The Kinetic Ion-Temperature-Gradient-driven instability and its localisation

E. Rodríguez    A. Zocco Max Planck Institute for Plasma Physics, 17491 Greifswald, Germany
Abstract

We construct a description of Ion Temperature Gradient (ITG) driven localised linear modes which retains both wave-particle and magnetic drift resonant effects while capturing the field-line dependence of the electrostatic potential. We exploit the smallness of the magnetic drift and the strong localisation of the mode to resolve the problem with a polynomial-gaussian expansion in the field-following co-ordinate. A simple semi-analytical formula for the spectrum of the mode is shown to capture long wavelength Landau dam**, ion-scale Larmor radius stabilization, weakening of Larmor radius effects at short-wavelengths and magnetic-drift resonant stabilisation. These elements lead to linear spectra with multiple maxima as observed in gyrokinetic simulations in stellarators. Connections to the transition to extended eigenfunctions and those localized by less unfavourable curvature regions (hop** solutions) are also made. The model provides a clear qualitative framework with which to interpret numerically simulated ITG modes linear spectra with realistic geometries, despite its limitations for exact quantitative predictions.

1 Introduction

The ion-temperature-gradient (ITG) is one of the primary modes driving turbulence in magnetic confinement fusion devices (Rudakov & Sagdeev, 1961; Coppi et al., 1967; Terry et al., 1982). As such, much literature exists devoted to understanding this type of instability. The most basic understanding of this mode can be gained by studying the linear response of the system, as described by the linearised gyrokinetic equation (Connor et al., 1980; Romanelli, 1989) used to evaluate quasineutrality. Extending our understanding of the mode and its response to the geometry is particularly important in the context of stellarator physics.

Linear studies of the ITG driven mode, either analytical or numerical, are numerous in the literature, but they are all fundamentally traceable to the review of Kadomtsev & Pogutse (1970b). Analytical progress often requires some simplifying assumptions. These occur at two levels. First, by considering separately the various destabilising mechanisms in the problem. Second, by reducing the complexity of the magnetic field along field lines (e.g., curvature, flux compression or the magnetic field magnitude), often described as constant. Our knowledge of the ITG builds on consideration of special cases that respond to different approximations of the first kind (Zocco et al., 2018). When the bad curvature, together with the temperature gradient, drives the mode unstable, we have a toroidal ITG (Terry et al., 1982). When destabilisation does not require curvature, but is enabled by the propagation of density along the field lines, with a specific relative phase with respect to temperature (Cowley et al., 1991), we have a slab ITG Rudakov & Sagdeev (1961); Kadomtsev & Pogutse (1970b). While any simplified perspective helps shedding light on the physics of the ITG mode, the selective treatment of the physics can hinder their scope. A pressing case of this is the overlook of the mode localisation along the field line in the toroidal branch (Terry et al., 1982), where the presence of bad curvature is paramount. Magnetic fields generally exhibit alternating regions of good and bad curvature every connection length.

A simple theoretical description of the inhomogeneity along the magnetic field line, and the consequent behaviour of the ITG mode, is approachable through a fluid description (Horton Jr et al., 1981; Hahm & Tang, 1988; Romanelli, 1989; Plunk et al., 2014). Assuming particle parallel streaming to be small (slow ion transit time), a strong drive and small curvature drift, one can describe the ITG through a second order ordinary differential equation along the field line. Such description incorporates key physics ingredients to the problem, and crucially couples the behaviour of the instability to its structure along the field line. However, while localisation seems so important for a mathematical characterisation of the fluid branches (Wesson & Campbell, 2011; Zocco et al., 2016) and interpretation of numerical results, its analytical treatment becomes increasingly intricate when Landau dam** and the magnetic drift resonance are included. Without a consistent inclusion of these kinetic elements, the fluid description suffers from fundamental shortcomings in describing the behaviour of microinstabilities at small scales (either large perpendicular wavenumber or small parallel scale). This limitation can lead to wrong expectations in the mode response to changes in geometry. But the latter being of crucial importance, especially in the context of current activities such as stellarator optimisation, we are in need to integrate all elements together. While much progress has been made in this direction for the development of the theory of Geodesic Acoustic Modes (Zonca et al., 1996; Sugama & Watanabe, 2006b, a; Qiu et al., 2018), a manageable kinetic theory that retains resonant effects and includes geometric localisation is desirable. We address this problem in this work.

In this article we introduce a formal approach to the gyrokinetic equation to describe localised ITG modes and its structure in a simple geometry with good and bad curvature alongside kinetic effects, both resonant (due to particles parallel streaming and magnetic drift) and non-resonant (due to Larmor radius effects). In Section 2 we start from the linearised gyrokinetic equation and transform it into a linear system of equations, introducing the key ordering in the curvature drift and localisation of the mode. In the following section, Section 3, we focus on constructing the dispersion equation for a model simple Gaussian shaped mode, and detail its derivation and numerical solution. Section 4 then studies the physical elements of the model constructed, including the roles of Landau dam**, finite Larmor radius (FLR) stabilisation and the curvature drift resonance. Estimates for important features such as critical temperature gradients are also given. Section 5 then discusses the behaviour of other modes other than that with the simplest Gaussian structure, to end with some numerical gyrokinetic simulations. Although the model is too simple to be quantitatively correct, we show it serves as a blueprint to interpret the results from the simulations. We use this to make connections to the presence of de-localised (slab) and hop** modes.

2 Localised description of the gyrokinetic equation

2.1 Rewriting the gyrokinetic equation

The starting point of our model construction is the linear gyrokinetic (GK) equation, which we write as follows for the ions,

ivg+(ωω~d)g=qiTiF0iJ0(ωω~)ϕ.𝑖subscript𝑣parallel-tosubscript𝑔𝜔subscript~𝜔𝑑𝑔subscript𝑞𝑖subscript𝑇𝑖subscript𝐹0𝑖subscript𝐽0𝜔subscript~𝜔italic-ϕiv_{\parallel}\partial_{\ell}g+(\omega-\tilde{\omega}_{d})g=\frac{q_{i}}{T_{i}% }F_{0i}J_{0}(\omega-\tilde{\omega}_{\star})\phi.italic_i italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT italic_g + ( italic_ω - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) italic_g = divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ω - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ) italic_ϕ . (1)

