figure 2 a comparison of the metrics in six quasisymmetric configurations:Footnote 1 GAR (Garabedian Reference Garabedian2008; Garabedian & McFadden Reference Garabedian and McFadden2009), HSX (Anderson et al. Reference Anderson, Almagri, Anderson, Matthews, Talmadge and Shohet1995), NCSX (Zarnstorff et al. Reference Zarnstorff, Berry, Brooks, Fredrickson, Fu, Hirshman, Hudson, Ku, Lazarus and Mikkelsen2001), ESTELL (Drevlak et al. Reference Drevlak, Brochard, Helander, Kisslinger, Mikhailov, Nührenberg, Nührenberg and Turkin2013), QHS48 (Ku & Boozer Reference Ku and Boozer2010) and ARIESCS (Najmabadi et al. Reference Najmabadi, Raffray, Abdel-Khalik, Bromberg, Crosatti, El-Guebaly, Garabedian, Grossman, Henderson and Ibrahim2008). Within a given device, the overall correlation between cost functions is good. All measures qualitatively show a similar behaviour: QS improves as we move from the boundary inwards. However, in regions closest to the axis, this correspondence is not as strong (especially in configurations such as GAR and QHS48). However, the same cost function correspondence does not seem to hold if we make comparisons between different configurations.

Figure 2. Comparison of defined cost functions in quasisymmetric designs. Plots in logarithmic scale including a comparison of the cost functions $\hat {f}_B$, $\hat {f}_C$ and $\hat {f}_T$ for a number of quasisymmetric devices. The scatter points for each device indicate values at a number of equally spaced magnetic flux surfaces joined by lines in order from the centre to the edge of the configurations (the rightmost points represent the edge).

To quantify this lack of ordering across devices, we compute the Spearman correlation coefficient, which considers correlation only taking order into account. We show this in figure 3. The high correlation between cost functions within a particular device (which in all cases was ${\sim }0.9$$1$) clearly does not hold across devices. Although $\hat {f}_C$ and $\hat {f}_T$ retain a high degree of correlation (${\sim }0.9$), the conventional $\hat {f}_B$ measure seems largely uncorrelated to the others. (Of course, in this inter-configuration comparison the normalisation factors chosen are particularly important.)

Figure 3. Correlation between edge values of cost functions and $\epsilon _\mathrm {eff}$ of different quasisymmetric configurations. Diagram showing the Spearman correlation between the edge values of the different cost functions and $\epsilon _\mathrm {eff}$ of the configurations in figure 2. The colour represents the coefficient (as does the diameter of the coloured circles).

The loss of correlation across machines emphasises the main point on universality from a practical perspective. Depending on the form of cost function taken, one could say either A or B is more quasisymmetric. The statement has only a relative meaning at best for these optimised configurations. Without a physical basis to the measures it is difficult to argue in favour of one definition over the other. For a more meaningful statement, one should instead compare these functions through some physical property. Ultimately we are interested in finding configurations that are close to QS because we seek some of the physical properties that QS confers upon magnetic configurations. In figure 3 we compare $\epsilon _\mathrm {eff}$ (which also contains some arbitrary normalisation factors in its definition; Nemov et al. Reference Nemov, Kasilov, Kernbichler and Heyn1999) with the three cost functions for the devices of figure 2. Other physics measures such as alpha particle confinement could also be suited for such comparison (Bader et al. Reference Bader, Drevlak, Anderson, Faber, Hegna, Likin, Schmitt and Talmadge2019). There exists some correlation between the transport and the normalised values of $f_C$ and $f_T$. Especially remarkable is the seemingly little correlation between $\hat {f}_B$, which is widely used for optimisation, and $\epsilon _\mathrm {eff}$. Overall, for these ‘quasisymmetric’ configurations, the QS cost functions appear not to be good indicators of their transport levels. Similar behaviour has been observed in previous studies of quasisymmetric configurations (Martin & Landreman Reference Martin and Landreman2020).

In summary, when comparing quasisymmetric configurations, there is no universal measure of QS. The comparison depends on the metric used, and these differ from each another. To make a comparison physically meaningful, especially close to optimised configurations, we are in need of resorting to direct physical measures of the configuration (e.g. neoclassical transport or alpha particle confinement).

4.2 Optimising for QS

The differences in the formal structure of the cost functions in conjunction with the difference in behaviour close to QS suggest that the different measures of QS will behave quite differently as optimisation cost functions.

