Hostname: page-component-848d4c4894-hfldf Total loading time: 0 Render date: 2024-05-09T21:29:25.370Z Has data issue: false hasContentIssue false

Nonlinear drift-wave and energetic particle long-time behaviour in stellarators: solution of the kinetic problem

Published online by Cambridge University Press:  09 June 2023

Alessandro Zocco*
Affiliation:
Max Planck Institute for Plasma Physics, 17491 Greifswald, Germany
Alexey Mishchenko
Affiliation:
Max Planck Institute for Plasma Physics, 17491 Greifswald, Germany
Axel Könies
Affiliation:
Max Planck Institute for Plasma Physics, 17491 Greifswald, Germany
Matteo Falessi
Affiliation:
Center for Nonlinear Plasma Science and ENEA C. R. Frascati, Frascati, Italy
Fulvio Zonca
Affiliation:
Center for Nonlinear Plasma Science and ENEA C. R. Frascati, Frascati, Italy Institute for Fusion Theory and Simulation and Department of Physics, Zhejiang University, Hangzhou, PR China
*
Email address for correspondence: alessandro.zocco@ipp.mpg.de
Rights & Permissions [Opens in a new window]

Abstract

We propose a theoretical scheme for the study of the nonlinear interaction of drift-wave-like turbulence and energetic particles in stellarators. The approach is based on gyrokinetics, and features a separation of time and scales, for electromagnetic fluctuations, inspired by linear ballooning theory. Two specific moments of the gyrokinetic equation constitute the main equations of the system, which requires a full kinetic nonlinear solution. This is found iteratively, expanding in the smallness of the bounce-average radial drift frequency, and nonlinear $\boldsymbol {E}\times \boldsymbol {B}$ drift frequency, compared with the inverse time scales of the resonantly interacting energetic particles. Our analysis is therefore valid for neoclassically optimised stellators. The resummation of all iterative and perturbative nonlinear kinetic solutions is discussed in terms of Feynman diagrams. Particular emphasis is put on the role of collisionlessly undamped large-scale structures in phase space, the kinetic equivalent of zonal flows, i.e. phase-space zonal structures, and on wave-like fluctuations generated by energetic particles.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press

1. Introduction

The purpose of any magnetic confinement fusion device is to access energetically favourable scenarios where fusion reactions are self-sustained by alpha particles. This has not been attained yet, but the study of energetic particle minorities, produced by auxiliary heating, gives us a glimpse of what could be the behaviour of a burning plasma.

A burning plasma is expected to be a kinetic turbulent environment in which energetic particles experience resonant interactions with fluctuating electromagnetic fields. These are believed to be faithfully described by gyrokinetic theory (Frieman & Chen Reference Frieman and Chen1982). Here, fluctuations are assumed to vary slowly along and strongly across the confining magnetic field, while evolving on a time scale which is slow when compared with the gyromotion, but fast when compared with the long-time behaviour of the plasma density and temperature background. Thus, within a gyrokinetic description, plasma transport properties are determined through a turbulence average (in time and space) (Frieman & Chen Reference Frieman and Chen1982; Abel & Cowley Reference Abel and Cowley2013) which allows nonlinear fluxes to act as sources in the evolution of mean quantities. Thereby, in a sense, nonlinear gyrokinetic simulations provide insight into the detailed properties of kinetic turbulence in magnetised plasmas, but do not fully characterise them from the point of view of transport, and therefore possible reactor performance, unless one knows how fluctuations, on a long time scale, impact mean quantities such as temperature and density.

Among the many approaches that one can pursue to attack the problem of turbulent transport in magnetically confined plasmas (which is naturally multiscale, and tackled with an asymptotic expansion for small Larmor radii), the one reviewed by Chen & Zonca (Reference Chen and Zonca2019), perhaps more than others, kept the physics of fast particles at heart. In the first of a series of works there cited (Zonca et al. Reference Zonca, Chen, Briguglio, Fogaccia, Vlad and Wang2015), the question was proposed, and partially addressed, as to what would be the dominant contribution to the possible long-time behaviour of gyrokinetic fluctuations driven by wave–particle resonant interaction. By exploiting the existence of action-angle variables in axisymmetric toroidal geometry, Zonca et al. (Reference Zonca, Chen, Briguglio, Fogaccia, Vlad and Wang2015) introduced a way of integrating perturbed particle orbits by using the unperturbed action-angle variables, thus capturing cumulative wave–particle resonance effects of bounce- and transit-averaged processes on unperturbed particle motion, for sufficiently small nonlinear dynamics. This approach allows one to describe several nonlinear wave–particle phenomena, such as phase locking, resonance detuning and broadening. It gives a powerful means of predicting chorus emission in planetary magnetospheres (Zonca, Tao & Chen Reference Zonca, Tao and Chen2021b), but it also gives the basis for the study of what the authors call phase-space zonal structures, i.e. the counterpart of zonal flows (Rosenbluth & Hinton Reference Rosenbluth and Hinton1998) or, more generally, zonal field structures in phase space. These are structures that are constant over magnetic flux surfaces, thus, in many circumstances, immune to collisionless Landau damping. They play a major role in the regulation of turbulence levels (Rogers, Dorland & Kotschenreuther Reference Rogers, Dorland and Kotschenreuther2000; Diamond et al. Reference Diamond, Itoh, Itoh and Hahm2005).

Further advancing this line of research, Falessi & Zonca (Reference Falessi and Zonca2019) have focused on the long-time behaviour of phase-space zonal structures in axisymmetry. Their results have been successfully applied to particle-in-cell codes (Bottino et al. Reference Bottino, Falessi, Hayward-Schneider, Biancalani, Briguglio, Hatzky, Lauber, Mishchenko, Poli and Rettino2022), and completed with the evaluation of the mesoscale radial extent of fluctuations (Zonca et al. Reference Zonca, Chen, Falessi and Qiu2021a). These are evaluated as first-order corrections with a subsidiary limit of gyrokinetics that mimics the original radial envelope evaluation of ballooning modes (Connor, Hastie & Taylor Reference Connor, Hastie and Taylor1978), Alfvèn waves (Zonca & Chen Reference Zonca and Chen1992, Reference Zonca and Chen1993) and the ion-temperature-gradient-driven modes (Romanelli & Zonca Reference Romanelli and Zonca1993).

We are starting to see the first numerical results on the behaviour of energetic particles embedded into turbulence (Biancalani et al. Reference Biancalani, Bottino, Siena, Gürcan, Hayward-Schneider, Jenko, Lauber, Mishchenko, Morel and Novikau2021; Di Siena et al. Reference Di Siena, Görler, Poli, Bañón Navarro, Biancalani, Bilato, Bonanomi, Novikau, Vannini and Jenko2021; Carlevaro et al. Reference Carlevaro, Meng, Montani, Zonca, Hayward-Schneider, Lauber, Lu and Wang2022; Hayward-Schneider et al. Reference Hayward-Schneider, Lauber, Bottino and Mishchenko2022; Lu et al. Reference Lu, Meng, Hatzky, Hoelzl and Lauber2023), but, to date, the aforementioned fundamental theoretical aspects have only been covered in the context of tokamak physics (Zonca et al. Reference Zonca, Chen, Briguglio, Fogaccia, Vlad and Wang2015; Chen & Zonca Reference Chen and Zonca2016, Reference Chen and Zonca2019; Falessi & Zonca Reference Falessi and Zonca2019). A generalisation to stellarators is not straightforward mainly for two reasons. The first being that within a multiscale gyrokinetic approach in stellarators, the strong variation of the strength of the equilibrium magnetic field on flux surfaces is very likely more important than the radial variation of thermodynamic equilibrium quantities (Zocco et al. Reference Zocco, Plunk, Xanthopoulos and Helander2016; Zocco, Aleynikova & Xanthopoulos Reference Zocco, Aleynikova and Xanthopoulos2018a; Zocco, Plunk & Xanthopoulos Reference Zocco, Plunk and Xanthopoulos2020), the latter being the most important mesoscale effect encountered in asymmetry, and taken into account when evaluating the fluctuation radial envelope. The second problem encountered when extending the work of Falessi & Zonca (Reference Falessi and Zonca2019) to stellarators is the absence of a conserved toroidal canonical momentum, which complicates considerably the analysis.

Nevertheless, in this work, we find a way of extending to stellarators the kinetic theory first put forth by Zonca et al. (Reference Zonca, Chen, Briguglio, Fogaccia, Vlad and Wang2015). Whilst realising that we are engaging with innumerable aspects of a complex endeavour, we decide to particularly focus on the kinetic solution of the problem, devising a method, given the electromagnetic fields, to solve iteratively the gyrokinetic equation while retaining the nonlinear wave–particle interaction. The method is diagrammatic, and applies not only to the physics of energetic particles, but also to the study of micro-instability-driven transport as the electrostatic ion temperature gradient at marginality (Zocco et al. Reference Zocco, Xanthopoulos, Doerk, Connor and Helander2018b), for instance, when wave–particle resonance effects are important.

This work is organised as follows. In § 2 we introduce the basic gyrokinetic equation and the equations for fluctuating fields and set up the relation between wave-like fluctuations and the underlying kinetics. In § 3 we introduce the coordinates and specify some geometric features that allow us to solve the gyrokinetic equation. In § 4 we introduce its iterative diagrammatic solution. In §§ 5 and 6 we apply the kinetic solution to the evaluation of transport due to wave-like fluctuating fields, and the phase-space zonal structure. In § 7 we conclude.

2. Nonlinear transport from finite-$\beta$ fluctuations

In this section we highlight the mathematical formulation for the study of transport from electromagnetic fluctuations. Our starting point is the evaluation of the quasi-neutrality equation and the divergence of the perturbed plasma current (Frieman et al. Reference Frieman, Rewoldt, Tang and Glasser1980; Tang, Connor & Hastie Reference Tang, Connor and Hastie1980; Qin, Tang & Rewoldt Reference Qin, Tang and Rewoldt1998; Zonca & Chen Reference Zonca and Chen2014; Chen & Zonca Reference Chen and Zonca2016; Zocco, Helander & Connor Reference Zocco, Helander and Connor2015; Aleynikova & Zocco Reference Aleynikova and Zocco2017):

(2.1)\begin{equation} \sum_{s}q_{s}\int \,{\rm d}^{3}\boldsymbol{v}F_{0s}\frac{q_{s}\phi}{T_{0s}} =\sum_{s}q_{s}\int \,{\rm d}^{3}\boldsymbol{v}\sum_{\boldsymbol{k}} {\rm e}^{{\rm i}\boldsymbol{k}\boldsymbol{\cdot}\boldsymbol{r}}{\rm J}_{0}\left(a_{s}\right)\delta G_{s\boldsymbol{k}} \end{equation}

and

(2.2)\begin{align} B\boldsymbol{\nabla}_{{\parallel}}\frac{j_{{\parallel},\boldsymbol{k}}}{B}& ={-}\sum_{s}\frac{e_{s}^{2}}{T_{0s}}\int \,{\rm d}^{3}\boldsymbol{v}F_{0s}\left[\left(1-{\rm J}_{0}^{2}\left(a_{s}\right)\right)\frac{\partial}{\partial t}-{\rm i}\varOmega_{*,s}^{T}{\rm J}_{0}^{2}\left(a_{s}\right)\right]\phi_{\boldsymbol{k}}\nonumber\\ & \quad -\sum_{s}\frac{e_{s}^{2}}{T_{s}}\int \,{\rm d}^{3}\boldsymbol{v}F_{0s}\left(\frac{\partial}{\partial t}-{\rm i}\varOmega_{*,s}^{T}\right)4\mu B\frac{{\rm J}_{0}\left(a_{s}\right){\rm J}_{1}\left(a_{s}\right)}{a_{s}}\frac{T_{0s}}{q_{s}}\frac{\delta B_{{\parallel}\boldsymbol{k}}}{B}\nonumber\\ & \quad +{\rm i}\frac{c}{B}\hat{\boldsymbol{b}}\boldsymbol{\cdot}\boldsymbol{\boldsymbol{\kappa}\times\boldsymbol{k}} \sum_{s}\int \,{\rm d}^{3}\boldsymbol{v}m_{s}\left(\mu B+v_{{\parallel}}^{2}\right){\rm J}_{0}(a_{s})\delta G_{s\boldsymbol{k}}\nonumber\\ & \quad +\sum_{s}q_{s}\int \,{\rm d}^{3}\boldsymbol{v}{\rm J}_{0}(a_{s})\sum_{{\boldsymbol{k}}^{\prime}}\frac{c}{B}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{k}_{{\perp}}^{\prime}\times\boldsymbol{k}_{{\perp}}\left[{\rm J}_{0}(a_{s}^{\prime})\left(\phi_{\boldsymbol{k^{\prime}}}-\frac{v_{{\parallel}}}{c}A_{{\parallel}\boldsymbol{k}^{\prime}}\right)\right.\nonumber\\ & \quad+ \left.4\mu B\frac{{\rm J}_{1}(a_{s}^{\prime})}{a_{s}^{\prime}}\frac{T_{0s}}{q_{s}}\frac{\delta B_{{\parallel}\boldsymbol{k}^{\prime}}}{B}\right]\delta G_{s,\boldsymbol{k}-\boldsymbol{k}^{^{\prime}}}+B\sum_{s}\int \,{\rm d}^{3}\boldsymbol{v}\delta G_{\boldsymbol{k}}\boldsymbol{\nabla}_{{\parallel}}{\rm J}_{0}(a_{s}), \end{align}