The equation is to be understood as a first order ordinary differential equation in \ellroman_ℓ (Taylor, 1976), the distance along a field-line, for the non-adiabatic part of the distribution function, denoted by g𝑔gitalic_g, which is a perturbation with respect to the background ion Maxwellian F0isubscript𝐹0𝑖F_{0i}italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT. To write the equation in this one-dimensional form, the ballooning transform has been considered (Taylor, 1976; Connor et al., 1978; Tang et al., 1980; Antonsen Jr & Lane, 1980) so that 𝐤=kαα+kψψsubscript𝐤perpendicular-tosubscript𝑘𝛼𝛼subscript𝑘𝜓𝜓\mathbf{k}_{\perp}=k_{\alpha}\nabla\alpha+k_{\psi}\nabla\psibold_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∇ italic_α + italic_k start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ∇ italic_ψ. Here α𝛼\alphaitalic_α and ψ𝜓\psiitalic_ψ are defined to be the straight field line Clebsch variables (D’haeseleer et al., 2012), such that 𝐁=ψ×α𝐁𝜓𝛼\mathbf{B}=\nabla\psi\times\nabla\alphabold_B = ∇ italic_ψ × ∇ italic_α and the toroidal magnetic flux is 2πψ2𝜋𝜓2\pi\psi2 italic_π italic_ψ. The response of the system is coupled to the electrostatic potential ϕitalic-ϕ\phiitalic_ϕ, and is modulated by the geometry of the magnetic field. Elements of the geometry are present in ω~dsubscript~𝜔𝑑\tilde{\omega}_{d}over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT and J0subscript𝐽0J_{0}italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The former represents the ion magnetic particle drifts, which in the small β𝛽\betaitalic_β limit may be written as ω~d=ωd(x2+x2/2)subscript~𝜔𝑑subscript𝜔𝑑superscriptsubscript𝑥parallel-to2superscriptsubscript𝑥perpendicular-to22\tilde{\omega}_{d}=\omega_{d}(x_{\parallel}^{2}+x_{\perp}^{2}/2)over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 ), where ωd=𝐯D𝐤=vTiρiκ×𝐁𝐤/Bsubscript𝜔𝑑subscript𝐯𝐷subscript𝐤perpendicular-tosubscript𝑣𝑇𝑖subscript𝜌𝑖𝜅𝐁subscript𝐤perpendicular-to𝐵\omega_{d}=\mathbf{v}_{D}\cdot\mathbf{k}_{\perp}=v_{Ti}\rho_{i}\kappa\times% \mathbf{B}\cdot\mathbf{k}_{\perp}/Bitalic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = bold_v start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT ⋅ bold_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_κ × bold_B ⋅ bold_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / italic_B, where x=v/vTisubscript𝑥parallel-tosubscript𝑣parallel-tosubscript𝑣𝑇𝑖x_{\parallel}=v_{\parallel}/v_{Ti}italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT and x=v/vTisubscript𝑥perpendicular-tosubscript𝑣perpendicular-tosubscript𝑣𝑇𝑖x_{\perp}=v_{\perp}/v_{Ti}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT are velocities normalised to the ion thermal speed vTi=2Ti/misubscript𝑣𝑇𝑖2subscript𝑇𝑖subscript𝑚𝑖v_{Ti}=\sqrt{2T_{i}/m_{i}}italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT = square-root start_ARG 2 italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG, and ρi=mivT/qiBsubscript𝜌𝑖subscript𝑚𝑖subscript𝑣𝑇subscript𝑞𝑖𝐵\rho_{i}=m_{i}v_{T}/q_{i}Bitalic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT / italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_B is the ion Larmor radius, where Tisubscript𝑇𝑖T_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, qisubscript𝑞𝑖q_{i}italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and misubscript𝑚𝑖m_{i}italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are the ion temperature, charge and mass. For simplicity, we will focus on the simpler limit kψ=0subscript𝑘𝜓0k_{\psi}=0italic_k start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT = 0. The second element of geometry is included in the finite Larmor radius term J0=J0(x2b)subscript𝐽0subscript𝐽0subscript𝑥perpendicular-to2𝑏J_{0}=J_{0}(x_{\perp}\sqrt{2b})italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT square-root start_ARG 2 italic_b end_ARG ), where J0subscript𝐽0J_{0}italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the Bessel function of the first kind, and b=(kρi)2/2𝑏superscriptsubscript𝑘perpendicular-tosubscript𝜌𝑖22b=(k_{\perp}\rho_{i})^{2}/2italic_b = ( italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2. The drive of the instability is included in the diamagnetic drift, represented here by ω~=ω[1+η(x23/2)]subscript~𝜔subscript𝜔delimited-[]1𝜂superscript𝑥232\tilde{\omega}_{\star}=\omega_{\star}[1+\eta(x^{2}-3/2)]over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT [ 1 + italic_η ( italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 3 / 2 ) ], where η=dlnTi/dlnni𝜂dsubscript𝑇𝑖dsubscript𝑛𝑖\eta=\mathrm{d}\ln T_{i}/\mathrm{d}\ln n_{i}italic_η = roman_d roman_ln italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / roman_d roman_ln italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the ratio of ion temperature to ion density gradients, and ω=(kαTi/qi)(dlnni/dψ)subscript𝜔subscript𝑘𝛼subscript𝑇𝑖subscript𝑞𝑖dsubscript𝑛𝑖d𝜓\omega_{\star}=(k_{\alpha}T_{i}/q_{i})(\mathrm{d}\ln n_{i}/\mathrm{d}\psi)italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = ( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ( roman_d roman_ln italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / roman_d italic_ψ ).

Because the GK equation is a first order ODE, a solution for g𝑔gitalic_g can be written in its most general form using an integrating factor, as originally presented by Connor et al. (1980). However, the resulting integral expressions, in their generality, do not always manifest clearly the role played by the different physical elements in the problem. After imposing quasineutrality, the relation between the mode structure and the degree of instability is not clear, as an involved integral eigenvalue problem ensues (Romanelli, 1989). To circumvent this complexity, we present here an approach that makes the resonant kinetic problem as close as possible to a second order ordinary differential equation, much in the way that it occurs in the fluid limit 111Some classes of integral eigenvalue equations can be easily related to linear differential problems (Tricomi, 1985), but this is not straightforward in our kinetic case..

To that end, we first invoke symmetry arguments to simplify the construction of the solution as much as possible. Given the explicit involvement of vsubscript𝑣parallel-tov_{\parallel}italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT in the GK equation, it is convenient to separate g𝑔gitalic_g into even and odd parts in vsubscript𝑣parallel-tov_{\parallel}italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT, namely,

ge(,𝐯)=12[g(v)+g(v)],superscript𝑔𝑒𝐯12delimited-[]𝑔subscript𝑣parallel-to𝑔subscript𝑣parallel-to\displaystyle g^{e}(\ell,\mathbf{v})=\frac{1}{2}\left[g(v_{\parallel})+g(-v_{% \parallel})\right],italic_g start_POSTSUPERSCRIPT italic_e end_POSTSUPERSCRIPT ( roman_ℓ , bold_v ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG [ italic_g ( italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) + italic_g ( - italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) ] , (2a)
go(,𝐯)=12[g(v)g(v)].superscript𝑔𝑜𝐯12delimited-[]𝑔subscript𝑣parallel-to𝑔subscript𝑣parallel-to\displaystyle g^{o}(\ell,\mathbf{v})=\frac{1}{2}\left[g(v_{\parallel})-g(-v_{% \parallel})\right].italic_g start_POSTSUPERSCRIPT italic_o end_POSTSUPERSCRIPT ( roman_ℓ , bold_v ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG [ italic_g ( italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) - italic_g ( - italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) ] . (2b)

With these well-defined parity functions, the original GK equation, which had mixed parity, can be separated into two coupled first order differential equations,

ivgo+(ωω~d)ge𝑖subscript𝑣parallel-tosubscriptsuperscript𝑔𝑜𝜔subscript~𝜔𝑑superscript𝑔𝑒\displaystyle iv_{\parallel}\partial_{\ell}g^{o}+(\omega-\tilde{\omega}_{d})g^% {e}italic_i italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT italic_g start_POSTSUPERSCRIPT italic_o end_POSTSUPERSCRIPT + ( italic_ω - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) italic_g start_POSTSUPERSCRIPT italic_e end_POSTSUPERSCRIPT =qiTiF0iJ0(ωω~)ϕ,absentsubscript𝑞𝑖subscript𝑇𝑖subscript𝐹0𝑖subscript𝐽0𝜔subscript~𝜔italic-ϕ\displaystyle=\frac{q_{i}}{T_{i}}F_{0i}J_{0}(\omega-\tilde{\omega}_{\star})\phi,= divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ω - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ) italic_ϕ , (3a)
ivge+(ωω~d)go𝑖subscript𝑣parallel-tosubscriptsuperscript𝑔𝑒𝜔subscript~𝜔𝑑superscript𝑔𝑜\displaystyle iv_{\parallel}\partial_{\ell}g^{e}+(\omega-\tilde{\omega}_{d})g^% {o}italic_i italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT italic_g start_POSTSUPERSCRIPT italic_e end_POSTSUPERSCRIPT + ( italic_ω - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) italic_g start_POSTSUPERSCRIPT italic_o end_POSTSUPERSCRIPT =0.absent0\displaystyle=0.= 0 . (3b)

The latter is used to eliminate go(,𝐯)superscript𝑔𝑜𝐯g^{o}(\ell,\mathbf{v})italic_g start_POSTSUPERSCRIPT italic_o end_POSTSUPERSCRIPT ( roman_ℓ , bold_v ) from the former, to give an equation that only involves the even part of g𝑔gitalic_g in vsubscript𝑣parallel-tov_{\parallel}italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT,222 A word of caution should be raised about the presence of the resonant denominator ωω~d𝜔subscript~𝜔𝑑\omega-\tilde{\omega}_{d}italic_ω - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT. The GK equation Eq. (1) can be understood as a Laplace transformed version of the its original time dependent form. It is thus well defined for {ω}>0𝜔0\Im\{\omega\}>0roman_ℑ { italic_ω } > 0 (the usual Bromwich contour), which avoids the ‘division by zero’. The extension to the remainder of ω𝜔\omegaitalic_ω-space is treated in Sec. 3.2. Retaining the resonance is important, especially as it can generate a finite imaginary part of a marginally stable fluid limit.