From a formal perspective, we would expect to find $f_B$ and $f_C$ to perform similarly, while the nonlinear $f_T$ to differ significantly. We now present some examples to illustrate the differences in the optimisation. We consider an example optimisation problem in which QS is targeted throughout the plasma volume. The optimisation parameters correspond to the Fourier modes describing the outermost surface:

(4.4)\begin{equation} \left. \begin{aligned} R & = \sum_{nm} R_{nm} \cos(m \vartheta - n N_p \phi_c), \\ Z & = \sum_{nm} Z_{nm} \sin(m \vartheta - n N_p \phi_c), \end{aligned} \right\} \end{equation}

where $N_p$ is the number of field periods, $\phi _c$ is the cylindrical angle and $\vartheta$ is a general poloidal angle. The configuration optimised is based on a quasihelicallysymmetric configurationFootnote 2 with $N_p=4$ field periods. We investigate two aspects of the problem. First, we consider slices of parameter space with the purpose of observing the differences in QS measures in practice beyond the region close to the minima. This should give a partial idea of how the problem changes as a result of changing the form of the optimisation cost function. Then, we perform various optimisations using STELLOPT (Lazerson et al. Reference Lazerson, Schmitt, Zhu and Breslau2020) with implementations of the three different forms of the cost functions. Let us start with the exploration of parameter space. Figure 4 presents the evaluation of the QS metrics and $\epsilon _\mathrm {eff}$ as a function of two parameters characterising the surface (the $n=2,~m=1$ modes of the boundary (4.4)), and Figure 5 the magnitude of the associated gradient. The metrics are evaluated at six equally spaced magnetic flux surfaces, and their contributions are added together.

Figure 4. Parameter space spanned by $n=2,m=1$ surface mode for $\hat {f}_B$, $\hat {f}_C$, $\hat {f}_T$ and $\epsilon _\mathrm {eff}$. The plots show, clockwise, the cost functions $\hat {f}_B$, $\hat {f}_C$, $\epsilon _\mathrm {eff}$ and $\hat {f}_T$ as a function of the $(2,1)$ surface modes. The parameter scan is performed using the map option of STELLOPT around the stellarator design in the footnote of p. 28. The minima for each of these are found (clockwise) at: $(-0.031,0.031)$, $(-0.031,0.031)$, $(-0.037,0.045)$ and $(-0.037,0.037)$.

Figure 5. Gradients in parameter space spanned by $n=2,m=1$ surface mode for $\hat {f}_B$, $\hat {f}_C$, $\hat {f}_T$ and $\epsilon _\mathrm {eff}$. The plots show, clockwise, the parameter space gradients of $\hat {f}_B$, $\hat {f}_C$, $\epsilon _\mathrm {eff}$ and $\hat {f}_T$ as a function of the $(2,1)$ surface modes. Scales of the gradients are significantly different between measures.

The objective landscape varies depending on the metrics used. Qualitatively, though, all the measures show a similar behaviour. The metrics share a region of reduced cost, which seem to roughly coincide. In a global sense, the metrics seem not to disagree with each other. This might appear as a surprise given the observations of the QS configurations and some of the formal discussion above. We should, however, remember that there exist certain changes that affect all metrics favourably. Away from a QS configuration it thus seems that there are directions in which one may reduce many of these modes such that all metrics decrease. This is supported by the parameter space representations in figure 4. However, note that the similarity in global trends does not preclude differences in local trends.

First, the relative variation of the cost functions in space is significantly different. Gradients seem to be relatively strong for the highly nonlinear measures $\hat {f}_T$ and $\epsilon _\mathrm {eff}$, while the other two metrics behave more smoothly. This distinction in the gradients also seems to leave a region close to the minima much shallower for $\hat {f}_T$ and $\epsilon _\mathrm {eff}$, comparatively, which may lead to difficulties in accessing the exact minima. Second, beyond these differences in magnitude, there are also differences in the shape and location of the minima. The region near a minimum can be more or less elongated depending on the measure, and, importantly, the minimiser is located at different values. In this particular case, minimisers vary by up to ${\sim }20\,\% - 50$ % in the boundary coefficients between different measures (although $\hat {f}_B$ and $\hat {f}_C$ have, in this case, roughly the same minimiser). In brief, although far from a minimum the various metrics may exhibit a qualitatively similar behaviour, close to the minimum there are significant differences. These latter differences are consistent with the discussion on the lack of universality in § 4.1.

Of course, this illustration of the optimisation space is only one slice of a multidimensional manifold. The large number of degrees of freedom means that the parameter space slices will change as other parameters do. In general, this makes optimisation a highly complex process in which the interaction between different dimensions may amplify the differences between metrics. In addition, the algorithm employed to move around in optimisation space will also affect the result of a given optimisation. A discussion on how the optimisation is likely to perform is therefore extremely difficult, and we instead present an illustrative example. With that being said, from our discussion we expect the optimisation of $\hat {f}_B$ and $\hat {f}_C$ to be similar, while the nonlinearity of $\hat {f}_T$ to change the optimisation significantly. We hypothesise here that there is a rough relation between changes in boundary shape and the Fourier content of $B$.

We present the resulting cross-sections and values associated with the various metrics from optimisations in which QS is targeted in a volume using the different forms of QS metrics as cost functions in figure 6 and table 2. We start from a configuration that is deliberately chosen to be some distance away in parameter space from the QS configuration found in the footnote of p. 28. The optimisation uses a BFGS gradient-based method (as part of the suite MANGOFootnote 3) in which there is a single scalarised cost function with 32 degrees of freedom in the form of Fourier components (from 0 through 3, keeping the major radius $R_{00}$ fixed) defining the surface as in (4.4). The process does not penalise any additional property such as the commonly targeted rotational transform or aspect ratio. By studying a simplified problem, we hope to gain more insight into the differences between the various QS metrics. The optimisation is performed for several values of the finite-difference step size to ensure convergence.