where $F_{0s}$ is the equilibrium distribution function (taken to be Maxwellian), thus $f_{s}=F_{0s}+\delta f_{s}\equiv F_{0s}(1-q_{s}\phi (\boldsymbol {r},t)/T_{0s})+\delta G_{s}(\boldsymbol {R}_{s},\mu,\mathcal {E},t)+\mathcal {O}(\epsilon ^{2}),$ and $\delta f_{s}/F_{0s}\sim k_{\parallel }/k_{\perp }\sim \epsilon \equiv \rho _*=\rho _s/a,$ where $\rho _s$ is the Larmor radius and $a$ a macroscopic length. Here $\boldsymbol {R}_{s}=\boldsymbol {r}+\boldsymbol {v}_{\perp }\times \hat {\boldsymbol {b}}/\varOmega _{s}$ is the gyrocentre position, where $\boldsymbol {r}$ is the particle position, ${\varOmega _{s}=q_{s}B/(m_{s}c)},$ $\hat {\boldsymbol {b}}=\boldsymbol {B}/B,$ with $\boldsymbol {B}$ the equilibrium magnetic field, and $\mu =v_{\perp }^{2}/(2B),$ $\mathcal {E}=v^{2}/2$ the velocity-space coordinates. The Bessel function, ${\rm J}_{0}={\rm J}_{0}(a_{s}),$ with $a^2_{s}=2B\mu k_{\perp }^{2}/\varOmega _{s}^{2},$ relates the Fourier transform with respect to $\boldsymbol {R}_{s}$ of $\delta G_{s}$ with its Fourier transform with respect to $\boldsymbol {r}$ (velocity-space integrals in (2.1)–(2.2) are taken at constant $\boldsymbol {r}$). We also have

(2.3)\begin{equation} \varOmega_{*,s}^{T}=\frac{cT_{0s}}{Be_{s}}\hat{\boldsymbol{b}}\boldsymbol{\cdot}\boldsymbol{k}\times\frac{\boldsymbol{\nabla} F_{0s}}{F_{0s}}, \end{equation}

which takes account of diamagnetic effects, while $\boldsymbol {\kappa =}\hat {\boldsymbol {b}}\boldsymbol {\cdot }\boldsymbol {\nabla }\hat {\boldsymbol {b}}$ is the magnetic field curvature. Finally,

(2.4)\begin{equation} \frac{\delta B_{{\parallel},\boldsymbol{k}}}{B}={-}\frac{4{\rm \pi}}{c}\sum_{s}\int \,{\rm d}^{3}\boldsymbol{v}m_{s}2\mu B\frac{{\rm J}_{1}(a_{s})}{a_{s}}\delta G_{s,\boldsymbol{k}} \end{equation}

and

(2.5)\begin{equation} j_{{\parallel}\boldsymbol{k}}=\sum_{s}q_{s}\int \,{\rm d}^{3}\boldsymbol{v}v_{{\parallel}}{\rm J}_0(a_{s})\delta G_{s,\boldsymbol{k}}, \end{equation}

where $j_{\parallel }=(c/4{\rm \pi} )\hat {\boldsymbol {b}}\boldsymbol {\cdot }\boldsymbol {\nabla }\times (\boldsymbol {\nabla }\times \boldsymbol {A}).$ We take $\boldsymbol {\kappa }\approx \boldsymbol {\nabla } B/B,$ consistent with linear cancellations due to the perpendicular pressure balance (Tang et al. Reference Tang, Connor and Hastie1980; Hasegawa & Sato Reference Hasegawa and Sato1981; Zocco et al. Reference Zocco, Helander and Connor2015; Chen & Zonca Reference Chen and Zonca2016; Aleynikova & Zocco Reference Aleynikova and Zocco2017), thus allowing us to eliminate (but not neglect!) magnetic compressibility altogether. The final term in (2.2) can be neglected in most practical applications, since the current is mostly carried by the electrons (Connor et al. Reference Connor, Hastie and Taylor1978; Zocco et al. Reference Zocco, Helander and Connor2015; Aleynikova & Zocco Reference Aleynikova and Zocco2017). The kinetic information is contained in the velocity-space integrals of the function $\delta G_{s}.$

The function $\delta G_{s}$ follows the nonlinear Frieman–Chen equation (Frieman & Chen Reference Frieman and Chen1982) for the non-adiabatic part of the perturbed distribution function, $\delta G_{s}=\delta f_{s}+F_{0s}e_{s}\phi /T_{0s}$:

(2.6)\begin{align} & \left(\frac{\partial}{\partial t}+v_{{\parallel}}\boldsymbol{\nabla}_{{\parallel}}+\boldsymbol{v}_{d,s}\boldsymbol{\cdot}\boldsymbol{\nabla}\right)\delta G_{s}\nonumber\\ & \quad= \frac{e_{s}F_{0s}}{T_{0s}}\frac{\partial}{\partial t}\left\langle \chi\right\rangle _{\boldsymbol{R}_{s}}-\frac{c}{B}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{\nabla}\left\langle \chi\right\rangle _{\boldsymbol{R}_{s}}\times\boldsymbol{\nabla} F_{0_{s}}-\frac{c}{B}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{\nabla}\left\langle \chi\right\rangle _{\boldsymbol{R}_{s}}\times\boldsymbol{\nabla}\delta G, \end{align}

where $\langle \chi \rangle _{\boldsymbol {R}_{s}}=\sum _{\boldsymbol {k}}\langle \chi \rangle _{\boldsymbol {R}_{s},\boldsymbol {k}}\exp {\rm i}\boldsymbol {k}\boldsymbol {\cdot }\boldsymbol {R_{s}},$ with $\langle \chi \rangle _{\boldsymbol {R}_{s},\boldsymbol {k}}={\rm J}_{0}(a_{s})(\phi _{\boldsymbol {k}}-v_{\parallel }A_{\parallel \boldsymbol {k}}/c)$ being the gyroaveraged kinetic potential, $\boldsymbol {\nabla }=\partial /\partial \boldsymbol {R}_{s},$ $\boldsymbol {v}_{d_{s}}=-v_{\parallel }\boldsymbol {b}\times \boldsymbol {\nabla }(v_{\parallel }/\varOmega _{s})$ and

(2.7)\begin{equation} \frac{c}{B}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{\nabla}\chi\times\boldsymbol{\nabla}\delta G=\frac{c}{B}\left[\chi,\delta G\right]. \end{equation}

In a stellarator context, it is wise to introduce Boozer coordinates (Boozer Reference Boozer1981) $\boldsymbol {B}=\boldsymbol {\nabla }\psi \times \boldsymbol {\nabla }\theta +\iota \boldsymbol {\nabla }\varphi \times \boldsymbol {\nabla }\psi =G(\psi )\boldsymbol {\nabla }\varphi +I(\psi )\boldsymbol {\nabla }\theta +K(\psi,\theta,\varphi )\boldsymbol {\nabla }\psi.$ Here, the magnetic equilibrium satisfies force balance:

(2.8)\begin{equation} \frac{4{\rm \pi}}{c}{\boldsymbol J}\times{\boldsymbol B}=\frac{1}{\sqrt{g}}\left(\iota I'+G'-\iota\frac{\partial K}{\partial\theta}-\frac{\partial K}{\partial\varphi}\right)\boldsymbol{\nabla}\psi=4{\rm \pi}\boldsymbol{\nabla} p_{0}, \end{equation}

where $1/\sqrt {g}=B^{2}/(G+\iota I)\simeq B^{2}/G$ so that

(2.9)\begin{equation} \iota I^{\prime}+G^{\prime}-\iota\frac{\partial K}{\partial\theta}-\frac{\partial K}{\partial\varphi}=\frac{4{\rm \pi} G}{B^{2}}\frac{{\rm d}p_{0}}{{\rm d}\psi}, \end{equation}

and $p_{0}$ and $\boldsymbol {J}$ are the equilibrium pressure and plasma current, respectively.

It is furthermore useful to see that

(2.10)\begin{align} \boldsymbol{v}_{d}\boldsymbol{\cdot}\boldsymbol{\nabla}\delta G& ={-}v_{{\parallel}}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{\nabla}\left(\frac{v_{{\parallel}}}{\varOmega}\right)\times\boldsymbol{\nabla}\delta G\nonumber\\ & ={-}\frac{v_{{\parallel}}}{\sqrt{g}B}\left\{ G\left[\left(\frac{\partial}{\partial\psi}\frac{v_{{\parallel}}}{\varOmega}\right)\frac{\partial}{\partial\theta}-\left(\frac{\partial}{\partial\theta}\frac{v_{{\parallel}}}{\varOmega}\right)\frac{\partial}{\partial\psi}\right]\right.\nonumber\\ & \quad -I\left[\left(\frac{\partial}{\partial\psi}\frac{v_{{\parallel}}}{\varOmega}\right)\frac{\partial}{\partial\varphi}-\left(\frac{\partial}{\partial\varphi}\frac{v_{{\parallel}}}{\varOmega}\right)\frac{\partial}{\partial\psi}\right]\nonumber\\ & \quad \left.+K\left[\left(\frac{\partial}{\partial\varphi}\frac{v_{{\parallel}}}{\varOmega}\right)\frac{\partial}{\partial\theta}-\left(\frac{\partial}{\partial\theta}\frac{v_{{\parallel}}}{\varOmega}\right)\frac{\partial}{\partial\varphi}\right]\right\} \delta G\nonumber\\ & \equiv{-}v_{{\parallel}}\left[\frac{v_{{\parallel}}}{\varOmega},\delta G\right]. \end{align}

2.1. Nonlinear fluctuations

Equations (2.1)–(2.2) accommodate finite-$\beta$ Alfvénic and drift-wave ballooning-like instabilities. Here we set up a model for nonlinear transport associated with balloning-type modes supported by them (Chen & Zonca Reference Chen and Zonca2016).

As in ballooning theory (Connor et al. Reference Connor, Hastie and Taylor1978; Connor, Hastie & Taylor Reference Connor, Hastie and Taylor1979), one introduces an eikonal representation for perturbed quantities (Antonsen & Lane Reference Antonsen and Lane1980):

(2.11)\begin{equation} \delta Q(\psi,\theta,\varphi,t)=\delta\hat{Q}\left(\psi,\alpha(\theta,\varphi),\theta,t\right) \exp({{\rm i}S\left(\psi,\alpha(\theta,\varphi),t\right)}), \end{equation}

where $\alpha =\theta -\iota \varphi,$

(2.12)\begin{equation} \frac{1}{S}\frac{\partial S}{\partial t}\gg\frac{1}{\delta\hat{Q}}\frac{\partial\delta\hat{Q}}{\partial t} \end{equation}

and

(2.13)\begin{equation} \frac{1}{S}\boldsymbol{\nabla} S\gg\frac{1}{\delta\hat{Q}}\boldsymbol{\nabla}\delta\hat{Q}. \end{equation}

To leading order, (2.2) becomes a nonlinear drift-wave-ballooning equation:

(2.14)\begin{align} & B\boldsymbol{\nabla}_{{\parallel}}\frac{c}{4{\rm \pi}}k_{{\perp}}^{2}\frac{\delta\hat{A}_{{\parallel},\boldsymbol{k}}}{B}-{\rm i}\omega\sum_{s}\frac{e_{s}^{2}}{T_{s}}\int \,{\rm d}^{3}\boldsymbol{v}F_{0s}\left[1-{\rm J}_{0}^{2}\left(a_{s}\right)-\frac{\varOmega_{*,s}^{T}}{\omega}{\rm J}_{0}^{2}\left(a_{s}\right)\right]\phi_{\boldsymbol{k}}\nonumber\\ & \qquad +\frac{c}{B}\hat{\boldsymbol{b}}\boldsymbol{\cdot}\boldsymbol{\boldsymbol{\kappa}\times\boldsymbol{k}_{{\perp}}}\sum_{s}\int \,{\rm d}^{3}\boldsymbol{v}m_{s}\left(\mu B+v_{{\parallel}}^{2}\right){\rm J}_{0}(a_{s})\delta G_{s\boldsymbol{k}}\nonumber\\ & \quad ={\rm i}\,\mathcal{N}_{F,\boldsymbol{k}}, \end{align}

where $\boldsymbol {k}=\partial _{\psi }S\boldsymbol {\nabla }\psi +\partial _{\alpha }S\boldsymbol {\nabla }\alpha,$ ${\rm i}\omega =\partial _{t}S$ and $\mathcal {N}_{F}$ contains all the terms that are nonlinear in the field amplitudes:

(2.15)\begin{align} \mathcal{N}_{F,\boldsymbol{k}}& = \frac{c}{B}\sum_{\boldsymbol{k}}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{k}_{{\perp}}^{\prime}\times\boldsymbol{k}_{{\perp}}\sum_{s}q_{s}\int \,{\rm d}^{3}\boldsymbol{v}{\rm J}_{0}(a_{s}){\rm J}_{0}(a_{s}^{\prime})\hat \phi_{\boldsymbol{k^{\prime}}}\delta \hat G_{s,\boldsymbol{k}-\boldsymbol{k}^{^{\prime}}}\nonumber\\ & \quad -\frac{1}{B}\sum_{\boldsymbol{k}}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{k}_{{\perp}}^{\prime}\times\boldsymbol{k}_{{\perp}}\sum_{s}q_{s}\int \,{\rm d}^{3}\boldsymbol{v}{\rm J}_{0}(a_{s}){\rm J}_{0}(a_{s}^{\prime})\hat A_{{\parallel},\boldsymbol{k^{\prime}}}v_\parallel\delta \hat G_{s,\boldsymbol{k}-\boldsymbol{k}^{^{\prime}}}. \end{align}

In the linear limit, $\mathcal {N_{F,\boldsymbol {k}}}\rightarrow 0,$ (2.14) reduces to the equation used by Tang et al. in the study of linear kinetic ballooning modes in the intermediate-frequency regime (Tang et al. Reference Tang, Connor and Hastie1980; Aleynikova & Zocco Reference Aleynikova and Zocco2017). Similarly, quasi-neutrality remains virtually unchanged:

(2.16)\begin{equation} n_{i}\left(\sum_{s}\frac{q_{s}}{\tau_{s}}\frac{n_{s}}{n_{i}}\right)\frac{q_{s}\delta\hat{\phi}_{\boldsymbol{k}}}{T_{0e}}=\sum_{s}\frac{q_{s}}{n_{s}}\int \,{\rm d}^{3}\boldsymbol{v}{\rm J}_{0}(a_{s})\delta\hat{G}_{s,\boldsymbol{k}}. \end{equation}

Equations (2.14)–(2.16) constitute a local eigenvalue problem for the fields $\delta \hat {A}_{{{\parallel }}}$ and $\delta \hat {\phi }$ where nonlinearities are retained.

Antonsen & Lane (Reference Antonsen and Lane1980) have proposed a way of attacking this type of eigenvalue problems, by applying a method introduced by Weinberg (Reference Weinberg1962), and further explored by Dewar & Glasser (Reference Dewar and Glasser1983). One considers, for a given radial location and a given field line, the local eigenvalue

(2.17)\begin{equation} \omega_{l}=\omega_{l}\left(\psi,\alpha,S_{\psi},S_{\alpha}\right), \end{equation}

and determines $S_{\psi }$ and $S_{\alpha }$ by solving Hamilton equations

(2.18)\begin{equation} \frac{\partial\psi}{\partial\ell}=\frac{\partial\omega_{l}}{\partial S_{\psi}},\quad \frac{\partial\alpha}{\partial\ell}=\frac{\partial\omega_{l}}{\partial S_{\alpha}} \end{equation}

and

(2.19)\begin{equation} \frac{\partial S_{\psi}}{\partial\ell}={-}\frac{\partial\omega_{l}}{\partial\psi},\quad \frac{\partial S_{\alpha}}{\partial\ell}={-}\frac{\partial\omega_{l}}{\partial\alpha}, \end{equation}

where $\ell$ parametrises the trajectories that make $\omega _{l}$ constant. This determines $S_{\psi }$ and $S_{\alpha }.$ A global eigenvalue could further be determined, but we do not proceed along these lines.

Instead, we point out that, to next order in the

(2.20)\begin{equation} \omega_{l}^{{-}1}\delta\hat{Q}^{{-}1}\partial_{t}\delta\hat{Q}\sim k_{{\perp}}^{{-}1}\delta\hat{Q}^{{-}1}\boldsymbol{\nabla}\delta\hat{Q}\ll1 \end{equation}

expansion, one obtains the equation for the global structure of field amplitudes. The outcome of this calculation is a nonlinear Schrödinger-type equation that integrates the description of finite-$\beta$ fluctuations with nonlinear wave–particle effects. Zonca et al. (Reference Zonca, Chen, Falessi and Qiu2021a) have recently derived the expressions for the radial envelope of fluctuations in tokamak geometry by implementing a specific ordering between linear and nonlinear time scales. The radial envelope of fluctuations, being a higher-order perturbative quantity, will follow an equation in which corrections to the radial variation of equilibrium quantities will need to be considered. These ‘radially global’ corrections to the turbulence description supported by the gyrokinetic equation, (2.6), need to be considered extremely carefully in a stellarator, since the strong variation of the modulus of the magnetic field on the magnetic surface introduces an intermediate equilibrium scale, $\rho \ll L_B\ll R,$ which induces important ‘surface-global effects’ that have no real equivalent in axisymmetry (Zocco et al. Reference Zocco, Plunk, Xanthopoulos and Helander2016, Reference Zocco, Plunk and Xanthopoulos2020). Thus, the generalisation to non-sxisymmetry of the analysis of Zonca et al. (Reference Zonca, Chen, Falessi and Qiu2021a), while conceptually unproblematic, needs its own dedicated work. We therefore dedicate the rest of the article to the kinetic side of the problem, that is, the evaluation of the function $\delta G_{s}.$

3. Kinetics

In this section we solve the kinetic equation with nonlinear wave–particle interaction in stellarators. Before proceeding, a couple of technical comments are in order.

3.1. Choice of coordinates

Let us be more precise with the choice of coordinates. We introduce

(3.1)\begin{equation} \left(\begin{array}{@{}c@{}} \theta\\ \varphi \end{array}\right)\rightarrow\left(\begin{array}{@{}c@{}} \theta_{c}\\ \alpha_{c} \end{array}\right)=\left(\begin{array}{@{}c@{}} \theta\\ \varphi-q\theta \end{array}\right), \end{equation}

where $q=\iota ^{-1},$ so that

(3.2)\begin{equation} f(\psi,\theta,\varphi)\mapsto \hat f\left(\psi,\theta_{c}(\theta),\alpha_{c}(\theta,\varphi)\right). \end{equation}

Then

(3.3)\begin{equation} \left.\frac{\partial}{\partial\theta}\right|_{\varphi}\hat f\left(\psi,\theta_{c}(\theta),\alpha_{c}(\theta,\varphi)\right)=\left(\frac{\partial}{\partial\theta_{c}}-q\frac{\partial}{\partial\alpha_{c}}\right)\hat f\left(\psi,\theta_{c},\alpha_{c}\right) \end{equation}

and

(3.4)\begin{equation} \left.\frac{\partial}{\partial\varphi}\right|_{\theta}\hat f\left(\psi,\theta_{c}(\theta),\alpha_{c}(\theta,\varphi)\right)=\frac{\partial}{\partial\alpha_{c}}\hat f\left(\psi,\theta_{c},\alpha_{c}\right). \end{equation}

The purpose of these coordinates is to reduce the parallel streaming term operator into a first-order partial derivative in one angle only:

(3.5)\begin{equation} v_{{\parallel}}\boldsymbol{\nabla}_{{\parallel}} =\frac{v_{{\parallel}}\iota}{B\sqrt{g}}\frac{\partial}{\partial\theta_{c}}, \end{equation}

as can easily be verified. Then the gyrokinetic equation reads

(3.6)\begin{align} & \frac{\partial\delta G}{\partial t}+\frac{v_{{\parallel}}\iota}{B\sqrt{g}}\frac{\partial}{\partial\theta_{c}}\delta G+\frac{v_{{\parallel}}G}{B\sqrt{g}}\left(\frac{\partial\rho_{{\parallel}}}{\partial\theta_{c}}\frac{\partial}{\partial\psi}-\frac{\partial\rho_{{\parallel}}}{\partial\psi}\frac{\partial}{\partial\theta_{c}}\right)\delta G\nonumber\\ & \qquad +\iota^{{-}1}Bv_{{\parallel}}\left(\frac{\partial\rho_{{\parallel}}}{\partial\psi}\frac{\partial}{\partial\alpha_{c}}-\frac{\partial\rho_{{\parallel}}}{\partial\alpha_{c}}\frac{\partial}{\partial\psi}\right)\delta G={-}\frac{e}{m}\frac{\partial F_{0}}{\partial\mathcal{E}}\frac{\partial}{\partial t}\chi\nonumber\\ & \qquad -\frac{c}{\sqrt{g}B^{2}}\left[G\left(\frac{\partial\chi}{\partial\psi}\frac{\partial}{\partial\theta_{c}}-\frac{\partial\chi}{\partial\theta_{c}}\frac{\partial}{\partial\psi}\right)-\iota\sqrt{g}B^{2}\left(\frac{\partial\chi}{\partial\psi}\frac{\partial}{\partial\alpha_{c}}-\frac{\partial\chi}{\partial\alpha_{c}}\frac{\partial}{\partial\psi}\right)\right]\nonumber\\ & \qquad +\frac{c}{\sqrt{g}B^{2}}\left[G\frac{\partial\chi}{\partial\theta_{c}}\frac{\partial}{\partial\psi}+\iota\sqrt{g}B^{2}\frac{\partial\chi}{\partial\alpha_{c}}\frac{\partial}{\partial\psi}\right]F_{0}, \end{align}

where we introduced $\rho _{\parallel }=v_{\parallel }/\varOmega,$ we are dropping the symbol of gyroaverage and the species index and we are working in real space. The $\boldsymbol {v}_{d}\boldsymbol {\cdot }\boldsymbol {\nabla }\theta$ term has generated the

(3.7)\begin{equation} -\frac{v_{{\parallel}}G}{B\sqrt{g}}\frac{\partial\rho_{{\parallel}}}{\partial\psi}\frac{\partial}{\partial\theta_{c}} \end{equation}

contribution, and this can be discarded if compared with the streaming term (a common approximation). However, $\partial _{\psi }\rho _{\parallel }\partial _{\theta _{c}}\sim \partial _{\theta _{c}}\rho _{\parallel }\partial _{\psi }$ so, formally, $\boldsymbol {v}_{d}\boldsymbol {\cdot }\boldsymbol {\nabla }\theta$ can be neglected for

(3.8)\begin{equation} k_{\theta_c}\frac{\partial\rho_{{\parallel}}}{\partial\psi}\ll k_{\psi}\frac{\partial\rho_{{\parallel}}}{\partial\theta_{c}}, \end{equation}

where $k_{\theta _c}$ and $k_{\psi }$ are the characteristic fast-varying wavelengths of fluctuations, which are assumed to be of the same order. The physical meaning of this statement is that finite banana width effects are caused by the poloidal variation of the equilibrium quantities that enter the definition of $\rho _{\parallel }$. We stress here that $\rho _{\parallel }$ is an equilibrium quantity. While its ‘fast’ variation might impact mode structure, it is not allowed to interfere with the gyrokinetic ordering $k_{\parallel }/k_{\perp }\ll 1$. The term $\partial \rho _{\parallel }/\partial \alpha _{c}$ is purely non-axisymmetric, and cannot be neglected if $\partial _{\theta _{c}}\rho _{\parallel }\partial _{\psi }\sim \partial _{\alpha }\rho _{\parallel }\partial _{\psi }.$ The gyrokinetic equation that we study is therefore

(3.9)\begin{align} & \frac{\partial\delta G}{\partial t}+\frac{v_{{\parallel}}\iota}{B\sqrt{g}}\frac{\partial}{\partial\theta_{c}}\delta G+\frac{v_{{\parallel}}G}{B\sqrt{g}}\frac{\partial\rho_{{\parallel}}}{\partial\theta_{c}}\frac{\partial}{\partial\psi}\delta G-\iota^{{-}1}Bv_{{\parallel}}\frac{\partial\rho_{{\parallel}}}{\partial\alpha_{c}}\frac{\partial}{\partial\psi}\delta G={-}\frac{e}{m}\frac{\partial F_{0}}{\partial\mathcal{E}}\frac{\partial}{\partial t}\chi\nonumber\\ & \qquad -\frac{c}{\sqrt{g}B^{2}}\left[G\left(\frac{\partial\chi}{\partial\psi}\frac{\partial}{\partial\theta_{c}}-\frac{\partial\chi}{\partial\theta_{c}}\frac{\partial}{\partial\psi}\right)-\iota\sqrt{g}B^{2}\left(\frac{\partial\chi}{\partial\psi}\frac{\partial}{\partial\alpha_{c}}-\frac{\partial\chi}{\partial\alpha_{c}}\frac{\partial}{\partial\psi}\right)\right]\delta G\nonumber\\ & \qquad +\frac{c}{\sqrt{g}B^{2}}\left[G\frac{\partial\chi}{\partial\theta_{c}}\frac{\partial}{\partial\psi}-\iota\sqrt{g}B^{2}\frac{\partial\chi}{\partial\alpha_{c}}\frac{\partial}{\partial\psi}\right]F_{0}, \end{align}

where we are basically assuming that equilibrium quantities vary more strongly in $\alpha$ and $\theta$ than in $\psi.$ To better understand what this approximation means, we revert to the original Boozer coordinates to obtain