v[vωω~dge]+(ωω~d)ge=qiTiF0iJ0(ωω~)ϕ.subscript𝑣parallel-tosubscriptdelimited-[]subscript𝑣parallel-to𝜔subscript~𝜔𝑑subscriptsuperscript𝑔𝑒𝜔subscript~𝜔𝑑superscript𝑔𝑒subscript𝑞𝑖subscript𝑇𝑖subscript𝐹0𝑖subscript𝐽0𝜔subscript~𝜔italic-ϕv_{\parallel}\partial_{\ell}\left[\frac{v_{\parallel}}{\omega-\tilde{\omega}_{% d}}\partial_{\ell}g^{e}\right]+(\omega-\tilde{\omega}_{d})g^{e}=\frac{q_{i}}{T% _{i}}F_{0i}J_{0}(\omega-\tilde{\omega}_{\star})\phi.italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT [ divide start_ARG italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG ∂ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT italic_g start_POSTSUPERSCRIPT italic_e end_POSTSUPERSCRIPT ] + ( italic_ω - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) italic_g start_POSTSUPERSCRIPT italic_e end_POSTSUPERSCRIPT = divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ω - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ) italic_ϕ . (4)

We must not forget that the linearised GK equation, Eq. (1), does not come on its own. First, gesuperscript𝑔𝑒g^{e}italic_g start_POSTSUPERSCRIPT italic_e end_POSTSUPERSCRIPT must satisfy vanishing boundary conditions at ±plus-or-minus\ell\rightarrow\pm\inftyroman_ℓ → ± ∞ for a physically reasonable ballooning solution of the equation (Connor et al., 1978). Second, it must be complemented by the quasineutrality condition, the condition preventing charge separation from building up in the system. This imposes an additional relation between the velocity-space function gesuperscript𝑔𝑒g^{e}italic_g start_POSTSUPERSCRIPT italic_e end_POSTSUPERSCRIPT and the real space function ϕitalic-ϕ\phiitalic_ϕ, which is necessary to complete the ω𝜔\omegaitalic_ω-eigenvalue equation. As the velocity-space average of the odd gosuperscript𝑔𝑜g^{o}italic_g start_POSTSUPERSCRIPT italic_o end_POSTSUPERSCRIPT vanishes due to parity, the quasineutrality condition reduces to,

J0ged3𝐯=n¯(1+τ)ϕ,subscript𝐽0superscript𝑔𝑒superscriptd3𝐯¯𝑛1𝜏italic-ϕ\int J_{0}g^{e}\mathrm{d}^{3}\mathbf{v}=\bar{n}(1+\tau)\phi,∫ italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_g start_POSTSUPERSCRIPT italic_e end_POSTSUPERSCRIPT roman_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_v = over¯ start_ARG italic_n end_ARG ( 1 + italic_τ ) italic_ϕ , (5)

where τ=Ti/ZTe𝜏subscript𝑇𝑖𝑍subscript𝑇𝑒\tau=T_{i}/ZT_{e}italic_τ = italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_Z italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and n¯¯𝑛\bar{n}over¯ start_ARG italic_n end_ARG is the equilibrium density. We are taking the electron response to be adiabatic here. Equations (4) and (5), describing our ion response, will be our starting point.

2.2 Approximations: localisation and weak curvature

So far, the problem defined by Eqs. (4) and (5) is as general as the standard form of the linearised GK equation. To proceed, we introduce a number of simplifying assumptions that will make it tractable while retaining the main physical elements of the problem.

2.2.1 Simplified geometry

The first element of simplification in the problem is the geometry along the field line. We restrict all the inhomogeneity in the field to the \ellroman_ℓ dependence of the curvature drift ωd()subscript𝜔𝑑\omega_{d}(\ell)italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( roman_ℓ ). Doing so lets us capture the key aspect of having good and bad curvature regions along the field line. On physical grounds (Terry et al., 1982), we expect the unstable toroidal mode to have a tendency to localise around bad curvature regions, which are energetically favourable, while being repelled from good curvature ones.

To introduce a sense of this feature, we describe a single region of bad curvature, taken to be symmetric in \ellroman_ℓ and to have a finite extent, ΛΛ\Lambdaroman_Λ, beyond which the curvature changes sign. In the usual jargon, this length-scale may be deemed the connection length, and the region within can be thought as a bad curvature well. We model the well as a two-parameter simple symmetric quadratic function in \ellroman_ℓ, where the parameters represent its depth and width. The magnitude of the bad curvature at the origin is defined to be ω¯d<0subscript¯𝜔𝑑0-\bar{\omega}_{d}<0- over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT < 0. The curvature remains bad, i.e., negative, in the region ||<ΛΛ|\ell|<\Lambda| roman_ℓ | < roman_Λ. We then write ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT as

ωd=ω¯d[(Λ)21].subscript𝜔𝑑subscript¯𝜔𝑑delimited-[]superscriptΛ21\omega_{d}=\bar{\omega}_{d}\left[\left(\frac{\ell}{\Lambda}\right)^{2}-1\right].italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT [ ( divide start_ARG roman_ℓ end_ARG start_ARG roman_Λ end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ] . (6)

Because ΛΛ\Lambdaroman_Λ is the only existing length-scale along the field line, as both B𝐵Bitalic_B and kρisubscript𝑘perpendicular-tosubscript𝜌𝑖k_{\perp}\rho_{i}italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are assumed to be constant, we shall use it to normalise lengths and write ¯=/Λ¯Λ\bar{\ell}=\ell/\Lambdaover¯ start_ARG roman_ℓ end_ARG = roman_ℓ / roman_Λ.

Before moving on, we should reflect on the consequences of our simplified geometry. Choosing the magnetic field magnitude to be flat along the field line eliminates any contribution from trapped ions, which do not exist in our model. The physics associated to the variation of kρisubscript𝑘perpendicular-tosubscript𝜌𝑖k_{\perp}\rho_{i}italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are also lost, and with it the possible modulation of finite Larmor radius effects. In particular, this approximation erases the effects of global magnetic shear, which would have appeared as a secular term in ksubscript𝑘perpendicular-tok_{\perp}italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT. The global shear can become important especially in localising significantly extended, slab-like modes, and thus forcing the right behaviour of g𝑔gitalic_g at ±plus-or-minus\ell\rightarrow\pm\inftyroman_ℓ → ± ∞. Instead, in our problem, in the limit of large |¯|¯|\bar{\ell}|| over¯ start_ARG roman_ℓ end_ARG | we have a strongly positive good curvature. Physically, we expect this to play a stabilising role for the typical toroidal ITG that precludes modes from becoming completely delocalised. In that sense, our model of ωd()subscript𝜔𝑑\omega_{d}(\ell)italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( roman_ℓ ) mimics in part the action of global shear.333The unbounded behaviour of the curvature drift appears in this context as a rather artificial construct. However, note that in the gyrokinetic equation ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT has in fact a secular piece in \ellroman_ℓ. This piece is in fact proportional to the global magnetic shear, and thus there lies an additional connection between the ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT behaviour and the global shear. However, the secularity of ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT is not quadratic in \ellroman_ℓ, but rather linear, as follows from 𝐤kααιkαφψsimilar-tosubscript𝐤perpendicular-tosubscript𝑘𝛼𝛼similar-tosuperscript𝜄subscript𝑘𝛼𝜑𝜓\mathbf{k}_{\perp}\sim k_{\alpha}\nabla\alpha\sim-\iota^{\prime}k_{\alpha}% \varphi\nabla\psibold_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ∼ italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∇ italic_α ∼ - italic_ι start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_φ ∇ italic_ψ. Unavoidably, this simplification couples the local and global behaviour of the system. All the simplifications considered will render less accurate a quantitative comparison of the analytical results to real fields, but will serve as an important qualitative and semi-quantitative tool. We prove numerically that this particular approximation gives excellent results for the real geometry of the quasisymmetric (Nührenberg & Zille, 1988; Boozer, 1983; Rodriguez et al., 2020) HSX stellarator (Anderson et al., 1995).

Reducing all the field inhomogeneity to ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT is a significant formal simplification, making the differential operator subscript\partial_{\ell}∂ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT commute with everything in Eq. (4) except: g𝑔gitalic_g (we shall drop the superscript e𝑒eitalic_e describing parity for simplicity), ϕitalic-ϕ\phiitalic_ϕ and ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT. The assumption of a constant B𝐵Bitalic_B is of particular importance here. The partial derivative respect to \ellroman_ℓ is meant to be taken kee** the velocity space variables μ=miv2/2B𝜇subscript𝑚𝑖superscriptsubscript𝑣perpendicular-to22𝐵\mu=m_{i}v_{\perp}^{2}/2Bitalic_μ = italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_B and =miv2/2subscript𝑚𝑖superscript𝑣22\mathcal{E}=m_{i}v^{2}/2caligraphic_E = italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 constant. Only when B𝐵Bitalic_B is constant is vsubscript𝑣parallel-tov_{\parallel}italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT independent of real space.