Figure 6. Comparison of cross-sections of optimisation results for different cost functions. Plot showing the cross-section of the stellarators as obtained from the optimisation using different forms of the cost function. The optimisation stops when the optimiser is unable to make further improvement (the characteristic change in the parameters during an iteration is $10^{-4}$$10^{-3}$). The broken line represents the starting point, while the black, blue and red lines represent the $f_B$, $f_T$ and $f_C$ optimisation results, respectively. The cross-sections are only presented in one half of the plane but the missing plane can be reconstructed symmetrically.

Table 2. Averaged metric values for optimised stellarators. Summary of the average value of the metric over the volume of the optimised stellarators with respect to the different cost functions.

Comparing cross-sections in figure 6 and the values in table 2 we observe that the resulting optima are quite different from one another. This emphasises the points discussed in this paper: that the precise form of the cost function has important qualitative implications. The optimisation using the standard form of $\hat {f}_B$ seems, in this case, to perform best in allowing the configuration to reach a state of minimal transport and QS measures. The optimisation by $\hat {f}_C$ is not far away, and a comparison of the cross sections shows similar shaping features. Interestingly, optimising for $\hat {f}_C$ does not seem to yield the smallest value for $\hat {f}_C$ (which $\hat {f}_B$ optimisation does). This suggests that the optimisation for $\hat {f}_C$ gets stuck in some effective local minimum.

In light of the comparison made in § 3, and with the hypothesis that the surface variations are roughly related to variations in $B$, we can think of optimisation as a process that rearranges energy in the space of Fourier modes of $|\boldsymbol {B}|$. The relation is generally complex and nonlinear, but thinking in these terms may still shed some light on the differences in optimisation. In this rearrangement of energy in Fourier space, $\hat {f}_B$ contains a large space of degenerate interchanges. Energy can be freely exchanged between asymmetric modes of $B$, regardless of their mode numbers. In that sense, reducing $\hat {f}_B$ means, necessarily, reducing some asymmetric mode. On the other hand, the cost function $\hat {f}_C$ distinguishes between modes. Different weights at different mode numbers lift part of the degeneracy in the exchange of mode energy that occurs for $\hat {f}_B$. This distinction effectively excludes certain regions of the mode space from being used for the restacking of the mode energy. In addition to decreasing the magnitude of an asymmetric mode, moving energy to modes closer to the symmetry direction will decrease $\hat {f}_C$. This could be a mixed blessing. Having this additional mechanism could be a way to avoid certain local minima by adding a new direction of descent. However, it could also present additional local minima if the optimisation prematurely reduces modes very close to the symmetry direction.

Relating changes in the mode content to changes in the boundary, we expect these two metrics to facilitate different ways of modifying the boundary. It is not clear which metric will lead to improved optimisation. We have here presented a single numerical exercise in which $\hat {f}_B$ appeared to perform better, but the performance of these metrics will likely depend on the particulars of the optimisation problem as well as other aspects such as the optimisation algorithm used.

The optimisation of $\hat {f}_T$ in the presented example exhibits the worst performance of all. The cross-sections show large differences with respect to the other configurations, suggesting that optimisation has stopped at a distinct local minimum. That the measure $\hat {f}_T$ is lower in the $\hat {f}_B$-optimised case serves as evidence that the optimisation of $\hat {f}_T$ terminated at a different local minimum. The presence of nonlinearities in $f_T$ is perhaps responsible for this additional minimum. In the light of the Fourier space energy picture, $\hat {f}_T$ not only depends on the magnitude and the location of the mode energy, but also on the relative location of modes. This nonlinear coupling complicates the problem, potentially leading to additional local minima. Furthermore, the nonlinearity of $\hat {f}_T$ results in an objective landscape which appears flatter near the minimum (see figure 4), which may make difficult reaching an accurate optimum. Although this qualitative picture fits with observations made for this optimisation example, we remark that the behaviour is likely to vary between problems. An important additional point of distinction of the $\hat {f}_T$ optimisation with respect to the others is, as emphasised in § 2, that the form of symmetry of the configuration is not specified. Hence, the optimisation must locally decide upon the type of QS towards which to evolve. If close to a configuration with good QS, optimisation will naturally bring the system towards the corresponding helicity. However, generally it is difficult to predict given an initial configuration the form of QS towards which optimisation will tend.