(3.10)\begin{align} & \frac{\partial\delta G}{\partial t}+\frac{v_{{\parallel}}}{B\sqrt{g}}\left(\frac{\partial}{\partial\varphi}+\iota\frac{\partial}{\partial\theta}\right)\delta G+\frac{v_{{\parallel}}}{B\sqrt{g}}\left(G\frac{\partial\rho_{{\parallel}}}{\partial\theta}-I\frac{\partial\rho_{{\parallel}}}{\partial\varphi}\right)\frac{\partial}{\partial\psi}\delta G\nonumber\\ & \qquad ={-}\left\{ \frac{e}{m}\frac{\partial F_{0}}{\partial\mathcal{E}}\frac{\partial}{\partial t}\chi+\frac{c}{B}\left[\chi,\delta G\right]+\frac{c}{B}\left[\chi,F_{0}\right]\right\} . \end{align}

The terms left that are proportional to $v_{\parallel }$ are recognisable as

(3.11)\begin{equation} \frac{v_{{\parallel}}}{B\sqrt{g}}\left(\frac{\partial}{\partial\varphi}+\iota\frac{\partial}{\partial\theta}\right)\rightarrow v_{{\parallel}}\boldsymbol{\nabla}_{{\parallel}}, \end{equation}

while the last term of the first line is

(3.12)\begin{equation} \frac{v_{{\parallel}}}{B\sqrt{g}}\left(G\frac{\partial\rho_{{\parallel}}}{\partial\theta}-I\frac{\partial\rho_{{\parallel}}}{\partial\varphi}\right)\frac{\partial}{\partial\psi}\rightarrow\boldsymbol{v}_{d}\boldsymbol{\cdot}\boldsymbol{\nabla}\psi\frac{\partial}{\partial\psi}. \end{equation}

In field-following coordinates, the left-hand side of (3.10) would then be

(3.13)\begin{equation} \frac{\partial\delta G}{\partial t}+v_{{\parallel}}\boldsymbol{\nabla}_{{\parallel}}\delta G+k_{\psi}\boldsymbol{v}_{d}\boldsymbol{\cdot}\boldsymbol{\nabla}\psi \delta G. \end{equation}

One can choose the poloidal angle to be the field-following coordinate, thus accounting for a slow variation of perturbations along the field line, in the spirit of the underlying gyrokinetic ordering. However, this fact does not preclude fluctuations from being most important at short wavelengths, that is, when

(3.14)\begin{equation} k_{\psi}\sim k_{{\alpha}_c}\sim k_{{\theta}_c}\sim n/R\sim m/R\sim\rho_{i}^{{-}1}\gg 1, \end{equation}

for any set or coordinates used, where $n$ and $m$ are some poloidal and toroidal wavenumbers. This discussion is largely concerned with the influence of the strong variations of magnetic equilibrium quantities in gyrokinetics, and is specific to stellarators. Alternatively, one could state that (3.10) can be used when the radial derivatives of the fluctuating fields are much larger than any other derivative, and $\boldsymbol {v}_{d}\boldsymbol {\cdot }\boldsymbol {\nabla }\alpha _c$ does not play a role in their destabilisation.

The inclusion in (3.10) of the $\boldsymbol {v}_d \boldsymbol {\cdot } \boldsymbol {\nabla } \theta$ and $\boldsymbol {v}_d \boldsymbol {\cdot } \boldsymbol {\nabla } \varphi$ components of the magnetic drift, which would drive the toroidal branch of the ion-temperature-gradient-driven instability, would not be problematic. This could be done by simply upgrading the ordering of the radial derivatives of $\rho _{\parallel }$. However, in this work, we rely on linear drives provided by wave–particle resonant excitation, thus naturally energetic particle modes, and work with (3.10) as it is. From now on, we refer to the form of (3.9) as ‘formulation A’, and to (3.10) as ‘formulation B’. The first is more useful to treat explicitly the streaming term for the nonlinear investigation of wave–particle nonlinear interaction, the second is more practical for the study of the late-time behaviour of zonal flows in a flux tube.

4. Solution of the kinetic problem

We consider formulation B (equation (3.10)), and introduce $\delta G=\exp (-H\partial _{\psi })\delta g$, to obtain the compact form

(4.1)\begin{align} & \frac{\partial}{\partial t}\delta g+v_{{\parallel}}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{\nabla}\delta g\nonumber\\ & \qquad ={-}\bar{v}_{\psi}\frac{\partial}{\partial\psi}\delta g-{\rm e}^{{\rm i}Q}\left\{ \frac{e}{m}\frac{\partial F_{0}}{\partial\mathcal{E}}\frac{\partial}{\partial t}\chi+\frac{c}{B} \left[\chi,{\rm e}^{-{\rm i}Q}\delta g\right]+\frac{c}{B}\left[\chi,F_{0}\right]\right\} , \end{align}

where $Q=-iH\partial _{\psi }$ (this radial derivative acts on fluctuating quantities only) and $H$ solves for the magnetic differential equation

(4.2)\begin{equation} v_{{\parallel}}\boldsymbol{\nabla}_{{\parallel}}H=\frac{v_{{\parallel}}}{\sqrt{g}B}\left[G(\psi)\frac{\partial}{\partial\theta}\left(\frac{v_{{\parallel}}}{\varOmega}\right)-I(\psi)\frac{\partial}{\partial\varphi}\left(\frac{v_{{\parallel}}}{\varOmega}\right)\right]-\bar{v}_{\psi}. \end{equation}

Here $\bar {v}_{\psi }=\bar {v}_{\psi }(\mathcal {E},\psi )\equiv \overline {\boldsymbol {v}_{d}\boldsymbol {\cdot }\boldsymbol {\nabla }\psi },$ since $\overline {v_{\parallel }\boldsymbol {\nabla }_{\parallel }H}=0.$ Equation (4.1) was also derived by Sugama & Watanabe (Reference Sugama and Watanabe2005), for a specific helical magnetic equilibrium. Our equation is completely general from the point of view of geometry, provided a solution to the magnetic differential equation is given. For this, see  Mishchenko, Helander & Könies (Reference Mishchenko, Helander and Könies2008). Let us introduce the ‘rippling time’, $\tau,$ which satisfies the equation

(4.3)\begin{equation} \left(\frac{\partial}{\partial\varphi}+\iota\frac{\partial}{\partial\theta}\right)\tau(\theta,\varphi)=\tau_b, \end{equation}

where $\tau _{b}^{-1}=v_{\parallel }/(B\sqrt {g}).$ In these new variables, the left-hand side of (4.1) becomes

(4.4)\begin{equation} \frac{\partial}{\partial t}\delta g+v_{{\parallel}} \boldsymbol{\hat{b}\boldsymbol{\cdot}\boldsymbol{\nabla}}\delta g= \left(\frac{\partial}{\partial t}+\frac{\partial}{\partial\tau}\right)\delta g. \end{equation}

We can then introduce the ripple-bounce expansion (Sugama & Watanabe Reference Sugama and Watanabe2005):

(4.5)\begin{equation} \delta g(\tau)=\sum_{l}h_{l}{\rm e}^{{\rm i}l\omega_{b}\tau},\,\,h_{l}=T_{b}^{{-}1}\oint \,{\rm d}\tau\delta g(\tau) {\rm e}^{-{\rm i}l\omega_{b}\tau}, \end{equation}

with $\omega _{b}=2{\rm \pi} T_{b}^{-1}$ and $T_{b}=\oint \,{\rm d}\theta _{c}^{\prime }\tau _{b}(\theta _{c}^{\prime }),$ where the $\oint$ integral is from $\theta _{1}$ to $\theta _{2}$ and back to $\theta _{1},$ along a trapped particle trajectory, where $\theta _{1}$ and $\theta _{2}$ are two consecutive bounce points whose exact definition depends on the class of trapped particles considered (Sugama & Watanabe Reference Sugama and Watanabe2005; Mishchenko, Helander & Könies Reference Mishchenko, Helander and Könies2008). Then, the ripple-bounce and Laplace transforms are applied to find

(4.6)\begin{equation} \int_{0}^{\infty}\,{\rm d}t{\rm e}^{{-}pt}\frac{1}{T_{b}}\oint \,{\rm d}\tau {\rm e}^{-{\rm i}l\omega_{b}\tau}\left(\frac{\partial}{\partial t}+\frac{\partial}{\partial\tau}\right)\delta g=\left(p+{\rm i}l\omega_{b}\right)\hat{h}_{l}(p)-h_{l}(0), \end{equation}

for the first term.

All other terms in the kinetic equation are better represented by Fourier modes. We introduce

(4.7)\begin{equation} \varPhi_{mn}=\varXi_{mn}^{l}\varPhi_{l} \end{equation}

and

(4.8)\begin{equation} \varPhi_{l}=\varTheta_{l}^{mn}\varPhi_{mn}, \end{equation}

where

(4.9)\begin{equation} \varTheta_{l}^{mn}=\frac{1}{\oint\tau_{b}}\oint \,{\rm d}\tau \exp({-{\rm i}(l\tau\omega_{b}+n\varphi(\tau)-m\theta(\tau))}) \end{equation}

and

(4.10)\begin{equation} \varXi_{mn}^{l}=\int_{0}^{2{\rm \pi}}\,{\rm d}\theta\int_{0}^{2{\rm \pi}}\,{\rm d}\varphi \exp\left({{\rm i}\left(l\tau(\theta,\varphi)+n\varphi-m\theta\right)}\right). \end{equation}

We also use the radial local approximation:

(4.11)\begin{equation} \chi(\psi,\theta,\varphi,t)=\sum_{k}\varPhi_{k}(\theta,\varphi,t){\rm e}^{{\rm i}k\psi} \end{equation}

and

(4.12)\begin{equation} \delta g(\psi,\theta,\varphi,t)=\sum_{k}h_{k}(\theta,\varphi,t){\rm e}^{{\rm i}k\psi}. \end{equation}

Then, the nonlinear equation expanded in bounce harmonics is

(4.13)\begin{equation} \frac{\partial h_{k}^{l}}{\partial t}+{\rm i}l\omega_{b}h_{k}^{l}={-}{\rm i}k\bar{v}_{\psi}h_{k}^{l}-\frac{e}{m}\frac{\partial F_{0}}{\partial\mathcal{E}}q_{l^{\prime}}\frac{\partial\varPhi_{k}^{l-l^{\prime}}}{\partial t}+\frac{\partial F_{0}}{\partial\mathcal{\psi}}H_{l,l^{\prime}}\varPhi_{k}^{l^{\prime}}+\mathcal{N}_{k^{\prime}}^{ll^{\prime}l^{\prime\prime}}\varPhi_{k^{\prime}}^{l^{\prime}}h_{k-k^{\prime}}^{l^{\prime\prime}},\end{equation}

where

(4.14)\begin{align} H_{l,l^{\prime}}={\rm i}\frac{c}{\sqrt{g}B^{2}}\left(nI+mG\right)q_{l^{\prime\prime}}\varTheta_{l-l^{\prime\prime}}^{mn}\varXi_{l^{\prime}}^{mn}, \end{align}
(4.15)\begin{align} \mathcal{N}_{k^{\prime}}^{ll^{\prime}l^{\prime\prime}}=\frac{c}{\sqrt{g}B^{2}}\left(\mathcal{S}_{k^{\prime}}^{ll^{\prime}l^{\prime\prime}}+{\rm i}\mathcal{R}^{ll^{\prime}l^{\prime\prime}}kk^{\prime}\right), \end{align}

with

(4.16)\begin{equation} \mathcal{S}_{k^{\prime}}^{ll^{\prime}l^{\prime\prime}}=\left\{ G\left[k^{\prime}m^{\prime}-\left(k-k^{\prime}\right)m\right]+I\left[k^{\prime}n^{\prime}-\left(k-k^{\prime}\right)n\right]\right\} \varXi_{l^{\prime}}^{mn}\varXi_{l^{\prime\prime}}^{mn}\varTheta_{l}^{m+m^{{{\prime}}},n+n^{\prime}} \end{equation}

and

(4.17)\begin{equation} \mathcal{R}^{ll^{\prime}l^{\prime\prime}}\varPhi_{k^{\prime}}^{l^{\prime}}h_{k-k^{\prime}}^{l^{\prime\prime}}=\left(mG+nI\right)Q_{l_{\prime}}\varXi^{l^{\prime}}_{mn}\varTheta_{l}^{m+m^{\prime}+m^{\prime\prime},n+n^{\prime}+n^{\prime\prime}}\varXi_{l^{\prime\prime}}^{m^{\prime}n^{\prime}}\varXi_{l^{\prime\prime\prime}}^{m^{\prime\prime}n^{\prime\prime}}\varPhi_{k^{\prime}}^{l^{\prime\prime}}h_{k-k^{\prime}}^{l^{\prime\prime\prime}} \end{equation}

and $q_{l}$ are the coefficients of the bounce expansion of $\exp [{\rm i}Q].$ We use the Einstein convention for repeated indices.