With this observation, Eq. (4) becomes, upon commutation of the differential operator

(1ω~d/ω)1v2ω22g+(1ω~dω)g+v2ω2(ω~d/ω)(1ω~d/ω)2g=qiTiF0iJ0(1ω~ω)ϕ.superscript1subscript~𝜔𝑑𝜔1superscriptsubscript𝑣parallel-to2superscript𝜔2superscriptsubscript2𝑔1subscript~𝜔𝑑𝜔𝑔superscriptsubscript𝑣parallel-to2superscript𝜔2subscriptsubscript~𝜔𝑑𝜔superscript1subscript~𝜔𝑑𝜔2subscript𝑔subscript𝑞𝑖subscript𝑇𝑖subscript𝐹0𝑖subscript𝐽01subscript~𝜔𝜔italic-ϕ\left(1-\tilde{\omega}_{d}/\omega\right)^{-1}\frac{v_{\parallel}^{2}}{\omega^{% 2}}\partial_{\ell}^{2}g+\left(1-\frac{\tilde{\omega}_{d}}{\omega}\right)g+% \frac{v_{\parallel}^{2}}{\omega^{2}}\frac{\partial_{\ell}\left(\tilde{\omega}_% {d}/\omega\right)}{\left(1-\tilde{\omega}_{d}/\omega\right)^{2}}\partial_{\ell% }g=\frac{q_{i}}{T_{i}}F_{0i}J_{0}\left(1-\frac{\tilde{\omega}_{\star}}{\omega}% \right)\phi.( 1 - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT divide start_ARG italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∂ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_g + ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) italic_g + divide start_ARG italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG ∂ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ( over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω ) end_ARG start_ARG ( 1 - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∂ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT italic_g = divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) italic_ϕ . (7)

We have a second order ODE in which all the explicit spatial dependence can be expressed in terms of polynomials in \ellroman_ℓ.

2.2.2 Weak curvature drift, strong drive and strong localisation

It is clear from the above equation that the drift frequency in this scenario plays a central role in prescribing the behaviour of the instability. The mode will adapt to the geometry described by ωd(¯)subscript𝜔𝑑¯\omega_{d}(\bar{\ell})italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( over¯ start_ARG roman_ℓ end_ARG ), and thus to make the treatment more manageable, it is natural to consider ordering the drift. We introduce the ordering parameter δω¯d/ω1similar-to𝛿subscript¯𝜔𝑑𝜔much-less-than1\delta\sim\bar{\omega}_{d}/\omega\ll 1italic_δ ∼ over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω ≪ 1. Such a restriction is not completely artificial, and it is particularly appropriate in scenarios in which the turbulence is strongly driven, namely, in cases where the ωsubscript𝜔\omega_{\star}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT drive (either in its density or temperature gradient form) is much larger than ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT. We shall be focusing on this strongly driven scenario.

Note however that ordering ω¯d/ωsubscript¯𝜔𝑑𝜔\bar{\omega}_{d}/\omegaover¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω is not enough to simplify Eq. (7). Although ω¯d/ωsubscript¯𝜔𝑑𝜔\bar{\omega}_{d}/\omegaover¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω may be small, this is only its value at the bottom of the bad curvature well. Thus, there always exists a sufficiently large value of ¯¯\bar{\ell}over¯ start_ARG roman_ℓ end_ARG such that ωd/ω1greater-than-or-equivalent-tosubscript𝜔𝑑𝜔1\omega_{d}/\omega\gtrsim 1italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω ≳ 1. This appears hopeless for an approximated approach to Eq. (7). However, we must not consider ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT on its own when doing so, but rather alongside g𝑔gitalic_g and ϕitalic-ϕ\phiitalic_ϕ. If the distribution function and the potential are finite only over some finite length scale, then the effective value of ¯¯\bar{\ell}over¯ start_ARG roman_ℓ end_ARG should reflect that finite extent. Considering a Gaussian-like envelope of the form eλ¯2/2similar-toabsentsuperscript𝑒𝜆superscript¯22\sim e^{-\lambda\bar{\ell}^{2}/2}∼ italic_e start_POSTSUPERSCRIPT - italic_λ over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 end_POSTSUPERSCRIPT, so that g𝑔gitalic_g and ϕitalic-ϕ\phiitalic_ϕ have lengthscales 1/{λ}similar-toabsent1𝜆\sim 1/\sqrt{\Re\{\lambda\}}∼ 1 / square-root start_ARG roman_ℜ { italic_λ } end_ARG, where {λ}𝜆\Re\{\lambda\}roman_ℜ { italic_λ } denotes the real part of λ𝜆\lambdaitalic_λ which is generally complex, we limit this problem with a new ordering parameter,

ϵ=1{λ}ω¯dω1.italic-ϵ1𝜆subscript¯𝜔𝑑𝜔much-less-than1\epsilon=\frac{1}{\Re\{\lambda\}}\frac{\bar{\omega}_{d}}{\omega}\ll 1.italic_ϵ = divide start_ARG 1 end_ARG start_ARG roman_ℜ { italic_λ } end_ARG divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ≪ 1 . (8)

The parameter ϵitalic-ϵ\epsilonitalic_ϵ limits the construction that follows to modes that are sufficiently localised. Modes may have some degree of de-localisation, but limited by the ordering parameter δ𝛿\deltaitalic_δ. We thus shall be managing two simultaneous ordering assumptions, namely δ,ϵ1much-less-than𝛿italic-ϵ1\delta,~{}\epsilon\ll 1italic_δ , italic_ϵ ≪ 1. The precise implications of these orderings and how we proceed with them is detailed in what follows.

First, let us employ these newly introduced approximations to simplify Eq. (7). Expanding the equation to, and kee**, order O(ϵ,δ)𝑂italic-ϵ𝛿O(\epsilon,~{}\delta)italic_O ( italic_ϵ , italic_δ ),

2(ωtxω)2¯2g+(1ω~dω(¯21))g+4ω~d(ωtxω)2¯¯g=qiTiF0iJ0(1ω~ω)ϕ,2superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2superscriptsubscript¯2𝑔1subscript~𝜔𝑑𝜔superscript¯21𝑔4subscript~𝜔𝑑superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2¯subscript¯𝑔subscript𝑞𝑖subscript𝑇𝑖subscript𝐹0𝑖subscript𝐽01subscript~𝜔𝜔italic-ϕ2\left(\frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}\partial_{\bar{\ell}}^% {2}g+\left(1-\frac{\tilde{\omega}_{d}}{\omega}(\bar{\ell}^{2}-1)\right)g+4% \tilde{\omega}_{d}\left(\frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}\bar{% \ell}\partial_{\bar{\ell}}g=\frac{q_{i}}{T_{i}}F_{0i}J_{0}\left(1-\frac{\tilde% {\omega}_{\star}}{\omega}\right)\phi,2 ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∂ start_POSTSUBSCRIPT over¯ start_ARG roman_ℓ end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_g + ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ( over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) ) italic_g + 4 over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over¯ start_ARG roman_ℓ end_ARG ∂ start_POSTSUBSCRIPT over¯ start_ARG roman_ℓ end_ARG end_POSTSUBSCRIPT italic_g = divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) italic_ϕ , (9)

where allowing some looseness in the notation, ω~d=ω¯d(x2+x2/2)subscript~𝜔𝑑subscript¯𝜔𝑑superscriptsubscript𝑥parallel-to2superscriptsubscript𝑥perpendicular-to22\tilde{\omega}_{d}=\bar{\omega}_{d}(x_{\parallel}^{2}+x_{\perp}^{2}/2)over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 ) here, and ωtsubscript𝜔𝑡\omega_{t}italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is the transit frequency defined as,

ωt=vTiΛ2.subscript𝜔𝑡subscript𝑣𝑇𝑖Λ2\omega_{t}=\frac{v_{Ti}}{\Lambda\sqrt{2}}.italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = divide start_ARG italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT end_ARG start_ARG roman_Λ square-root start_ARG 2 end_ARG end_ARG . (10)

As of now, it is not clear what the consistent ordering of ωt/ωsubscript𝜔𝑡𝜔\omega_{t}/\omegaitalic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / italic_ω must be, but we may deal with that explicitly later. For now, we take it to be order 1, but consider ω¯d/ωsubscript¯𝜔𝑑𝜔\bar{\omega}_{d}/\omegaover¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω corrections to the first term in Eq. (9) small.