In the example presented, the aspect ratio of the configuration does not seem to change significantly between and along optimisations (values lie within ${\sim }15\,\%$ of the initial point). This seems counterintuitive, as from a near-axis perspective one would expect a larger aspect ratio to allow for a more quasisymmetric solution. This follows from the possibility of satisfying QS conditions sufficiently close to the magnetic axis, as the relevant expansion parameter for near axis scales as $\sqrt {\psi /B}/R$ (Landreman & Sengupta Reference Landreman and Sengupta2019). Since we are fixing the total flux, decreasing the minor radius of the toroidal configuration can increase the differential flux over which the QS error is small. In practice, however, the optimiser has difficulty in finding these large-aspect-ratio minima. The difficulty in changing the aspect ratio without sacrificing QS along the way and the choice of parameterisation (Henneberg, Helander & Drevlak Reference Henneberg, Helander and Drevlak2021) (4.4) might provide an explanation for this behaviour.

What conclusions can we thus draw from this illustrative example? First, the form of the cost function must be carefully considered. Here we only examined QS and observed that the optimisation result varies significantly depending on the form of the cost function. Similar behaviour will likely affect the optimisation of other physical measures. Second, we see in this example that the ‘simplest’ forms (in this case, $f_B$ and $f_C$) seemed to provide a better global performance. (By simpler we mean a lower number of field derivatives and a simpler dependence on the Fourier modes.) Increased complexity seems to make optimisation harder. However, it is difficult to make general statements based only on this example, and there might be exceptions not treated in this paper.

As discussed in the previous section, close to local QS minima, improvements in QS metrics do not always lead to improvements in physical properties. Thus insisting on making marginal improvements on $\hat {f}_B$ (and for that matter, any of the other QS measures) does not seem to be a useful exercise. Close to a QS minimum it does seem appropriate to prioritise optimisation with respect to more physical measures. Optimising stellarators based solely on physical properties has been shown to be successful in the literature (Hirshman et al. Reference Hirshman, Spong, Whitson, Lynch, Batchelor, Carreras and Rome1998, Reference Hirshman, Spong, Whitson, Nelson, Batchelor, Lyon, Sanchez, Brooks, Y.-Fu and Goldston1999; Spong et al. Reference Spong, Hirshman, Whitson, Batchelor, Sanchez, Carreras, Lynch, Lyon, Valanju and Miner2000). Here we propose a hybrid approach between optimising for QS alone and optimising for physical metrics (without QS). The QS optimisation should bring the search close to a region of configurations with QS-like properties following a simpler form of cost function. Then for refinement, it will follow the more detailed physical cost function optimisation. Perhaps insisting on QS at all times may stand its ground if an absolute QS zero could be reached (a point with clear physical meaning), such as in the optimisation for QS on a single magnetic surface. But even then, stellarator design usually requires additional physical considerations.

We seem then to need to alternate between an optimisation dominated by the QS metrics and an optimisation guided by other physics metrics. A way forward is to impose QS as an inequality constraint, so that physical metrics can dominate optimisation when QS is ‘small enough’. Of course there is the unavoidable difficulty of choosing the value of a QS metric to indicate ‘sufficient QS.’ Picking a fixed value will modify the shape of optimisation space in a non-trivial way. Instead of choosing a fixed value, a more dynamic alternative would be to turn QS optimisation on and off guided by the relative changes or gradient norms of the QS metrics. One might choose to not optimise for QS at all, instead initialising the search from an already fairly quasisymmetric configuration. This would place the optimisation in the right basin, but lead to refinements in it based on physically relevant features. Possible initial seeds could be constructed from near-axis solutions.

As explored, the non-universality of the QS metrics close to minima seems to require amendment by introducing additional physics metrics. At the same time, detailed physical calculations are likely to lead to complicated parameter spaces including multiple local minima compared with those of the simpler QS metrics. Using QS metrics in conjunction with a more complex physical metric may provide the best of both worlds. This could shed some light on recent results in Bader et al. (Reference Bader, Drevlak, Anderson, Faber, Hegna, Likin, Schmitt and Talmadge2019). In this work, optimisation of QS through $f_B$ together with a more complex physical proxy for $\alpha$ particle losses seems to perform better than each of the pieces separately. We hypothesise that $f_B$ allows for a smoother global navigation (avoiding some potential local minima), while the physics metric helps in finding a more physically meaningful final minimum. Using both metrics in an appropriately balanced way seems to lead to a convenient compromise in which benefits of both aspects are brought together.

Stellarator design often requires multi-objective optimisation. Given the nonlinearity of the objectives of interest, adding additional terms to the cost function can change the problem dramatically. Thus, the differences observed above for the QS metrics will only be more complex in the presence of other physics objectives. Those differences may be amplified or reduced depending on their relation to the behaviour of the other metrics. Such an analysis will be necessary to reason how to best compose QS with these other metrics. In these complex scenarios, the idea of initially seeding optimisation with an approximately QS configuration gains strength, as it would make the problem simplest. Further study is required to verify these proposed optimisation strategies.

5 Conclusion

In this paper we have explored three main measures of QS: the standard Fourier form, the two-term formulation and the triple product form. We performed an analytical comparison of their origin and mathematical structure. Particularly important are their differences in weighting the asymmetric modes of $B$ in Boozer coordinates. While the standard Fourier form treats all modes equally, the two-term form weights them differently, and the triple vector formulation involves a mode convolution.