Dynamically, for any $l,$ we choose

(4.18)\begin{equation} p\sim l\omega_{b}\gg k\bar{v}_{\psi},\omega_{\boldsymbol{E}\times\boldsymbol{B}}. \end{equation}

This means that wave–particle resonance effects are kept to all orders. However, subsidiary expansions in $p/(l \omega _b)$ are allowed. In particular, one could consider the ion sound approximation, $p \gg l\omega _{b,i}$, and derive the fluid ion-temperature-gradient-driven instability (after neglecting nonlinearities). The bounce-averaged radial drift is supposed to be small since we are not interested in unoptimised machines. Our ordering is most appropriate for turbulence driven by high-frequency resonant phenomena, such as those caused by energetic particle (or electron) instabilities, or for weak turbulence close to marginality (where resonant effects are important). Strongly driven fluid regimes would require consideration of $p\sim \omega _{\boldsymbol E \times \boldsymbol B},$ and at present we are not in the position to conclude if our solution can efficiently describe them.

Let us now introduce a perturbative nonlinear solution. In order to explore all time scales supported by our ordering, and in particular long-time turbulent behaviour, such solution must be known to all orders. We treat perturbatively the stellarator contribution and the nonlinearity, in a fashion somewhat similar to that of Al'tshu’l & Karpman (Reference Al'tshu’l and Karpman1966) for the study of nonlinear oscillations in collisionless plasmas. As already stated, the same type of treatment has proved fruitful for the study of phase-space zonal structures in tokamaks (Zonca et al. Reference Zonca, Chen, Briguglio, Fogaccia, Vlad and Wang2015; Falessi & Zonca Reference Falessi and Zonca2019). Reverting to (4.13), we immediately realise that its solution is iterative. Indeed, after introducing the Laplace transform, we find

(4.19)\begin{equation} \left(p+{\rm i}l\omega_{b}\right)\hat{h}_{k}^{(n)l}={-}{\rm i}k\bar{v}_{\psi}\hat{h}_{k}^{(n-1)l}+\mathcal{N}_{k^{\prime}}^{ll^{\prime}l^{\prime\prime}}\int \,{\rm d}p^{\prime}\hat{\varPhi}_{k^{\prime}}^{l^{\prime}}(p^{\prime})\hat{h}_{k-k^{\prime}}^{(n-1)l^{\prime\prime}}(p-p^{\prime}), \end{equation}

where

(4.20)\begin{equation} \frac{\hat{h}_{k}^{(n+1)l}}{\hat{h}_{k}^{(n)l}}\sim \frac{\omega_{\boldsymbol{E}\times\boldsymbol{B}}}{l \omega_b}\sim \frac{k {\bar{v}}_{\psi}}{l \omega_b} \ll 1. \end{equation}

This solution has a very attractive diagrammatic interpretation. Given an $\hat {h}^{(n-1)},$ the next-order iterative solution is the creation of a new field $\hat {h}_{k}^{l,(n)}=\hat {h}_{st}^{(n)}+\hat {h}_{nl}^{(n)},$ given by two modifications of the momentum-space propagator $(p+{\rm i}l\omega _{b})^{-1},$ that is,

(4.21)\begin{equation} \hat{h}_{k}^{l,(n)}=\frac{1}{p+{\rm i}l\omega_{b}}\left\{ \text{stellarator} + \text{nonlinearity}\right\}. \end{equation}

The first term is obtained via interaction of $\hat {h}_{k}^{(n-1)}$ at the vertex $-{\rm i}k\bar {v}_{\psi }$ (resulting in the emission of the curly propagator emerging from $-{\rm i}k\bar {v}_{\psi }$ in figure 1); the other is given by the annihilation of $\hat {h}^{(n-1)}$ and $\varPhi$ (see figure 1). At each step, the full result is the sum of all the fields produced by the iteration. We analyse in more detail the nonlinearity of the first-order solution of figure 1. This is

(4.22)\begin{equation} \left(p+{\rm i}l\omega_{b}\right)\hat{h}_{nl}^{(n)}=\mathcal{N}_{k^{\prime}}^{ll^{\prime}l^{\prime\prime}}\int \,{\rm d}p^{\prime}\hat{\varPhi}_{k^{\prime}}^{l^{\prime}}(p^{\prime})\hat{h}_{k-k^{\prime}}^{(n-1)l^{\prime\prime}}(p-p^{\prime}). \end{equation}

The right-hand side is a state characterised by two momentum-like quantities $(p,k)$ and an index $l.$ This is the result of the interaction at the circular vertex of figure 1 of two incoming fields: a $\hat {\varPhi }^{l^{\prime }}$ with ‘momentum’ $(p^{\prime },k^{\prime })$ and an $\hat {h}^{(n-1)l^{\prime \prime }}$ with ‘momentum’ $(p-p^{\prime },k-k^{\prime }).$ At the vertex momenta are conserved. The interaction is integral and its action mixes bounce harmonics via the $\mathcal {N}_{k^{\prime }}^{ll^{\prime }l^{\prime \prime }}$ tensor.

Figure 1. First-order iterative solution of (4.13).

Figure 2. Second-order iterative solution of (4.13).

Perturbatively, we might write, after introducing the regularised zeroth-order solution

(4.23)\begin{equation} \tilde{h}_{k}^{(0)l}=\left(p+{\rm i}l\omega_{B}\right)\hat{h}_{k}^{(0)l}, \end{equation}

with

(4.24)\begin{equation} \hat{h}_{k}^{(0)l}=\frac{1}{p+{\rm i}l\omega_{B}}\left\{ -\frac{e}{m}\frac{\partial F_{0}}{\partial\mathcal{E}}q_{l^{\prime}}\frac{\partial\varPhi_{k}^{l-l^{\prime}}}{\partial t}+\frac{\partial F_{0}}{\partial\mathcal{\psi}}H_{l,l^{\prime}}\varPhi_{k}^{l^{\prime}}\right\} , \end{equation}

the first-order solution

(4.25)\begin{equation} \tilde{h}_{k}^{(1)l}=\frac{-{\rm i}k\bar{v}_{\psi}}{\left(p+{\rm i}l\omega_{B}\right)^{2}}\tilde{h}_{k}^{(0)l}+\frac{\mathcal{N}_{k^{\prime}}^{ll_{1}l_{2}}}{p+{\rm i}l\omega_{B}}\int \,{\rm d}p_{1}\frac{\hat{\varPhi}_{k_{1}}^{l_{1}}(p_{1})}{p-p_{1}+{\rm i}l_{2}\omega_{b}}\tilde{h}_{k-k_{1}}^{(0)l_{2}}(p-p_{1}). \end{equation}

Then, the second-order solution is

(4.26)\begin{align} & \tilde{h}_{k}^{(2)l}=\frac{-{\rm i}k\bar{v}_{\psi}}{\left(p+{\rm i}l\omega_{B}\right)}\left\{ \frac{-{\rm i}k\bar{v}_{\psi}}{\left(p+{\rm i}l\omega_{B}\right)^{2}}\tilde{h}_{k}^{(0)l}+\frac{\mathcal{N}_{k^{\prime}}^{ll_{1}l_{2}}}{p+{\rm i}l\omega_{B}}\int \,{\rm d}p_{1}\frac{\hat{\varPhi}_{k_{1}}^{l_{1}}(p_{1})}{p-p_{1}+{\rm i}l_{2}\omega_{b}}\tilde{h}_{k-k_{1}}^{(0)l_{2}}(p-p_{1})\right\}\nonumber\\ & \qquad +\frac{\mathcal{N}_{k^{\prime}}^{ll_{1}l_{2}}}{p+{\rm i}l\omega_{B}}\int \,{\rm d}p_{1}\hat{\varPhi}_{k_{1}}^{l_{1}}(p_{1})\left\{ \frac{-{\rm i}k\bar{v}_{\psi}}{\left(p-p_{1}+{\rm i}l_{2}\omega_{B}\right)^{2}}\tilde{h}_{k-k_{1}}^{(0)l_{2}}(p-p_{1})\right.\nonumber\\ & \qquad+ \left.\frac{\mathcal{N}_{k^{\prime}}^{l_{2}l_{1}l_{3}}}{p-p_{1}+{\rm i}l_{2}\omega_{B}}\int \,{\rm d}p_{2}\frac{\hat{\varPhi}_{k_{1}}^{l_{1}}(p_{2})}{p-p_{1}-p_{2}+{\rm i}l_{3}\omega_{b}}\tilde{h}_{k-k_{1}-k_{2}}^{(0)l_{3}}(p-p_{1}-p_{2})\right\} , \end{align}

and so on. Here all the poles are now expressed explicitly. For the $n$th iteration, we generate $2^{n}$ graphs. In the following we show how to exploit the newly found nonlinear solution for the transport studies introduced in § 2.

5. Wave-like fluctuations and transport

At this point, the nonlinear theory of energetic particles and drift-wave-like modes is assumed to be given as an input. This provides mode frequencies and amplitudes, as explained in § 2. Then, for all perturbations

(5.1)\begin{equation} \varPhi_{k}(\theta,\varphi,t)=\phi^{\nu}{\rm e}^{{\rm i}\omega_{\nu}t}, \end{equation}

where $\omega _{\nu }$ is the local eigenvalue (2.17) of the system of equations for the fields (2.14)–(2.16) and $\phi ^{\nu }$ solves for the nonlinear Schrödinger-type equation discussed in § 2 (notice that there is no summation for repeated indices). After applying the Laplace transform, one obtains

(5.2)\begin{equation} \hat{\varPhi}_{k}^{l}=\varTheta_{mn}^{l}\frac{\phi_{mn,k}^{\nu}}{p+{\rm i}\omega_{\nu}}\equiv\frac{\zeta_{k,\nu}^{l}}{p+{\rm i}\omega_{\nu}}, \end{equation}

and all the convolution $p$ integrals in (4.19) can be performed by applying the Cauchy theorem and evaluating the residuals. To see specific stellarator nonlinear behaviour, we need to evaluate the second-order correction $\hat {h}_{k}^{(2)l}.$ The result is the sum of four different combinations of momentum-space propagators which look like

(5.3)\begin{align} \frac{k^{2}\bar{v}_{{{\psi}}}^{2}}{\left(p+{\rm i}l\omega_{b}\right)^{3}}& \div\frac{k\bar{v}_{{{\psi}}}\mathcal{N}\zeta}{\left(p+{\rm i}l\omega_{b}\right)^{2}\left(p-p_{0}\right)}\div\frac{k\bar{v}_{{{\psi}}}\mathcal{N}\zeta}{\left(p+{\rm i}l\omega_{b}\right)\left(p-p_{0}\right)^{2}}\nonumber\\ & \div\frac{\mathcal{N}^{2}\zeta^{2}}{\left(p+{\rm i}l\omega_{b}\right)\left(p-p_{0}\right)\left(p-p_{1}\right)},\end{align}

where $p_{0}=-{\rm i}(l_{2}\omega _{b}+\omega _{\nu })$ and $p_{1}=-{\rm i}(\omega _{b}l_{3}+2\omega _{\nu }).$ How such poles generate is highlighted in Appendix B.

What is left is a sum of momentum-space propagators that combine the available vertices of the theory, the stellarator-specific $k\bar {v}_{{{\psi }}}$ and that coming from nonlinear interaction $\mathcal {N}\zeta$. The three types of second-order propagators are shown in figure 3.

Figure 3. The three types of momentum-space second-order propagators for wave-like perturbations when $-{\rm i} l \omega _b =p_0=p_1$. Signs are taken into account by arrows, and quantities are conserved at the vertices. At this order, stellarator corrections and nonlinearity interact.

Diagrammatically, the first term of (5.3) is a stellarator self-energy loop. The second and the third terms are stellarator radiative corrections to a ‘luftballon’ diagram and the last is a nonlinear self-energy.

(5.4)

 Keeping this in mind, we are able to construct all possible solutions to all orders. We notice that the sum of all the two-vertex functions in (

5.4

) gives the analogue of the radiative corrections to the electron mass in quantum electrodynamics (Peskin & Schroeder Reference Peskin and Schroeder1995) (sum à-la-Dyson). The choice of specific resonant conditions allows us to evaluate which diagram is dominant in the sum, at each order.

A physics-based choice of resonances that dominate the sum is the following. Given a pole $p_{0},$ we consider a resonant condition as $p=p_{0}+\delta p,$ with $\delta p\sim k\bar {v}_{\psi }\ll 1,$ and

(5.5)\begin{equation} k\bar{v}_{\psi}\ll\mathcal{N}\zeta\ll p_{0}. \end{equation}

That is, for a slowly evolving turbulence, and an optimised machine, we can select which diagrams enter the Dyson summation in the final transport theory.