Localised solutions to a second order ODE like Eq. (9) may be approximated by considering a representation of g𝑔gitalic_g and ϕitalic-ϕ\phiitalic_ϕ in a Taylor-Gauss basis,

g=n=0gn¯nN(n)eλ¯2/2,𝑔superscriptsubscript𝑛0subscript𝑔𝑛superscript¯𝑛𝑁𝑛superscript𝑒𝜆superscript¯22g=\sum_{n=0}^{\infty}\frac{g_{n}\bar{\ell}^{n}}{N(n)}e^{-\lambda\bar{\ell}^{2}% /2},italic_g = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_N ( italic_n ) end_ARG italic_e start_POSTSUPERSCRIPT - italic_λ over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 end_POSTSUPERSCRIPT , (11)

where N(n)=n!/{λ}nN(n)=\sqrt{n!/\Re\{\lambda\}^{n}}italic_N ( italic_n ) = square-root start_ARG italic_n ! / roman_ℜ { italic_λ } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG is a normalisation factor. The Taylor part of the basis (i.e., the expansion in powers of ¯¯\bar{\ell}over¯ start_ARG roman_ℓ end_ARG) naturally describes the mode near the bottom of the well, while the Gaussian part provides an overall envelope that localises the mode. The latter requires {λ}>0𝜆0\Re\{\lambda\}>0roman_ℜ { italic_λ } > 0, although it is consistent with having a non-zero imaginary part.444It could be tempting to use Hermite polynomials instead of unpaired powers of ¯¯\bar{\ell}over¯ start_ARG roman_ℓ end_ARG as has been done to deal with velocity space in the literature (Mandell et al., 2018). However, the Taylor form is more convenient here with the consideration of the solution in the limit ¯0¯0\bar{\ell}\rightarrow 0over¯ start_ARG roman_ℓ end_ARG → 0 in mind. The normalisation factor includes powers of {λ}𝜆\Re\{\lambda\}roman_ℜ { italic_λ } to account for the scale associated to the monomial ¯nsuperscript¯𝑛\bar{\ell}^{n}over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT. Note that the higher the mode considered, the increasingly hollower the shape of the mode is, providing a characteristic length-scale n/{λ}n/2\sqrt{n}/\Re\{\lambda\}^{n/2}square-root start_ARG italic_n end_ARG / roman_ℜ { italic_λ } start_POSTSUPERSCRIPT italic_n / 2 end_POSTSUPERSCRIPT.

Once we have this basis representation, the ordinary differential equation Eq. (9) becomes an infinite set of coupled algebraic linear equation on {ϕn}subscriptitalic-ϕ𝑛\{\phi_{n}\}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } and {gn}subscript𝑔𝑛\{g_{n}\}{ italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT }. The benefit of the particular basis used is the simplicity of the form that the operations in Eq. (9) take. The product by ¯2superscript¯2\bar{\ell}^{2}over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT in the second term of Eq. (9) simply upshifts the mode number by 2, making the regularising role of ϵitalic-ϵ\epsilonitalic_ϵ manifest. It restricts the coupling of the different {gn}subscript𝑔𝑛\{g_{n}\}{ italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } and {ϕn}subscriptitalic-ϕ𝑛\{\phi_{n}\}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } modes, and thus is critical in achieving a useful truncation of the linear system of equations. Differentiation plays a similar, albeit more involved, coupling role (see Appendix A).

2.3 The Taylor-Gauss form of the gyrokinetic equation

We are in a position now to resolve Eq. (9) in the new basis of Eq. (11). This may be achieved by substituting the expansion in Eq. (11), and applying the coupling rules in Eqs. (46) and (47). The details of the derivation may be seen in Appendix A. Upon substitution, the resulting equation takes the form

n=0En{λ}nn!¯neλ¯2/2=0,\sum_{n=0}^{\infty}E_{n}\sqrt{\frac{\Re\{\lambda\}^{n}}{n!}}\bar{\ell}^{n}e^{-% \lambda\bar{\ell}^{2}/2}=0,∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT square-root start_ARG divide start_ARG roman_ℜ { italic_λ } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG end_ARG over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_λ over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 end_POSTSUPERSCRIPT = 0 , (12a)
where, the general expression for the n𝑛nitalic_n-th mode equation is,
En=2{λ}(ωtxω)2(n+2)(n+1)gn+2++[1+ω~dω2λ(2n+1)(ωtxω)2+4nω~dω(ωtxω)2]gn++2{λ}[λ2(ωtxω)2ω~d2ω2λω~dω(ωtxω)2]n(n1)gn2qiTiF0iJ0(1ω~ω)ϕn.subscript𝐸𝑛2𝜆superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2𝑛2𝑛1subscript𝑔𝑛2delimited-[]1subscript~𝜔𝑑𝜔2𝜆2𝑛1superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔24𝑛subscript~𝜔𝑑𝜔superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2subscript𝑔𝑛2𝜆delimited-[]superscript𝜆2superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2subscript~𝜔𝑑2𝜔2𝜆subscript~𝜔𝑑𝜔superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2𝑛𝑛1subscript𝑔𝑛2subscript𝑞𝑖subscript𝑇𝑖subscript𝐹0𝑖subscript𝐽01subscript~𝜔𝜔subscriptitalic-ϕ𝑛E_{n}=2\Re\{\lambda\}\left(\frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}% \sqrt{(n+2)(n+1)}g_{n+2}+\\ +\left[1+\frac{\tilde{\omega}_{d}}{\omega}-2\lambda(2n+1)\left(\frac{\omega_{t% }x_{\parallel}}{\omega}\right)^{2}+4n\frac{\tilde{\omega}_{d}}{\omega}\left(% \frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}\right]g_{n}+\\ +\frac{2}{\Re\{\lambda\}}\left[\lambda^{2}\left(\frac{\omega_{t}x_{\parallel}}% {\omega}\right)^{2}-\frac{\tilde{\omega}_{d}}{2\omega}-2\lambda\frac{\tilde{% \omega}_{d}}{\omega}\left(\frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}% \right]\sqrt{n(n-1)}g_{n-2}-\\ -\frac{q_{i}}{T_{i}}F_{0i}J_{0}\left(1-\frac{\tilde{\omega}_{\star}}{\omega}% \right)\phi_{n}.start_ROW start_CELL italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 2 roman_ℜ { italic_λ } ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT square-root start_ARG ( italic_n + 2 ) ( italic_n + 1 ) end_ARG italic_g start_POSTSUBSCRIPT italic_n + 2 end_POSTSUBSCRIPT + end_CELL end_ROW start_ROW start_CELL + [ 1 + divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG - 2 italic_λ ( 2 italic_n + 1 ) ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 italic_n divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + end_CELL end_ROW start_ROW start_CELL + divide start_ARG 2 end_ARG start_ARG roman_ℜ { italic_λ } end_ARG [ italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG - 2 italic_λ divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] square-root start_ARG italic_n ( italic_n - 1 ) end_ARG italic_g start_POSTSUBSCRIPT italic_n - 2 end_POSTSUBSCRIPT - end_CELL end_ROW start_ROW start_CELL - divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT . end_CELL end_ROW (12b)

For Eq. (9) to be true, Eq. (12a) must hold for all ¯¯\bar{\ell}over¯ start_ARG roman_ℓ end_ARG, meaning that each of the equations Ensubscript𝐸𝑛E_{n}italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT must vanish separately. The original ODE becomes this way an infinite dimensional system of linear equations, {En=0}subscript𝐸𝑛0\{E_{n}=0\}{ italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 0 }, as promised.

Inspection of the structure of this equation shows two important features of the system. First, that even and odd modes in ¯¯\bar{\ell}over¯ start_ARG roman_ℓ end_ARG are completely decoupled. This is, of course, a direct consequence of the original form of the equation having well defined parity. Unless otherwise stated, we shall consider the even set, of which its most basic form is a Gaussian mode. The results follow similarly for the odd set. The second noteworthy property is that, within each of these subspaces, the system has a tridiagonal structure. That is, the equation couples every mode to the immediately adjacent ones.

3 Gaussian mode model of kinetic ITG

The resolution of the original equation into the Taylor-Gauss basis leaves us with an infinite set of algebraic equations to solve in phase space (i.e., in 𝐯𝐯\mathbf{v}bold_v and \mathbf{\ell}roman_ℓ). It is our goal now to truncate this hierarchy down to obtain a useful model of the problem.