The differences in these valid QS metrics show that there is no universal measure of QS. While the measures are all rigorously equivalent when QS is satisfied exactly (assuming $\boldsymbol {j} \boldsymbol {\cdot } \boldsymbol {\nabla } \psi = 0$), there is no unique measure that quantifies deviations from QS. This lack of universality makes the comparison between configurations particularly difficult near approximately QS solutions, and some physically relevant property (such as neoclassical transport through $\epsilon _\mathrm {eff}$ or alpha particle confinement) would be needed for a more meaningful comparison. Near those local minima that are not exactly QS, optimising QS cost functions for marginal improvements generally lacks physical meaning. Away from the minima, the global structures of the QS metrics are qualitatively similar, providing more physical meaning to the metrics.

We also show that the mathematical form of the measure affects the optimisation significantly via an example. In the particular example presented, the complexity of a cost function seems to lead to ending in alternative local minima.

In light of the above, we suggest enforcing approximate QS either through an inequality constraint (so that optimisation close to minima is dominated by some direct physics measure) or through an appropriate initial configuration. Additional attention should also be paid to the form of physics or engineering cost functions.

Acknowledgements

The authors would like to acknowledge Daniel Dudt for fruitful discussion.

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

Funding

This research is primarily supported by a grant from the Simons Foundation/SFARI (560651, A.B.) and DoE contract no. DE-AC02-09CH11466. E.R. was also supported by the Charlotte Elizabeth Procter Fellowship at Princeton University. E.P. was also supported by the Presidential Postdoctoral Research Fellowship at Princeton University.

Declaration of interests

The authors report no conflict of interest.

Data availability

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

Appendix A. Magnetic differential relation between $f_C$ and $f_T$

Although the triple vector product form of QS seems distinct from the other forms given its nonlinear character, we can establish a simple relation between $f_T$ and $f_C$. To obtain this relation it is most convenient to use the same tools that we used in the derivations of the different forms of QS in § 2. Away from QS, one may still construct the symmetry field as $\tilde {\boldsymbol {u}}=\boldsymbol {\nabla }\varPhi \times \boldsymbol {\nabla } B/(\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B)$, where we assume $\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B\neq 0$. At extremal points of the magnetic field magnitude along $\boldsymbol {B}$, this construction fails. We have seen the role of the extrema, for example, in the discussion of single-particle dynamics (§ 3). Requiring pseudosymmetry takes care of these, as it guarantees $\boldsymbol {B}\boldsymbol {\cdot }\tilde {\boldsymbol {u}}$ to be well-behaved. Constructing this vector field $\tilde {\boldsymbol {u}}$ in this way is equivalent to enforcing (2.1) and (2.2). With these as well as $\boldsymbol {\nabla }\boldsymbol {\cdot }\boldsymbol {B}=0$:

(A 1)\begin{equation} \boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla}(\boldsymbol{B}\boldsymbol{\cdot}\tilde{\boldsymbol{u}})=B^{2}\boldsymbol{\nabla}\boldsymbol{\cdot}\tilde{\boldsymbol{u}}, \end{equation}

using $\boldsymbol {j}\boldsymbol {\cdot }\boldsymbol {\nabla }\psi =0$ and appropriate vector identities to rewrite the left-hand side. Now, from $\boldsymbol {B}\boldsymbol {\cdot }\tilde {\boldsymbol {u}}/\varPhi '=C+f_C/(\boldsymbol {B}\boldsymbol {\cdot }\boldsymbol {\nabla } B)$, it then follows that $f_C$ and $f_T$ are related to each other through a magnetic differential equation:

(A 2)\begin{equation} \boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} \left(\frac{f_C}{\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} B}\right)={-}\frac{f_T}{(\boldsymbol{B}\boldsymbol{\cdot}\boldsymbol{\nabla} B)^{2}}. \end{equation}

This magnetic differential equation shows that $f_T$ does, in fact, contain a higher-order derivative in comparison with $f_C$.

Footnotes

2 The basis of the optimisation is the 34DOF_varyBoundary_targetQuasisymmetry scenario in https://github.com/landreman/stellopt_scenarios.