The perturbative nonlinear approach is a useful technique that helps in gaining insight into the problem. However, an equation for the complete infinite sum is preferable. The fact that the perturbative nonlinear solution can be constructed to all orders allows us to do so.

6. Phase-space zonal structure equation

Let us then focus our attention on the component of the perturbed distribution function that contributes to a ‘redefinition’ of the equilibrium quantities, in the spirit of what was originally proposed by Zonca et al. (Reference Zonca, Chen, Briguglio, Fogaccia, Vlad and Wang2015), and then derived for axisymmetric systems by Falessi & Zonca (Reference Falessi and Zonca2019). That is, we need to identify the specific component of the fluctuating distribution function which tends to have structure at the equilibrium scale and, very importantly, is not damped or suppressed by resonantly kinetic effects. Within our treatment, this is the $l=0$ bounce-harmonic component, as (4.13) shows, since here the wave–particle resonant term (the second term on the left-hand side) is absent, in total analogy with the more familiar case first studied by Rosenbluth & Hinton (Reference Rosenbluth and Hinton1998) (which we reproduce in Appendix A). Let us then consider the infinite sum

(6.1)\begin{equation} p\sum_{n=0}^{\infty}\hat{h}_{k}^{l_{0},(n)}\equiv p\hat{H}_{k}^{l_{0}}(p)=p\left(\hat{h}_{k}^{l_{0},(0)}+\hat{h}_{k}^{l_{0},(1)}+\sum_{n\geq2}^{\infty}\hat{h}_{k}^{l_{0},(n)}\right), \end{equation}

where we notice that stellarator corrections drive the nonlinearity in the perturbed distribution function, the terms $h\,h,$ only beyond second order (see the second line of the diagram in figure 2). We then iterate the perturbative $n$th solution twice, in order to isolate the linear drive and the $\phi h^{(0)}$ ‘nonlinearity’:

(6.2)\begin{gather} (p+{\rm i} l_0\omega_b)\hat{h}_{k}^{l_{0},(n)}=\int\int\frac{{\rm d}p_{1}\,{\rm d}p_2}{p-p_{1}+{\rm i} l_0 \omega_b} \left\{ -{\rm i}k\bar{v}_{\psi}\delta_{k_{1},0}\delta_{l_2,l_0}\delta(p_{1})+\mathcal{N}_{k_{1}}^{l_{0}l_{1}l_{2}}\hat{\varPhi}_{k_{1}}^{l_{1}}(p_{1})\right\}\nonumber\\ \times \left\{ -{\rm i}(k-k_1)\bar{v}_{\psi}\delta_{k_{2},0}\delta_{l_4,0}\delta(p_{2})+\mathcal{N}_{k_{2}}^{l_{2}l_{3}l_{4}}\hat{\varPhi}_{k_{2}}^{l_{3}}(p_{2})\right\} \hat{h}_{k-(k_{1}+k_{2})}^{l_{4},(n-2)}(p-p_1-p_2). \end{gather}

Plugging this result into (6.1), we obtain

(6.3)\begin{align} p\hat{H}_{k}^{0}(p)& =p\left(\hat{h}_{k}^{0,(0)}+\hat{h}_{k}^{0,(1)}\right)\nonumber\\ & \quad+ \int\int\frac{{\rm d}p_{1}\,{\rm d}p_{2}}{p-p_{1}}\left\{ -{\rm i}k\bar{v}_{\psi}\delta_{k_{1},0}\delta_{l_{2},0}\delta(p_{1})+\mathcal{N}_{k_{1}}^{0\,l_{1}l_{2}}\hat{\varPhi}_{k_{1}}^{l_{1}}(p_{1})\right\} \nonumber\\ & \quad\times \left\{ -{\rm i}(k-k_{1})\bar{v}_{\psi}\delta_{k_{2},0}\delta_{l_{4},0}\delta(p_{2})+\mathcal{N}_{k_{2}}^{l_{2}l_{3}l_{4}}\hat{\varPhi}_{k_{2}}^{l_{3}}(p_{2})\right\} \sum_{n\geq2}\hat{h}_{k-(k_{1}+k_{2})}^{l_{4},(n-2)}(p-p_{1}-p_{2}), \end{align}

where we set $l_0=0,$ since we are interested in the zonal part of the resummed perturbed distribution function. On the other hand, $\sum _{n\geq 2}^{\infty }\hat {h}_{k-(k_{1}+k_{2})}^{l_{4},(n-2)}=\hat {H}_{k-(k_{1}+k_{2})}^{l_{4}}(p)$; thus, for $l_{0}\equiv 0,$

(6.4)\begin{align} p\hat{H}_{k}^{0}(p)& =p\left(\hat{h}_{k}^{0,(0)}+\hat{h}_{k}^{0,(1)}\right)\nonumber\\ & \quad +\int \int\frac{{\rm d}p_{1} \,{\rm d}p_2}{p-p_{1}}\left\{ -{\rm i}k\bar{v}_{\psi}\delta_{k_{1},0}\delta_{l_2,0}\delta(p_{1})+\mathcal{N}_{k_{1}}^{0\,l_{1}l_{2}}\hat{\varPhi}_{k_{1}}^{l_{1}}(p_{1})\right\}\nonumber\\ & \quad\times \left\{ -{\rm i}(k-k_1)\bar{v}_{\psi}\delta_{k_{2},0}\delta_{l_4,0}\delta(p_{2})+\mathcal{N}_{k_{2}}^{l_{2}l_{3}l_{4}}\hat{\varPhi}_{k_{2}}^{l_{3}}(p_{2})\right\} \hat{H}^{l_4}_{k-(k_{1}+k_{2})}(p-p_1-p_2). \end{align}

This is the equivalent of the Dyson equation within the Al’tshu'l–Karpman treatment (their equation 2.23), which is quadratic in the field amplitude, but linear for the phase-space zonal structure, that is, the collisionlessly undamped ($l_0=0$) bounce harmonic. Equation (6.4) determines the nonlinear distortion to the local kinetic equilibrium, and its moments describe the sought-after transport theory which embeds nonlinear wave–particle resonant interactions which dominate energetic particle physics.

7. Discussion and conclusion

In this article we have introduced a new scheme for the study of the nonlinear interaction between drift-wave-like turbulence and energetic particles in stellarator geometry. The theory is fully electromagnetic. The equations for the fluctuating fields were derived by taking moments of the gyrokinetic equation and using an ordering inspired by linear ballooning theory. We therefore describe fluctuations as strongly varying across the equilibrium magnetic field, slowly varying along it, and characterised by an envelope that encapsulates some of their global features. The solution of the nonlinear distribution function is found by applying an iterative procedure which is based of the smallness of the bounce-averaged radial drift frequency and nonlinear $\boldsymbol {E}\times \boldsymbol {B}$ frequency. The resonant wave–particle interaction is treated with a mixed bounce harmonics/Fourier expansion. We introduced a diagrammatic interpretation of the iterative solutions, which can give guidance in the resummation of all perturbative nonlinear solutions derived at each iteration. We present a second-order perturbative nonlinear solution, for the case of wave-like fluctuating perturbations, and discuss the interplay between nonlinearity and specific stellarator features. We find that a finite bounce-averaged radial drift can couple to the nonlinearity in the fields and contribute to the evolution of turbulence. Finally we extended the concept of phase-space zonal structure to stellarator geometry, deriving the equivalent of the Dyson equation as proposed by Al'tshu’l & Karpman (Reference Al'tshu’l and Karpman1966) in the theory of nonlinear electrostatic fluctuations in plasmas. The result provides the basis for transport studies in which kinetically undamped nonlinear fluctuations contribute to the redefinition of the equilibrium distribution function. This behaviour is essential for the description of transport dynamics dominated by energetic particle physics.

Acknowledgements

A.Z. is grateful to Dr D. Rosco for illuminating discussions on the resummation of singularities that contributed to the development of the diagrammatic solutions here presented.

Editor A. Schekochihin thanks the referees for their advice in evaluating this article.

Funding

This work has been carried out within the framework of the EUROfusion Consortium, funded by the European Union via the Euratom Research and Training Programme (grant agreement nos. 633053 and 101052200 EUROfusion). Views and opinions expressed are, however, those of the authors only and do not necessarily reflect those of the European Union or the European Commission. Neither the European Union nor the European Commission can be held responsible for them.

Declaration of interests

The authors report no conflict of interest.

Appendix A. Linear theory and the Rosenbluth–Hinton response

We show how to recover the seminal result of Rosenbluth & Hinton (Reference Rosenbluth and Hinton1998) from our treatment. We consider the auxiliary functions $\delta g_{1}$ and $\delta g_{2},$ so that $\delta g=\delta g_{1}+\delta g_{2},$ where $\delta g_{1}$ follows the nonlinear gyrokinetic equation, but it is driven by flux-surface-averaged fluctuations only:

(A1)\begin{align} & \left(\frac{\partial}{\partial t}+\bar{v}_{\psi}\frac{\partial}{\partial\psi}+v_{{\parallel}}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{\nabla}\right)\delta g_{1}\nonumber\\ & \qquad={-}\left\{ \frac{e}{m}\frac{\partial F_{0}}{\partial\mathcal{E}} {\rm e}^{{\rm i}Q}\frac{\partial}{\partial t}\left\langle \chi\right\rangle _{\psi}+\frac{c}{\sqrt{g}B^{2}}\left\langle \chi\right\rangle _{\psi}^{\prime} {\rm e}^{{\rm i}Q}\left[G(\psi)\frac{\partial}{\partial\theta}-I(\psi)\frac{\partial}{\partial\varphi}\right] {\rm e}^{-{\rm i}Q}\delta g{}_{1}\right\} . \end{align}

Plugging back the result for $\delta G_{1}$ into the full gyrokinetic equation, one obtains the equation for $\delta G_{2}$:

(A2)\begin{align} & \left(\frac{\partial}{\partial t}+\bar{v}_{\psi}\frac{\partial}{\partial\psi} +v_{{\parallel}}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{\nabla}+\frac{c}{B}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{\nabla}\chi\times\boldsymbol{\nabla}\right)\delta g_{2}\nonumber\\ & \quad=-{\rm e}^{{\rm i}Q}\left\{ \frac{e}{m}\frac{\partial F_{0}}{\partial\mathcal{E}}\frac{\partial}{\partial t}\left(\chi-\left\langle \chi\right\rangle _{\psi}\right)+\frac{c}{B}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{\nabla}\chi\times\boldsymbol{\nabla} F_{0}\right.\nonumber\\ & \qquad \left.+\frac{c}{B}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{\nabla}\left(\chi-\left\langle \chi\right\rangle _{\psi}\right)\times\boldsymbol{\nabla} {\rm e}^{-{\rm i}Q}\delta g{}_{1}\vphantom{\frac{\partial F_{0}}{\partial\mathcal{E}}}\right\}. \end{align}

So, the non-adiabatic response is $\delta G=\exp (-{\rm i}Q)(\delta g_{1}+\delta g_{2}),$ where $\delta g_{1}$ and $\delta g_{2}$ solve (A1)–(A2).

Let us consider the local, linearised, electrostatic versions of (A1)–(A2):

(A3)\begin{gather} \frac{\partial\delta g_{1}}{\partial t}+{\rm i}k_{\psi}\bar{v}_{\psi}\delta g_{1}+v_{{\parallel}}\boldsymbol{\nabla}_{{\parallel}}\delta g_{1}=\frac{eF_{0}}{T} {\rm e}^{{\rm i}Q}\frac{\partial}{\partial t}\left\langle {\rm J}_{0}\phi\right\rangle _{\psi}, \end{gather}
(A4)\begin{gather} \frac{\partial\delta g_{2}}{\partial t}+{\rm i}k_{\psi}\bar{v}_{\psi}\delta g_{2}+v_{{\parallel}}\boldsymbol{\nabla}_{{\parallel}}\delta g_{2}=\frac{eF_{0}}{T} {\rm e}^{{\rm i}Q}\left\{ \frac{\partial}{\partial t}\left({\rm J}_{0}\phi-\left\langle {\rm J}_{0}\phi\right\rangle _{\psi}\right)-\frac{c}{B}\boldsymbol{b}\boldsymbol{\cdot}\boldsymbol{\nabla} {\rm J}_{0}\phi\times\boldsymbol{\nabla} F_{0}\right\}, \end{gather}

where we are setting $\partial _{\psi }\equiv {\rm i}k_{\psi },$ so $Q=k_{\psi }H,$ and we are using a hybrid notation for the diamagnetic term, in the sense that the explicit ${\rm J}_0$ is meant to remind the reader that the gyroaverage operation is taken at constant particle position. The electrostatic potential, $\phi,$ will be determined with a $k_{\perp }^{2}\rho _{i}^{2}\sim Q^{2}\ll 1$ expansion of the quasi-neutrality condition:

(A5)\begin{equation} -\frac{e}{T_{i}}\phi+\frac{1}{n}\int \,{\rm d}^{3}\boldsymbol{v}{\rm J}_{0}\delta G_{i}=\frac{e}{T_{e}}\phi+\frac{1}{n}\int \,{\rm d}^{3}\boldsymbol{v}\delta G_{e}, \end{equation}

where $\delta G_{s}$ are found for $\partial /\partial t\ll k_{\parallel }v_{thi}$ (Rosenbluth & Hinton Reference Rosenbluth and Hinton1998; Xiao & Catto Reference Xiao and Catto2006).