3.1 Constructing the dispersion function

Let us start first by understanding what the structure of the set of equations is when we truncate the system at order N𝑁Nitalic_N. That is, when we set all gn,ϕn=0subscript𝑔𝑛subscriptitalic-ϕ𝑛0g_{n},~{}\phi_{n}=0italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 0 for n>N𝑛𝑁n>Nitalic_n > italic_N. In that case the finite set of equations we are left with is,

{E0(0,2),,En(n2,n,n+2),,EN(N2,N),EN+2(N)},subscript𝐸002subscript𝐸𝑛𝑛2𝑛𝑛2subscript𝐸𝑁𝑁2𝑁subscript𝐸𝑁2𝑁\{E_{0}(0,2),\dots,E_{n}(n-2,n,n+2),\dots,E_{N}(N-2,N),E_{N+2}(N)\},{ italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 0 , 2 ) , … , italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_n - 2 , italic_n , italic_n + 2 ) , … , italic_E start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_N - 2 , italic_N ) , italic_E start_POSTSUBSCRIPT italic_N + 2 end_POSTSUBSCRIPT ( italic_N ) } ,

where each element must vanish and the numbers in parenthesis denote the modes (both in g𝑔gitalic_g and ϕitalic-ϕ\phiitalic_ϕ) involved in the equations. This system of equations may be rewritten by rearranging the first N/2+1𝑁21N/2+1italic_N / 2 + 1 equations to solve explicitly for {gn:nN}conditional-setsubscript𝑔𝑛𝑛𝑁\{g_{n}:~{}n\leq N\}{ italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT : italic_n ≤ italic_N } in terms of {ϕn:nN}conditional-setsubscriptitalic-ϕ𝑛𝑛𝑁\{\phi_{n}:~{}n\leq N\}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT : italic_n ≤ italic_N } by appropriate linear combinations. This is generally possible, and an explicit construction to order O(ϵ2)𝑂superscriptitalic-ϵ2O(\epsilon^{2})italic_O ( italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) is provided in Appendix B following Eq. (12b). We may write,

{gn(𝐯)=m=0N𝔻nm(N)(𝐯)ϕm(nN)gN(𝐯)=𝔻¯N(𝐯)ϕN,casessubscript𝑔𝑛𝐯absentsuperscriptsubscript𝑚0𝑁subscriptsuperscript𝔻𝑁𝑛𝑚𝐯subscriptitalic-ϕ𝑚𝑛𝑁subscript𝑔𝑁𝐯absentsubscript¯𝔻𝑁𝐯subscriptitalic-ϕ𝑁otherwise\begin{cases}\begin{aligned} g_{n}(\mathbf{v})=&\sum_{m=0}^{N}\mathbb{D}^{(N)}% _{nm}(\mathbf{v})\phi_{m}&(n\leq N)\\ g_{N}(\mathbf{v})=&\bar{\mathbb{D}}_{N}(\mathbf{v})\phi_{N},\end{aligned}\end{cases}{ start_ROW start_CELL start_ROW start_CELL italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_v ) = end_CELL start_CELL ∑ start_POSTSUBSCRIPT italic_m = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT blackboard_D start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ( bold_v ) italic_ϕ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_CELL start_CELL ( italic_n ≤ italic_N ) end_CELL end_ROW start_ROW start_CELL italic_g start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( bold_v ) = end_CELL start_CELL over¯ start_ARG blackboard_D end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( bold_v ) italic_ϕ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT , end_CELL end_ROW end_CELL start_CELL end_CELL end_ROW (13)

where 𝔻nmsubscript𝔻𝑛𝑚\mathbb{D}_{nm}blackboard_D start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT and 𝔻¯Nsubscript¯𝔻𝑁\bar{\mathbb{D}}_{N}over¯ start_ARG blackboard_D end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT are known functions of velocity space. The last isolated equation on gNsubscript𝑔𝑁g_{N}italic_g start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT, coming from EN+2subscript𝐸𝑁2E_{N+2}italic_E start_POSTSUBSCRIPT italic_N + 2 end_POSTSUBSCRIPT, is a result of the truncation at a finite N𝑁Nitalic_N, and leads to gNsubscript𝑔𝑁g_{N}italic_g start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT satisfying two equations simultaneously. The lack of degrees of freedom to salvage this overdetermination in velocity space is indicative of our failure to satisfy the original equation exactly with a finite number of modes. After all, we are attempting an approximate solution. We must therefore assess what is the error made by relaxing this last constraint. To do so, take M𝑀Mitalic_M to be the dominant mode in the system, which we shall take to be O(1)𝑂1O(1)italic_O ( 1 ). As shown in Appendix B explicitly, because the mode coupling terms are order ϵitalic-ϵ\epsilonitalic_ϵ, then we expect the magnitude of the error made by drop** that last equation to be ϵ(NM)/2similar-toabsentsuperscriptitalic-ϵ𝑁𝑀2\sim\epsilon^{(N-M)/2}∼ italic_ϵ start_POSTSUPERSCRIPT ( italic_N - italic_M ) / 2 end_POSTSUPERSCRIPT. The error made is thus small provided we restrict ourselves to MNNϵ=1/ϵmuch-less-than𝑀𝑁much-less-thansubscript𝑁italic-ϵ1italic-ϵM\ll N\ll N_{\epsilon}=1/\epsilonitalic_M ≪ italic_N ≪ italic_N start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT = 1 / italic_ϵ. The upper bound Nϵsubscript𝑁italic-ϵN_{\epsilon}italic_N start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT is set to preserve the presumed smallness of ϵitalic-ϵ\epsilonitalic_ϵ as it is amplified by the mode number of the solution to our system of equations. If N𝑁Nitalic_N is chosen to be sufficiently large, the matrix elements 𝔻nm(N)superscriptsubscript𝔻𝑛𝑚𝑁\mathbb{D}_{nm}^{(N)}blackboard_D start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_N ) end_POSTSUPERSCRIPT should then become largely independent of N𝑁Nitalic_N, and we may drop the (N)𝑁(N)( italic_N ) superscript. Having a model that is largely independent of the truncation is appropriate, and we shall see that a special choice of λ𝜆\lambdaitalic_λ enacts this truncation most snugly.

Our gyrokinetic problem has this way reduced to the main linear system of equations in Eq. (13). To complete the problem, we apply quasineutrality, Eq. (5). Taking the appropriate velocity moment of 𝔻nmsubscript𝔻𝑛𝑚\mathbb{D}_{nm}blackboard_D start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT, and defining,

𝒟nm=1n¯J0𝔻nm(𝐯)d3𝐯,subscript𝒟𝑛𝑚1¯𝑛subscript𝐽0subscript𝔻𝑛𝑚𝐯superscriptd3𝐯\mathcal{D}_{nm}=\frac{1}{\bar{n}}\int J_{0}\mathbb{D}_{nm}(\mathbf{v})\mathrm% {d}^{3}\mathbf{v},caligraphic_D start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG over¯ start_ARG italic_n end_ARG end_ARG ∫ italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT blackboard_D start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ( bold_v ) roman_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_v , (14)

the resulting eigenvalue problem becomes,

j=0N𝕄ijϕj=0,superscriptsubscript𝑗0𝑁subscript𝕄𝑖𝑗subscriptitalic-ϕ𝑗0\displaystyle\sum_{j=0}^{N}\mathbb{M}_{ij}\phi_{j}=0,∑ start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT blackboard_M start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 0 , (15a)
𝕄ij=(1+τ)δii𝒟ij.subscript𝕄𝑖𝑗1𝜏subscript𝛿𝑖𝑖subscript𝒟𝑖𝑗\displaystyle\mathbb{M}_{ij}=(1+\tau)\delta_{ii}-\mathcal{D}_{ij}.blackboard_M start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = ( 1 + italic_τ ) italic_δ start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT - caligraphic_D start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT . (15b)

Once we solve this system of equations, we may then construct gnsubscript𝑔𝑛g_{n}italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT for n<N𝑛𝑁n<Nitalic_n < italic_N, acknowledging the order ϵNMsuperscriptitalic-ϵ𝑁𝑀\epsilon^{N-M}italic_ϵ start_POSTSUPERSCRIPT italic_N - italic_M end_POSTSUPERSCRIPT error made as described above.

We now illustrate our way forward for the M=0𝑀0M=0italic_M = 0 mode (the approach for a general M𝑀Mitalic_M is given in Appendix B). This is to take the Gaussian centred about the bad curvature to be our dominant mode; formally, ϕ0O(1)similar-tosubscriptitalic-ϕ0𝑂1\phi_{0}\sim O(1)italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ italic_O ( 1 ). Following the principle of staying as distant as possible from the truncation point of the system, we may ask when the solution to the problem is consistent with ϕn=0subscriptitalic-ϕ𝑛0\phi_{n}=0italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 0 for all n>0𝑛0n>0italic_n > 0. We call this a pure mode, in so much as it is consistent with a ‘complete’ decoupling from higher mode numbers.