References

REFERENCES

Akçelik, V., Biros, G., Ghattas, O., Hill, J., Keyes, D. & van Bloemen Waanders, B. 2006 Parallel algorithms for PDE-constrained optimization. In Parallel Processing for Scientific Computing (eds. Heroux, M.A., Raghavan, P. & Simon, H.D.), pp. 291322. SIAM.CrossRefGoogle Scholar
Anderson, F.S.B., Almagri, A.F., Anderson, D.T., Matthews, P.G., Talmadge, J.N. & Shohet, J.L. 1995 The Helically Symmetric Experiment (HSX) goals, design and status. Fusion Technol. 27 (3 T), 273277.CrossRefGoogle Scholar
Bader, A., Drevlak, M., Anderson, D.T., Faber, B.J., Hegna, C.C., Likin, K.M., Schmitt, J.C. & Talmadge, J.N. 2019 Stellarator equilibria with reactor relevant energetic particle losses. J. Plasma Phys. 85 (5), 905850508.CrossRefGoogle Scholar
Bernardin, M.P., Moses, R.W. & Tataronis, J.A. 1986 Isodynamical (omnigenous) equilibrium in symmetrically confined plasma configurations. Phys. Fluids 29 (8), 26052611.CrossRefGoogle Scholar
Bhattacharjee, A., Hayashi, T., Hegna, C.C., Nakajima, N. & Sato, T. 1995 Theory of pressure-induced islands and self-healing in three-dimensional toroidal magnetohydrodynamic equilibria. Phys. Plasmas 2 (3), 883888.CrossRefGoogle Scholar
Boozer, A.H. 1981 Plasma equilibrium with rational magnetic surfaces. Phys. Fluids 24 (11), 19992003.CrossRefGoogle Scholar
Boozer, A.H. 1983 Transport and isomorphic equilibria. Phys. Fluids 26 (2), 496499.CrossRefGoogle Scholar
Burby, J.W., Kallinikos, N. & MacKay, R.S. 2020 Some mathematics for quasi-symmetry. J. Math. Phys. 61 (9), 093503.CrossRefGoogle Scholar
Calvo, I., Parra, F.I., Velasco, J.L. & Alonso, J.A. 2014 Flow damping in stellarators close to quasisymmetry. Plasma Phys. Control. Fusion 57 (1), 014014.CrossRefGoogle Scholar
Constantin, P., Drivas, T.D. & Ginsberg, D. 2021 On quasisymmetric plasma equilibria sustained by small force. J. Plasma Phys. 87 (1), 905870111.CrossRefGoogle Scholar
Dekeyser, W., Reiter, D. & Baelmans, M. 2014 Automated divertor target design by adjoint shape sensitivity analysis and a one-shot method. J. Comput. Phys. 278, 117132.CrossRefGoogle Scholar
Drevlak, M., Brochard, F., Helander, P., Kisslinger, J., Mikhailov, M., Nührenberg, C., Nührenberg, J. & Turkin, Y. 2013 ESTELL: a quasi-toroidally symmetric stellarator. Contrib. Plasma Phys. 53 (6), 459468.CrossRefGoogle Scholar
Fitzpatrick, R. 1993 Interaction of tearing modes with external structures in cylindrical geometry (plasma). Nucl. Fusion 33 (7), 10491084.CrossRefGoogle Scholar
Garabedian, P.R. 2008 Three-dimensional analysis of tokamaks and stellarators. Proc. Natl Acad. Sci. 105 (37), 1371613719.CrossRefGoogle ScholarPubMed
Garabedian, P.R. & McFadden, G.B. 2009 Design of the DEMO fusion reactor following ITER. J. Res. Natl Inst. Stand. Technol. 114 (4), 229.CrossRefGoogle ScholarPubMed
Garren, D.A. & Boozer, A.H. 1991 Existence of quasihelically symmetric stellarators. Phys. Fluids B 3 (10), 28222834.CrossRefGoogle Scholar
Grafakos, L. 2008 Fourier Analysis on the Torus, pp. 188. Springer.Google Scholar
Hegna, C.C. 2011 Kinetic shielding of magnetic islands in 3D equilibria. Plasma Phys. Control. Fusion 53 (2), 024003.CrossRefGoogle Scholar
Hegna, C.C. 2012 Plasma flow healing of magnetic islands in stellarators. Phys. Plasmas 19 (5), 056101.CrossRefGoogle Scholar
Helander, P. 2014 Theory of plasma confinement in non-axisymmetric magnetic fields. Rep. Prog. Phys. 77 (8), 087001.CrossRefGoogle ScholarPubMed
Helander, P. & Simakov, A.N. 2008 Intrinsic ambipolarity and rotation in stellarators. Phys. Rev. Lett. 101, 145003.CrossRefGoogle ScholarPubMed
Henneberg, S., Drevlak, M., Nuhrenberg, C., Beidler, C., Turkin, Y., Loizu, J. & Helander, P. 2019 Properties of a new quasi-axisymmetric configuration. Nucl. Fusion 59 (2), 026014.CrossRefGoogle Scholar
Henneberg, S., Helander, P. & Drevlak, M. 2021 Representing the boundary of stellarator plasmas. J. Plasma Phys. 87 (5), 905870503.CrossRefGoogle Scholar
Hirshman, S.P., Spong, D.A., Whitson, J.C., Lynch, V.E., Batchelor, D.B., Carreras, B.A. & Rome, J.A. 1998 Transport optimization and MHD stability of a small aspect ratio toroidal hybrid stellarator. Phys. Rev. Lett. 80, 528531.CrossRefGoogle Scholar
Hirshman, S.P., Spong, D.A., Whitson, J.C., Nelson, B., Batchelor, D.B., Lyon, J.F., Sanchez, R., Brooks, A., Y.-Fu, G., Goldston, R.J., et al. 1999 Physics of compact stellarators. Phys. Plasmas 6 (5), 18581864.CrossRefGoogle Scholar
Ho, D.D. & Kulsrud, R.M. 1987 Neoclassical transport in stellarators. Phys. Fluids 30 (2), 442461.CrossRefGoogle Scholar
Ku, L. & Boozer, A. 2010 New classes of quasi-helically symmetric stellarators. Nucl. Fusion 51 (1), 013004.CrossRefGoogle Scholar
Landreman, M. & Catto, P.J. 2012 Omnigenity as generalized quasisymmetry. Phys. Plasmas 19 (5), 056103.CrossRefGoogle Scholar