We anticipate an electrostatic potential which is a flux function to leading order in $k_\perp \rho _i\sim Q_i^2\ll 1,$ and introduce the Laplace transform of (A3)–(A4):

(A6)\begin{align} \left(p+{\rm i}k_{\psi}\bar{v}_{\psi}\right)\delta\hat{g}_{1}+v_{{\parallel}}\boldsymbol{\nabla}_{{\parallel}}\delta\hat{g}_{1}=\frac{eF_{0}}{T} {\rm e}^{{\rm i}Q}p\left\langle {\rm J}_{0}\hat{\phi}\right\rangle _{\psi} +\delta\hat{g}_{1}(0), \end{align}
(A7)\begin{align} \left(p+k_{\psi}\bar{v}_{\psi}\right)\delta g_{2}+v_{{\parallel}}\boldsymbol{\nabla}_{{\parallel}}\delta g_{2}=\frac{eF_{0e}}{T_{0e}} {\rm e}^{{\rm i}Q}\partial\left( {\rm J}_{0}\hat{\phi}-\left\langle {\rm J}_{0}\hat{\phi}\right\rangle \right). \end{align}

We then expand for $p/(k_{\parallel }v_{thi})\ll 1$ and obtain, to leading order (Rosenbluth & Hinton Reference Rosenbluth and Hinton1998; Xiao & Catto Reference Xiao and Catto2006),

(A8)\begin{equation} v_{{\parallel}}\boldsymbol{\nabla}_{{\parallel}}\delta\hat{g}_{j}^{(0)}=0, \end{equation}

which implies $\delta \hat {g}_{j}^{(0)}=\delta \hat {g}_{j}^{(0)}(p,\psi,\mu,\mathcal {E}),$ for $j=1,2.$ The actual form of $\delta \hat {g}_{j}^{(0)}$ is determined by bounce-/transit-averaging (A6)–(A7):

(A9)\begin{align} \delta\hat{g}_{1}^{(0)}=\frac{eF_{0}}{T}\frac{p}{p+{\rm i}k_{\psi}\bar{v}_{\psi}}\overline{{\rm e}^{{\rm i}Q}\left\langle {\rm J}_{0}\hat{\phi}\right\rangle _{\psi}}+\frac{\delta\hat{g}_{1}(0)}{p+{\rm i}k_{\psi}\bar{v}_{\psi}}, \end{align}
(A10)\begin{align} \delta\hat{g}_{2}^{(0)}=\frac{eF_{0}}{T}\frac{p}{p+{\rm i}k_{\psi}\bar{v}_{\psi}} \overline{{\rm e}^{{\rm i}Q}\left({\rm J}_{0}\hat{\phi}-\left\langle {\rm J}_{0}\hat{\phi}\right\rangle _{\psi}\right)}. \end{align}

For electrons, one simply sets $Q=0,$ ${\rm J}_{0}=1,$ $e\rightarrow -e$ and $\bar {v}_{\psi }\rightarrow -\bar {v}_{\psi }.$

Let us consider the $k_{\perp }^{2}\rho _{j}^{2}\sim Q^{2}\ll 1$ limit in (A10), and insert the result in (A7), for passing electrons, to see that

(A11)\begin{equation} \delta g_{1,e}^{(0)}={-}\frac{eF_{0e}}{T_{0e}}\frac{p}{p+k_{\psi}\bar{v}_{\psi,e}}\left\langle \hat{\phi}\right\rangle _{\psi}. \end{equation}

Similarly, to leading order, $\delta g_{2,e}^{(0)}=0.$ Then, in the tokamak (or helical symmetric case), for $k_\psi \bar {v}_\psi \rightarrow 0,$ we obtain the modified form of quasi-neutrality (Dorland & Hammett Reference Dorland and Hammett1993):

(A12)\begin{equation} \left(\tau+1\right)\frac{e\hat{\phi}}{T_i}=\frac{1}{ n}\int \,{\rm d}^{3}\boldsymbol{v}\delta G_i+\frac{e}{T_e}\left\langle \hat{\phi}\right\rangle _{\psi}, \end{equation}

otherwise known as (usually) proper electron response (Hammett et al. Reference Hammett, Beer, Dorland, Cowley and Smith1993). We now replace in (A5) the ion solution written in terms of (A9)–(A10):

(A13)\begin{align} \left(\frac{1}{\tau}+1\right)\hat{\phi}& =\frac{1}{\tau}\int \,{\rm d}^{3}\boldsymbol{v}\frac{F_{0i}}{n}\frac{p}{p+{\rm i}k_{\psi}\bar{v}_{\psi}} {\rm e}^{-{\rm i}Q}{\rm J}_{0}\left(\overline{{\rm e}^{{\rm i}Q}{\rm J}_{0}\hat{\phi}}+\frac{\delta\hat{g}_{1}(0)}{p}\right)\nonumber\\ & \quad +\int \,{\rm d}^{3}\boldsymbol{v}\frac{F_{0e}}{n}\frac{p}{p+{\rm i}k_{\psi}\bar{v}_{\psi,e}}\left\langle \hat{\phi}\right\rangle _{\psi}. \end{align}

Hence, up to second order,

(A14)\begin{align} \left(\frac{1}{\tau}+1\right)\hat{\phi}& =\frac{1}{\tau}\int \,{\rm d}^{3}\boldsymbol{v}\frac{F_{0i}}{n}\frac{p}{p+{\rm i}k_{\psi}\bar{v}_{\psi}}\left(1-{\rm i}Q-\frac{Q^{2}}{2}\right)\left(1-\frac{k_{{\perp}}^{2}\rho_{i}^{2}\hat{v}_{{\perp}}^{2}}{4}\right)\nonumber\\ & \quad \times \left[\left(1-{\rm i}\bar{Q}-\frac{\overline{Q^{2}}}{2}\right)\left(1-\frac{\overline{k_{{\perp}}^{2}\rho_{i}^{2}\hat{v}_{{\perp}}^{2}}}{4}\right)\left\langle \hat{\phi}\right\rangle _{\psi}+\frac{\delta\hat{g}_{1}(0)}{p}\right]\nonumber\\ & \quad +\int \,{\rm d}^{3}\boldsymbol{v}\frac{F_{0e}}{n}\frac{p}{p+{\rm i}k_{\psi}\bar{v}_{\psi,e}}\left\langle \hat{\phi}\right\rangle _{\psi}, \end{align}

where we now point out that $\delta \hat {g}_{1}(0)=\langle k_{\perp }^{2}\rho _{i}^{2}\rangle _{\psi }e\hat {\phi }(0)/T_{i}\sim k_{\perp }^{2}\rho _{i}^{2}e\phi /T_{i}.$ Notice that the velocity-space integrands of the tokamak result of Rosenbluth and Hinton are now simply multiplied by the factor $p/(p+{\rm i}k_{\psi }\bar {v}_{\psi }),$ and $p/(p+{\rm i}k_{\psi }\bar {v}_{\psi })\rightarrow 1,$ for $\bar {v}_{\psi }\rightarrow 0.$ Taking the flux surface average, we are left with

(A15)\begin{align} & \left\{ \left[\left(\tau+1\right)-\sum_{s}\frac{T_{i}}{T_{s}}\left\langle \int \,{\rm d}^{3}\boldsymbol{v}\frac{F_{0s}}{n}\frac{p}{p+{\rm i}k_{\psi}\bar{v}_{\psi,s}}\right\rangle _{\psi}\right]+\left\langle \frac{1}{4}\rho_{i}^{2}k_{{\perp}}^{2}\int \,{\rm d}^{3}\boldsymbol{v}\frac{F_{0i}}{n}\frac{p\left(\hat{v}_{{\perp}}^{2}+\overline{\hat{v}_{{\perp}}^{2}}\right)}{p+{\rm i}k_{\psi}\bar{v}_{\psi}}\right\rangle _{\psi}\right.\nonumber\\ & \qquad \left.+\left\langle \frac{1}{2}\int \,{\rm d}^{3}\boldsymbol{v}\frac{F_{0i}}{n}\frac{p}{p+{\rm i}k_{\psi}\bar{v}_{\psi}}\left(Q^{2}+\overline{Q^{2}}-2Q\bar{Q}\right)\right\rangle _{\psi}\right\} \left\langle \hat{\phi}\right\rangle _{\psi}\nonumber\\ & \qquad \quad = \left\langle k_{{\perp}}^{2}\rho_{i}^{2}\right\rangle _{\psi}\hat{\phi}(0)\left\langle \int \,{\rm d}^{3}\boldsymbol{v}\frac{F_{0i}}{n}\frac{1}{p+{\rm i}k_{\psi}\bar{v}_{\psi}}\right\rangle _{\psi}. \end{align}

Thus, for a vanishing bounce-averaged radial drift (tokamak case), we find the celebrated result of Rosenbluth and Hinton:

(A16)\begin{equation} \left\{ \frac{1}{2}\left\langle k_{{\perp}}^{2}\rho_{i}^{2}\right\rangle _{\psi}+\left\langle \int \,{\rm d}^{3}\boldsymbol{v}\frac{F_{0i}}{n}Q\left(Q-\bar{Q}\right)\right\rangle _{\psi}\right\} p\left\langle \hat{\phi}\right\rangle _{\psi}=\left\langle k_{{\perp}}^{2}\rho_{i}^{2}\right\rangle \hat{\phi}(0), \end{equation}

where we notice the exact cancellation of the adiabatic terms on the left-hand side of (A15). In a device with a finite bounce-averaged radial drift, the particles are not allowed to fully experience a Boltzmann response (nor non-adiabatic), since their motion along the field line, necessary to achieve an isothermal state, is interrupted by the radial particle excursion. This physical effect mathematically manifests in the lack of cancellation of the first term on the left-hand side of (A15), when $\bar {v}_{\psi }\neq 0.$ This is the fundamental reason why, at long wavelengths, zonal flows in stellarators are smaller than in tokamak geometry, since now

(A17)\begin{equation} \left\langle \hat{\phi}\right\rangle _{\infty}=\lim_{t\rightarrow\infty}\left\langle \hat{\phi}(t)\right\rangle \approx\left\langle k_{{\perp}}^{2}\rho_{i}^{2}\right\rangle \hat{\phi}(0), \end{equation}

for $\bar {v}_{\psi }\neq 0.$ Equation (A15) can also be obtained by calculating the neoclassical radial particle flux and using a long-wavelength approximation of Poisson's equation for the guiding centre density (Mishchenko et al. Reference Mishchenko, Helander and Könies2008):

(A18)\begin{equation} \left\langle k_{\psi}^{2}\rho_{i}^{2}\right\rangle _{\psi}\phi=\sum_{s}e_{s}N_{s}\equiv\sum_{s}e_{s}\int \,{\rm d}^{3}\boldsymbol{v}\delta G_{s}. \end{equation}

Obviously, in this case, the second term in (A16) gets replaced in the following way:

(A19)\begin{equation} \left\langle \frac{1}{4}\rho_{i}^{2}k_{{\perp}}^{2}\int \,{\rm d}^{3}\boldsymbol{v}\frac{F_{0i}}{n}\frac{p\left(\hat{v}_{{\perp}}^{2}+\overline{\hat{v}_{{\perp}}^{2}}\right)}{p+{\rm i}k_{\psi}\bar{v}_{\psi}}\right\rangle _{\psi}\rightarrow\frac{1}{2}\left\langle k_{{\perp}}^{2}\rho_{i}^{2}\right\rangle _{\psi}, \end{equation}

since Larmor radius corrections, in this case, are not evaluated by solving a gyrokinetic equation that retains the $p+{\rm i}k_{\psi }\bar {v}_{\psi }$ term.