To order ϵitalic-ϵ\epsilonitalic_ϵ, the system describing this ‘pure’ n=0𝑛0n=0italic_n = 0 mode reduces to the following two equations,

𝒟20=0,subscript𝒟200\displaystyle\mathcal{D}_{20}=0,caligraphic_D start_POSTSUBSCRIPT 20 end_POSTSUBSCRIPT = 0 , (16a)
1+τ𝒟00=0,1𝜏subscript𝒟000\displaystyle 1+\tau-\mathcal{D}_{00}=0,1 + italic_τ - caligraphic_D start_POSTSUBSCRIPT 00 end_POSTSUBSCRIPT = 0 , (16b)

where the expressions for 𝒟00subscript𝒟00\mathcal{D}_{00}caligraphic_D start_POSTSUBSCRIPT 00 end_POSTSUBSCRIPT and 𝒟20subscript𝒟20\mathcal{D}_{20}caligraphic_D start_POSTSUBSCRIPT 20 end_POSTSUBSCRIPT to order ϵitalic-ϵ\epsilonitalic_ϵ are,

𝒟20subscript𝒟20\displaystyle\mathcal{D}_{20}caligraphic_D start_POSTSUBSCRIPT 20 end_POSTSUBSCRIPT =22{λ}qiTid3𝐯J02ev2/vTi2(1ω~ω)[λ2(ωtxω)2ω~d2ω],absent22𝜆subscript𝑞𝑖subscript𝑇𝑖superscriptd3𝐯superscriptsubscript𝐽02superscript𝑒superscript𝑣2superscriptsubscript𝑣𝑇𝑖21subscript~𝜔𝜔delimited-[]superscript𝜆2superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2subscript~𝜔𝑑2𝜔\displaystyle=\frac{2\sqrt{2}}{\Re\{\lambda\}}\frac{q_{i}}{T_{i}}\int\mathrm{d% }^{3}\mathbf{v}J_{0}^{2}e^{-v^{2}/v_{Ti}^{2}}\left(1-\frac{\tilde{\omega}_{% \star}}{\omega}\right)\left[\lambda^{2}\left(\frac{\omega_{t}x_{\parallel}}{% \omega}\right)^{2}-\frac{\tilde{\omega}_{d}}{2\omega}\right],= divide start_ARG 2 square-root start_ARG 2 end_ARG end_ARG start_ARG roman_ℜ { italic_λ } end_ARG divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ∫ roman_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_v italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) [ italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG ] , (17a)
𝒟00subscript𝒟00\displaystyle\mathcal{D}_{00}caligraphic_D start_POSTSUBSCRIPT 00 end_POSTSUBSCRIPT =1n¯qiTid3𝐯J02F0i(1ω~ω)11+ω~dω2λ(ωtxω)2.absent1¯𝑛subscript𝑞𝑖subscript𝑇𝑖superscriptd3𝐯superscriptsubscript𝐽02subscript𝐹0𝑖1subscript~𝜔𝜔11subscript~𝜔𝑑𝜔2𝜆superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2\displaystyle=\frac{1}{\bar{n}}\frac{q_{i}}{T_{i}}\int\mathrm{d}^{3}\mathbf{v}% J_{0}^{2}F_{0i}\left(1-\frac{\tilde{\omega}_{\star}}{\omega}\right)\frac{1}{1+% \frac{\tilde{\omega}_{d}}{\omega}-2\lambda\left(\frac{\omega_{t}x_{\parallel}}% {\omega}\right)^{2}}.= divide start_ARG 1 end_ARG start_ARG over¯ start_ARG italic_n end_ARG end_ARG divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ∫ roman_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_v italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) divide start_ARG 1 end_ARG start_ARG 1 + divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG - 2 italic_λ ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (17b)

Both of these expressions constitute the governing dispersion relation for our mode. It might appear that these make the problem overconstrained, however, we must bear in mind that ω𝜔\omegaitalic_ω is not the only unknown here. The localisation λ𝜆\lambdaitalic_λ is as well, and it must be chosen alongside the frequency of the mode. This is analogous to what happens in the fluid limit (Hahm & Tang, 1988; Hastie, 2013; Plunk et al., 2014; Zocco et al., 2016). In fact, a closed form for λ𝜆\lambdaitalic_λ may be obtained from satisfying Eq. (17a). For simplicity, considering the small b𝑏bitalic_b limit and the leading order form of the expression in δ,ϵ𝛿italic-ϵ\delta,~{}\epsilonitalic_δ , italic_ϵ,

𝒟202{λ}[1ωω(1+η)][λ2ω¯dωωt2]0,subscript𝒟202𝜆delimited-[]1subscript𝜔𝜔1𝜂delimited-[]superscript𝜆2subscript¯𝜔𝑑𝜔superscriptsubscript𝜔𝑡20\displaystyle\mathcal{D}_{20}\approx\frac{\sqrt{2}}{\Re\{\lambda\}}\left[1-% \frac{\omega_{\star}}{\omega}(1+\eta)\right]\left[\lambda^{2}-\frac{\bar{% \omega}_{d}\omega}{\omega_{t}^{2}}\right]\approx 0,caligraphic_D start_POSTSUBSCRIPT 20 end_POSTSUBSCRIPT ≈ divide start_ARG square-root start_ARG 2 end_ARG end_ARG start_ARG roman_ℜ { italic_λ } end_ARG [ 1 - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ( 1 + italic_η ) ] [ italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT italic_ω end_ARG start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] ≈ 0 ,
λ=ωω¯dωt2.thereforeabsent𝜆𝜔subscript¯𝜔𝑑superscriptsubscript𝜔𝑡2\displaystyle\therefore~{}\lambda=\sqrt{\frac{\omega\bar{\omega}_{d}}{\omega_{% t}^{2}}}.∴ italic_λ = square-root start_ARG divide start_ARG italic_ω over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG . (18)

For physically meaningful modes, and to restrict ourselves to {λ}>0𝜆0\Re\{\lambda\}>0roman_ℜ { italic_λ } > 0, we define the square root in Eq. (18) with its branch cut along the negative real axis in the complex plane (its conventional definition). The mode envelope λ𝜆\lambdaitalic_λ indicates a balance between the parallel streaming, ωt2superscriptsubscript𝜔𝑡2\omega_{t}^{2}italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, and the curvature drift ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, as is well known to be behind the localisation mechanism in the fluid limit of the ITG (Hahm & Tang, 1988; Plunk et al., 2014; Zocco et al., 2016). Upon approaching marginal stability, the mode will tend to become increasingly de-localised, with an oscillating mode structure when it co-rotates with ions.

The basic scaling of λ𝜆\lambdaitalic_λ together with the ordering of ϵitalic-ϵ\epsilonitalic_ϵ implies that, ξ=ωt/ω𝜉subscript𝜔𝑡𝜔\xi=\omega_{t}/\omegaitalic_ξ = italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / italic_ω follows δ1/2ξϵ1superscript𝛿12𝜉italic-ϵmuch-less-than1\delta^{1/2}\xi\leq\epsilon\ll 1italic_δ start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT italic_ξ ≤ italic_ϵ ≪ 1. It is convenient then to take for the ordering arguments ξδ1/2ϵsimilar-to𝜉superscript𝛿12italic-ϵ\xi\delta^{1/2}\sim\epsilonitalic_ξ italic_δ start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ∼ italic_ϵ. This is consistent with the assumptions made to reach this point. Although such consistency is reassuring, the precise form of λ𝜆\lambdaitalic_λ in Eq. (18) is a direct consequence of the assumption of a ‘pure’ mode. This purity assumption is however not whimsical, but particularly representative. Not only is it consistent with the ordering, but it also brings an elegant and efficient closure to our system of equations (see a discussion on this in Appendix B), as well as granting the correct asymptotic fluid limit of the dispersion equation, as we shall later discuss. The choice of λ𝜆\lambdaitalic_λ may thus be interpreted as a ‘fluid’ choice, with all the approximations that this involves. However, as one may check upon lifting the purity assumption, this remains a simplifying choice. In fact, such a relaxation eliminates in our case the 𝒟20=0subscript𝒟200\mathcal{D}_{20}=0caligraphic_D start_POSTSUBSCRIPT 20 end_POSTSUBSCRIPT = 0 constrain, leaving λ𝜆\lambdaitalic_λ undetermined.555A straightforward way of seeing this is by taking the determinant of 𝕄𝕄\mathbb{M}blackboard_M to vanish. To order ϵitalic-ϵ\epsilonitalic_ϵ this is equal to the product of the principal diagonal. The M=0𝑀0M=0italic_M = 0 mode thus simply requires vanishing of Eq. (16b). Future work may be devoted to improving the model by determining λ𝜆\lambdaitalic_λ in some other way, perhaps treating it as a free parameter to optimise to maximise the growth rate of the ITG. However, in this first work we willingly sacrifice a precise quantitative prediction for simplicity. We will illustrate this expected quantitative mismatch comparing the model to some example gyrokinetic simulations. We shall however see that the model remains a highly useful tool to interpret complicated linear turbulence spectra.