Landreman, M. & Sengupta, W. 2018 Direct construction of optimized stellarator shapes. Part 1. Theory in cylindrical coordinates. J. Plasma Phys. 84 (6), 905840616.CrossRefGoogle Scholar
Landreman, M. & Sengupta, W. 2019 Constructing stellarators with quasisymmetry to high order. J. Plasma Phys. 85 (6), 815850601.CrossRefGoogle Scholar
Lazerson, S., Schmitt, J., Zhu, C., Breslau, J. & STELLOPT Developers and USDOE Office of Science 2020 STELLOPT, Version 2.7.5. https://www.osti.gov/biblio/1617636.Google Scholar
Littlejohn, R.G. 1983 Variational principles of guiding centre motion. J. Plasma Phys. 29 (1), 111125.CrossRefGoogle Scholar
Martin, M.F. & Landreman, M. 2020 Impurity temperature screening in stellarators close to quasisymmetry. J. Plasma Phys. 86 (3), 905860317.CrossRefGoogle Scholar
Mikhailov, M., Shafranov, V., Subbotin, A., Isaev, M., Hrenberg, J.N., Zille, R. & Cooper, W. 2002 Improved $\alpha$-particle confinement in stellarators with poloidally closed contours of the magnetic field strength. Nucl. Fusion 42 (11), L23L26.CrossRefGoogle Scholar
Mynick, H.E. 2006 Transport optimization in stellarators. Phys. Plasmas 13 (5), 058102.CrossRefGoogle Scholar
Najmabadi, F., Raffray, A., Abdel-Khalik, S., Bromberg, L., Crosatti, L., El-Guebaly, L., Garabedian, P., Grossman, A., Henderson, D., Ibrahim, A., et al. 2008 The ARIES-CS compact stellarator fusion power plant. Fusion Sci. Technol. 54 (3), 655672.CrossRefGoogle Scholar
Nemov, V.V., Kasilov, S.V. & Kernbichler, W. 2014 Collisionless high energy particle losses in optimized stellarators calculated in real-space coordinates. Phys. Plasmas 21 (6), 062501.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 (12), 46224632.CrossRefGoogle Scholar
Nührenberg, J. & Zille, R. 1988 Quasi-helically symmetric toroidal stellarators. Phys. Lett. A 129 (2), 113117.CrossRefGoogle Scholar
Paul, E.J., Antonsen, T., Landreman, M. & Cooper, W.A. 2020 Adjoint approach to calculating shape gradients for three-dimensional magnetic confinement equilibria. Part 2. Applications. J. Plasma Phys. 86 (1), 905860103.CrossRefGoogle Scholar
Plunk, G. & Helander, P. 2018 Quasi-axisymmetric magnetic fields: weakly non-axisymmetric case in a vacuum. J. Plasma Phys. 84 (2), 905840205.CrossRefGoogle Scholar
Reiman, A. & Boozer, A.H. 1984 Island formation and destruction of flux surfaces in three-dimensional MHD equilibria. Phys. Fluids 27 (10), 24462454.CrossRefGoogle Scholar
Rodríguez, E. & Bhattacharjee, A. 2021 a Islands and current singularities in quasisymmetric toroidal plasmas. Phys. Plasmas 28 (9), 092506.CrossRefGoogle Scholar
Rodríguez, E. & Bhattacharjee, A. 2021 b Solving the problem of overdetermination of quasisymmetric equilibrium solutions by near-axis expansions. I. Generalized force balance. Phys. Plasmas 28 (1), 012508.CrossRefGoogle Scholar
Rodríguez, E. & Bhattacharjee, A. 2021 c Solving the problem of overdetermination of quasisymmetric equilibrium solutions by near-axis expansions. II. Circular axis stellarator solutions. Phys. Plasmas 28 (1), 012509.CrossRefGoogle Scholar
Rodríguez, E., Helander, P. & Bhattacharjee, A. 2020 Necessary and sufficient conditions for quasisymmetry. Phys. Plasmas 27 (6), 062501.CrossRefGoogle Scholar
Rodríguez, E., Sengupta, W. & Bhattacharjee, A. 2021 Generalized boozer coordinates: a natural coordinate system for quasisymmetry. Phys. Plasmas 28 (9), 092510.CrossRefGoogle Scholar
Sanchez, R., Hirshman, S.P., Ware, A.S., Berry, L.A. & Spong, D.A. 2000 Ballooning stability optimization of low-aspect-ratio stellarators. Plasma Phys. Control. Fusion 42 (6), 641653.CrossRefGoogle Scholar
Skovoroda, A.A. 2005 3D toroidal geometry of currentless magnetic configurations with improved confinement. Plasma Phys. Control. Fusion 47 (11), 19111924.CrossRefGoogle Scholar
Spitzer, L. 1958 The stellarator concept. Phys. Fluids 1 (4), 253264.CrossRefGoogle Scholar
Spong, D., Hirshman, S., Whitson, J., Batchelor, D., Sanchez, R., Carreras, B., Lynch, V., Lyon, J., Valanju, P., Miner, W., et al. 2000 Design studies of low aspect ratio quasi-omnigenous stellarators. Nucl. Fusion 40 (3Y), 563567.CrossRefGoogle Scholar
Tessarotto, M., Johnson, J.L., White, R.B. & Zheng, L. 1996 Quasi-helical magnetohydrodynamic equilibria in the presence of flow. Phys. Plasmas 3 (7), 26532663.CrossRefGoogle Scholar
Tessarotto, M., Johnson, J.L. & Zheng, L.J. 1995 Hamiltonian approach to the magnetostatic equilibrium problem. Phys. Plasmas 2 (12), 44994512.CrossRefGoogle Scholar
Zarnstorff, M., Berry, L., Brooks, A., Fredrickson, E., Fu, G., Hirshman, S., Hudson, S., Ku, L., Lazarus, E., Mikkelsen, D., et al. 2001 Physics of the compact advanced stellarator NCSX. Plasma Phys. Control. Fusion 43 (12A), A237.CrossRefGoogle Scholar
Figure 0