Appendix B. Second-order momentum-space propagators

For wave-like perturbations, if we consider $\varPhi _{k}(t)=\phi ^{\nu }\exp ({\rm i}\omega _{\nu }t),$ then

(B1)\begin{equation} \hat{\varPhi}_{k}^{l}\equiv\frac{\zeta_{k,\nu}^{l}}{p+{\rm i}\omega_{\nu}}, \end{equation}

where $\zeta _{k,\nu }^{l}$ are now constants. After inserting this in the second-order kinetic solution (4.26), recalling that

(B2)\begin{equation} \tilde{h}_{k}^{(0)l}=\left(p+{\rm i}l\omega_{B}\right)h_{k}^{(0)l}, \end{equation}

with

(B3)\begin{equation} \left(p+{\rm i}l\omega_{B}\right)h_{k}^{(0)l}={-}\frac{e}{m}\frac{\partial F_{0}}{\partial\mathcal{E}}q_{l^{\prime}}\frac{\partial\varPhi_{k}^{l-l^{\prime}}}{\partial t}+\frac{\partial F_{0}}{\partial\mathcal{\psi}}H_{l,l^{\prime}}\varPhi_{k}^{l^{\prime}}, \end{equation}

we need to evaluate the integrals

(B4)\begin{gather} \mathcal{I}_{1}=\frac{1}{p+{\rm i}l\omega_{b}}\int \,{\rm d}p_{1}\frac{\zeta_{k_{1},\nu}^{l_{1}}}{p_{1}+{\rm i}\omega_{\nu}}\frac{\tilde{h}_{k-k_{1}}^{(0)l_{2}}(p-p_{1})}{p-p_{1}+{\rm i}l_{2}\omega_{b}}, \end{gather}
(B5)\begin{gather} \mathcal{I}_{2}=\frac{1}{p+{\rm i}l\omega_{b}}\int \,{\rm d}p_{1}\frac{\zeta_{k_{1},\nu}^{l_{1}}}{p_{1}+{\rm i}\omega_{\nu}}\frac{-{\rm i}\left(k-k_{1}\right)\bar{v}_{\psi}}{\left(p-p_{1}+{\rm i}l_{2}\omega_{b}\right)^{2}}\tilde{h}_{k-k_{1}}^{(0)l_{2}}(p-p_{1}) \end{gather}

and

(B6)\begin{align} \mathcal{I}_{3}=\frac{1}{p+{\rm i}l\omega_{b}}\int \,{\rm d}p_{1}\frac{\zeta_{k_{1},\nu}^{l_{1}}}{p_{1}+{\rm i}\omega_{\nu}}\frac{\mathcal{N}_{k-k_{1},k_{2}}^{l_{2},l_{1},l_{3}}}{p-p_{1}+{\rm i}l_{2}\omega_{b}}\int \,{\rm d}p_{2}\frac{\zeta_{k_{2},\nu}^{l_{1}}}{p_{2}+{\rm i}\omega_{\nu}}\frac{\tilde{h}_{k-k_{1}-k_{2}}^{(0)l_{3}}(p-p_{1}-p_{2})}{p-p_{1}-p_{2}+{\rm i}l_{3}\omega_{b}}, \end{align}

which produce the singularities found in the text.

References

Abel, I.G. & Cowley, S.C. 2013 Multiscale gyrokinetics for rotating tokamak plasmas. II. Reduced models for electron dynamics. New J. Phys. 15 (2), 023041.CrossRefGoogle Scholar
Al'tshu’l, L.M. & Karpman, V.I. 1966 Theory of nonlinear oscillations in a collisionless plasmas. Sov. Phys. JETP 22, 361.Google Scholar
Aleynikova, K. & Zocco, A. 2017 Quantitative study of kinetic ballooning mode theory in simple geometry. Phys. Plasmas 24 (9), 092106.CrossRefGoogle Scholar
Antonsen, T.M. & Lane, B. 1980 Kinetic equations for low frequency instabilities in inhomogeneous plasmas. Phys. Fluids 23 (6), 12051214.CrossRefGoogle Scholar
Biancalani, A., Bottino, A., Siena, A.D., Gürcan, Ö., Hayward-Schneider, T., Jenko, F., Lauber, P., Mishchenko, A., Morel, P., Novikau, I., et 2021 Gyrokinetic investigation of Alfvén instabilities in the presence of turbulence. Plasma Phys. Control. Fusion 63 (6), 065009.CrossRefGoogle Scholar
Boozer, A.H. 1981 Plasma equilibrium with rational magnetic surfaces. Phys. Fluids 24 (11), 19992003.CrossRefGoogle Scholar
Bottino, A., Falessi, M.V., Hayward-Schneider, T., Biancalani, A., Briguglio, S., Hatzky, R., Lauber, P., Mishchenko, A., Poli, E., Rettino, B., et al. 2022 Time evolution and finite element representation of Phase Space Zonal Structures in ORB5. J. Phys. Conf. Ser. 2397 (1), 012019.CrossRefGoogle Scholar
Carlevaro, N., Meng, G., Montani, G., Zonca, F., Hayward-Schneider, T., Lauber, P., Lu, Z. & Wang, X. 2022 One dimensional reduced model for ITER relevant energetic particle transport. Plasma Phys. Control. Fusion 64 (3), 035010 .CrossRefGoogle Scholar
Chen, L. & Zonca, F. 2016 Physics of Alfvén waves and energetic particles in burning plasmas. Rev. Mod. Phys. 88, 015008.CrossRefGoogle Scholar
Chen, L. & Zonca, F. 2019 Self-consistent kinetic theory with nonlinear wave-particle resonances. Plasma Sci. Technol. 21 (12), 125101.CrossRefGoogle Scholar
Connor, J.W., Hastie, R.J. & Taylor, J.B. 1978 Shear, periodicity, and plasma ballooning modes. Phys. Rev. Lett. 40 (6), 396.CrossRefGoogle Scholar
Connor, J.W., Hastie, R.J. & Taylor, J.B. 1979 High mode number stability of an axisymmetric toroidal plasma. Proc. R. Soc. Lond A 365 (1720), 117.Google Scholar
Dewar, R.L. & Glasser, A.H. 1983 Ballooning mode spectrum in general toroidal systems. Phys. Fluids 26 (10), 30383052.CrossRefGoogle Scholar
Di Siena, A., Görler, T., Poli, E., Bañón Navarro, A., Biancalani, A., Bilato, R., Bonanomi, N., Novikau, I., Vannini, F., Jenko, F., et 2021 Nonlinear electromagnetic interplay between fast ions and ion-temperature-gradient plasma turbulence. J. Plasma Phys. 87 (2), 555870201.CrossRefGoogle Scholar
Diamond, P.H., Itoh, S.-I., Itoh, K. & Hahm, T.S. 2005 Zonal flows in plasma—a review. Plasma Phys. Control. Fusion 47 (5), R35.CrossRefGoogle Scholar
Dorland, W. & Hammett, G.W. 1993 Gyrofluid turbulence models with kinetic effects. Phys. Fluids B 5 (3), 812835.CrossRefGoogle Scholar
Falessi, M.V. & Zonca, F. 2019 Transport theory of phase space zonal structures. Phys. Plasmas 26 (2), 022305.CrossRefGoogle Scholar
Frieman, E.A. & Chen, L. 1982 Nonlinear gyrokinetic equations for low‐frequency electromagnetic waves in general plasma equilibria. Phys. Fluids 25 (3), 502.CrossRefGoogle Scholar
Frieman, E.A., Rewoldt, G., Tang, W.M. & Glasser, A.H. 1980 General theory of kinetic ballooning modes. Phys. Fluids 23 (9), 17501769.CrossRefGoogle Scholar
Hammett, G.W., Beer, M.A., Dorland, W., Cowley, S.C. & Smith, S.A. 1993 Developments in the gyrofluid approach to tokamak turbulence simulations. Plasma Phys. Control. Fusion 35 (8), 973.CrossRefGoogle Scholar
Hasegawa, A. & Sato, T. 1981 Space Plasma Physics Stationary Processes, vol. 1. Springer.Google Scholar
Hayward-Schneider, T., Lauber, P., Bottino, A. & Mishchenko, A. 2022 Multi-scale analysis of global electromagnetic instabilities in iter pre-fusion-power operation plasmas. Nucl. Fusion 62 (11), 112007.CrossRefGoogle Scholar
Lu, Z., Meng, G., Hatzky, R., Hoelzl, M. & Lauber, P. 2023 Full f and gyrokinetic particle simulations of alfvén waves and energetic particle physics. Plasma Phys. Control. Fusion 65 (3), 034004.CrossRefGoogle Scholar
Mishchenko, A., Helander, P. & Könies, A. 2008 Collisionless dynamics of zonal flows in stellarator geometry. Phys. Plasmas 15 (7), 072309.CrossRefGoogle Scholar
Mishchenko, A. & Zocco, A. 2012 Global gyrokinetic particle-in-cell simulations of internal kink instabilities. Phys. Plasmas 19 (12), 122104.CrossRefGoogle Scholar
Peskin, M.E. & Schroeder, D.V. 1995 An Introduction to Quantum Field Theory. Perseus Books.Google Scholar
Qin, H., Tang, W.M. & Rewoldt, G. 1998 Gyrokinetic theory for arbitrary wavelength electromagnetic modes in tokamaks. Phys. Plasmas 5 (4), 10351049.CrossRefGoogle Scholar
Rogers, B.N., Dorland, W. & Kotschenreuther, M. 2000 Generation and stability of zonal flows in ion-temperature-gradient mode turbulence. Phys. Rev. Lett. 85, 53365339.CrossRefGoogle ScholarPubMed
Romanelli, F. & Zonca, F. 1993 The radial structure of the ion–temperature–gradient–driven mode. Phys. Fluids B 5 (11), 40814089.CrossRefGoogle Scholar
Rosenbluth, M.N. & Hinton, F.L. 1998 Poloidal flow driven by ion-temperature-gradient turbulence in tokamaks. Phys. Rev. Lett. 80, 724727.CrossRefGoogle Scholar
Sugama, H. & Watanabe, T.-H. 2005 Dynamics of zonal flows in helical systems. Phys. Rev. Lett. 94, 115001.CrossRefGoogle ScholarPubMed
Tang, W.M., Connor, J.W. & Hastie, R.J. 1980 Kinetic-ballooning-mode theory in general geometry. Nucl. Fusion 20 (11), 1439.CrossRefGoogle Scholar
Weinberg, S. 1962 Eikonal method in magnetohydrodynamics. Phys. Rev. 126 (6), 18991909.CrossRefGoogle Scholar
Xiao, Y. & Catto, P.J. 2006 Short wavelength effects on the collisionless neoclassical polarization and residual zonal flow level. Phys. Plasmas 13 (10), 102311.CrossRefGoogle Scholar
Zocco, A., Aleynikova, K. & Xanthopoulos, P. 2018 a Strongly driven surface-global kinetic ballooning modes in general toroidal geometry. J. Plasma Phys. 84 (3), 745840303.CrossRefGoogle Scholar
Zocco, A., Helander, P. & Connor, J.W. 2015 Magnetic compressibility and ion-temperature- gradient-driven microinstabilities in magnetically confined plasmas. Plasma Phys. Control. Fusion 57 (8), 085003.CrossRefGoogle Scholar
Zocco, A., Plunk, G.G. & Xanthopoulos, P. 2020 Geometric stabilization of the electrostatic ion-temperature-gradient driven instability. II. Non-axisymmetric systems. Phys. Plasmas 27 (2), 022507.CrossRefGoogle Scholar
Zocco, A., Plunk, G.G., Xanthopoulos, P. & Helander, P. 2016 Geometric stabilization of the electrostatic ion-temperature-gradient driven instability. I. Nearly axisymmetric systems. Phys. Plasmas 23 (8), 082516.CrossRefGoogle Scholar
Zocco, A., Xanthopoulos, P., Doerk, H., Connor, J.W. & Helander, P. 2018 b Threshold for the destabilisation of the ion-temperature-gradient mode in magnetically confined toroidal plasmas. J. Plasma Phys. 84 (1), 715840101.CrossRefGoogle Scholar
Zonca, F. & Chen, L. 1992 Resonant damping of toroidicity-induced shear-alfvén eigenmodes in tokamaks. Phys. Rev. Lett. 68, 592595.CrossRefGoogle ScholarPubMed
Zonca, F. & Chen, L. 1993 Theory of continuum damping of toroidal alfvén eigenmodes in finite—$\beta$ tokamaks. Phys. Fluids B 5 (10), 36683690.CrossRefGoogle Scholar
Zonca, F. & Chen, L. 2014 Theory on excitations of drift Alfvén waves by energetic particles. I. Variational formulation. Phys. Plasmas 21 (7), 072120.CrossRefGoogle Scholar
Zonca, F., Chen, L., Briguglio, S., Fogaccia, G., Vlad, G. & Wang, X. 2015 Nonlinear dynamics of phase space zonal structures and energetic particle physics in fusion plasmas. New J. Phys. 17 (1), 013052.CrossRefGoogle Scholar
Zonca, F., Chen, L., Falessi, M.V. & Qiu, Z. 2021 a Nonlinear radial envelope evolution equations and energetic particle transport in tokamak plasmas. J. Phys.: Conf. Ser. 1785 (1), 012005.Google Scholar
Zonca, F., Tao, X. & Chen, L. 2021 b Nonlinear dynamics and phase space transport by chorus emission. Rev. Mod. Plasma Phys. 5, 8.CrossRefGoogle Scholar
Figure 0

Figure 1. First-order iterative solution of (4.13).

Figure 1

Figure 2. Second-order iterative solution of (4.13).

Figure 2

Figure 3. The three types of momentum-space second-order propagators for wave-like perturbations when $-{\rm i} l \omega _b =p_0=p_1$. Signs are taken into account by arrows, and quantities are conserved at the vertices. At this order, stellarator corrections and nonlinearity interact.