Table 1. Basic properties of QS measures. Summary of the basic formal features of the QS measures introduced in § 2. These include: whether the helicity needs to be specified as an input (I) or not (O), whether Boozer coordinates are required to evaluate it, whether $\boldsymbol {j}\boldsymbol {\cdot }\boldsymbol {\nabla }\psi =0$ is a necessary assumption and whether the measure is local (L) or global (G).

Figure 1

Figure 1. Schematic of the loop integral for $\partial _\alpha \mathcal {J}_\parallel$. Schematic of the path integral followed to compute the field line dependence of the second adiabatic invariant $\mathcal {J}_\parallel$. The contours match field lines and contours of constant $|\boldsymbol {B}|=B_t$.

Figure 2

Figure 2. Comparison of defined cost functions in quasisymmetric designs. Plots in logarithmic scale including a comparison of the cost functions $\hat {f}_B$, $\hat {f}_C$ and $\hat {f}_T$ for a number of quasisymmetric devices. The scatter points for each device indicate values at a number of equally spaced magnetic flux surfaces joined by lines in order from the centre to the edge of the configurations (the rightmost points represent the edge).

Figure 3

Figure 3. Correlation between edge values of cost functions and $\epsilon _\mathrm {eff}$ of different quasisymmetric configurations. Diagram showing the Spearman correlation between the edge values of the different cost functions and $\epsilon _\mathrm {eff}$ of the configurations in figure 2. The colour represents the coefficient (as does the diameter of the coloured circles).

Figure 4

Figure 4. Parameter space spanned by $n=2,m=1$ surface mode for $\hat {f}_B$, $\hat {f}_C$, $\hat {f}_T$ and $\epsilon _\mathrm {eff}$. The plots show, clockwise, the cost functions $\hat {f}_B$, $\hat {f}_C$, $\epsilon _\mathrm {eff}$ and $\hat {f}_T$ as a function of the $(2,1)$ surface modes. The parameter scan is performed using the map option of STELLOPT around the stellarator design in the footnote of p. 28. The minima for each of these are found (clockwise) at: $(-0.031,0.031)$, $(-0.031,0.031)$, $(-0.037,0.045)$ and $(-0.037,0.037)$.

Figure 5

Figure 5. Gradients in parameter space spanned by $n=2,m=1$ surface mode for $\hat {f}_B$, $\hat {f}_C$, $\hat {f}_T$ and $\epsilon _\mathrm {eff}$. The plots show, clockwise, the parameter space gradients of $\hat {f}_B$, $\hat {f}_C$, $\epsilon _\mathrm {eff}$ and $\hat {f}_T$ as a function of the $(2,1)$ surface modes. Scales of the gradients are significantly different between measures.

Figure 6

Figure 6. Comparison of cross-sections of optimisation results for different cost functions. Plot showing the cross-section of the stellarators as obtained from the optimisation using different forms of the cost function. The optimisation stops when the optimiser is unable to make further improvement (the characteristic change in the parameters during an iteration is $10^{-4}$$10^{-3}$). The broken line represents the starting point, while the black, blue and red lines represent the $f_B$, $f_T$ and $f_C$ optimisation results, respectively. The cross-sections are only presented in one half of the plane but the missing plane can be reconstructed symmetrically.

Figure 7

Table 2. Averaged metric values for optimised stellarators. Summary of the average value of the metric over the volume of the optimised stellarators with respect to the different cost functions.