With this simple form for λ𝜆\lambdaitalic_λ as a function of ω𝜔\omegaitalic_ω, we then interpret Eq. (17b) as a dispersion relation for ω𝜔\omegaitalic_ω,

𝒟=1+τ+ζn¯J02(1ω~ω)f0x2ζ(1+ω¯d2ωx2)d3𝐯,𝒟1𝜏𝜁¯𝑛superscriptsubscript𝐽021subscript~𝜔𝜔subscript𝑓0superscriptsubscript𝑥parallel-to2𝜁1subscript¯𝜔𝑑2𝜔superscriptsubscript𝑥perpendicular-to2superscriptd3𝐯\mathcal{D}=1+\tau+\frac{\zeta}{\bar{n}}\int J_{0}^{2}\left(1-\frac{\tilde{% \omega}_{\star}}{\omega}\right)\frac{f_{0}}{x_{\parallel}^{2}-\zeta\left(1+% \frac{\bar{\omega}_{d}}{2\omega}x_{\perp}^{2}\right)}\mathrm{d}^{3}\mathbf{v},caligraphic_D = 1 + italic_τ + divide start_ARG italic_ζ end_ARG start_ARG over¯ start_ARG italic_n end_ARG end_ARG ∫ italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) divide start_ARG italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ζ ( 1 + divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG roman_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_v , (19)

where,

ζ=12[λ(ωtω)2ω¯d2ω]1.𝜁12superscriptdelimited-[]𝜆superscriptsubscript𝜔𝑡𝜔2subscript¯𝜔𝑑2𝜔1\zeta=\frac{1}{2}\left[\lambda\left(\frac{\omega_{t}}{\omega}\right)^{2}-\frac% {\bar{\omega}_{d}}{2\omega}\right]^{-1}.italic_ζ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG [ italic_λ ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT . (20a)
or using the form of λ𝜆\lambdaitalic_λ in Eq. (18),
ζ=ω22λ(1λ/2)ωt2.𝜁superscript𝜔22𝜆1𝜆2superscriptsubscript𝜔𝑡2\zeta=\frac{\omega^{2}}{2\lambda(1-\lambda/2)\omega_{t}^{2}}.italic_ζ = divide start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_λ ( 1 - italic_λ / 2 ) italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (20b)

The parameter ζ𝜁\zetaitalic_ζ can be interpreted as a measure of the kinetic effects, being important for |ζ|1.similar-to𝜁1|\zeta|\sim 1.| italic_ζ | ∼ 1 . Equation (20a) includes resonant kinetics that can come in either through a Landau-type resonance involving the structure along the field line (the λ𝜆\lambdaitalic_λ piece), or the magnetic drift resonance. The relative importance of these terms will depend on the scale of the mode structure along the field line, i.e., the magnitude of |λ|𝜆|\lambda|| italic_λ |

3.2 Simplifying the dispersion function

To evaluate the dispersion relation in Eq. (19) we must perform the necessary velocity space integrals, which includes resolving a resonant denominator. We shall deal with it by first performing the integral over xsubscript𝑥parallel-tox_{\parallel}italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT (and ignoring the resonance in xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT, as detailed in Appendix C). To that end, we need to resolve integrals of the following form

I,β(ζ)=1πx2βex2x2ζdx.I_{\parallel,\beta}(\zeta)=\frac{1}{\sqrt{\pi}}\int_{-\infty}^{\infty}x_{% \parallel}^{2\beta}\frac{e^{-x_{\parallel}^{2}}}{x_{\parallel}^{2}-\zeta}% \mathrm{d}x_{\parallel}.italic_I start_POSTSUBSCRIPT ∥ , italic_β end_POSTSUBSCRIPT ( italic_ζ ) = divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_π end_ARG end_ARG ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_β end_POSTSUPERSCRIPT divide start_ARG italic_e start_POSTSUPERSCRIPT - italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ζ end_ARG roman_d italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT . (21)

Using the difference of squares for the denominator, we rewrite the integral

I,β(ζ)=1ζ1πx2βex2xζdx,I_{\parallel,\beta}(\zeta)=\frac{1}{\sqrt[*]{\zeta}}\frac{1}{\sqrt{\pi}}\int_{% -\infty}^{\infty}\frac{x_{\parallel}^{2\beta}e^{-x_{\parallel}^{2}}}{x_{% \parallel}-\sqrt[*]{\zeta}}\mathrm{d}x_{\parallel},italic_I start_POSTSUBSCRIPT ∥ , italic_β end_POSTSUBSCRIPT ( italic_ζ ) = divide start_ARG 1 end_ARG start_ARG nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG end_ARG divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_π end_ARG end_ARG ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_β end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT - nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG end_ARG roman_d italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT , (22)

which will soon be related to the plasma dispersion function (Fried & Conte, 2015). Here, ζ𝜁\sqrt[*]{\zeta}nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG denotes a choice for a branch cut of the function f(ζ)=ζ𝑓𝜁𝜁f(\zeta)=\sqrt{\zeta}italic_f ( italic_ζ ) = square-root start_ARG italic_ζ end_ARG that maps the portion of the ω𝜔\omegaitalic_ω-plane {ω}𝜔\Im\{\omega\}\rightarrow\inftyroman_ℑ { italic_ω } → ∞ to {ζ}>0𝜁0\Im\{\sqrt[*]{\zeta}\}>0roman_ℑ { nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG } > 0. This choice is important for the problem to be consistent with the time-dependent description and the inverse Laplace transform. For large positive growth rates this avoids pole contributions to the Bromwich contour (and thus to the plasma dispersion function), making the inverse Laplace transform well defined. To evaluate the latter it is convenient to deform the Bromwich contour from its original position in the positive imaginary part of the plane downwards. Thus, in addition to the choice regarding the sign of {ζ}𝜁\Im\{\sqrt[*]{\zeta}\}roman_ℑ { nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG }, it is convenient to construct a Riemann sheet by placing branch-cuts southwards.

Refer to caption
Figure 1: Branch cuts for ζ𝜁\sqrt[*]{\zeta}nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG. Two different Riemann sheets are shown (a and b) for ζ𝜁\sqrt[*]{\zeta}nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG in complex ω𝜔\omegaitalic_ω-space. The plots on the left and right show the real and imaginary parts of ζ𝜁\sqrt[*]{\zeta}nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG, respectively. The plots in a) represent the natural Laplace continuation choice for the branch cuts, while b) is the choice that represents localised solutions everywhere in the ω𝜔\omegaitalic_ω plane ({λ}>0𝜆0\Re\{\lambda\}>0roman_ℜ { italic_λ } > 0). The branch cuts are denoted by the red wiggly lines across which the function is discontinuous. The function has an integrable singularity at ω=4ωt2/ωd𝜔4superscriptsubscript𝜔𝑡2subscript𝜔𝑑\omega=4\omega_{t}^{2}/\omega_{d}italic_ω = 4 italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT as indicated in the text. Frequency is normalised to ωtsubscript𝜔𝑡\omega_{t}italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and ωt/ωd=1/2subscript𝜔𝑡subscript𝜔𝑑12\omega_{t}/\omega_{d}=1/2italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 1 / 2 is chosen for illustration, with the colormaps normalised for appropriate visualisation.

To enforce the above, we choose two branch cuts in ω𝜔\omegaitalic_ω-space originating from the critical points λ=0𝜆0\lambda=0italic_λ = 0 and λ=1𝜆1\lambda=1italic_λ = 1. The latter is, in addition to a branch point, also a singular point, ζ1/1λ/2similar-to𝜁11𝜆2\sqrt[*]{\zeta}\sim 1/\sqrt{1-\lambda/2}nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG ∼ 1 / square-root start_ARG 1 - italic_λ / 2 end_ARG. The branch points represented will lead to some secular time dependence for damped modes (Kuroda et al., 1998; Sugama, 1999) which we shall not explore further in this work. With these branch points localised, the natural branch-cut choice in Figure 1a is problematic, as the branch cut emanating from ω=0𝜔0\omega=0italic_ω = 0 is directly linked to the definition of λ𝜆\lambdaitalic_λ, Eq. (18), and the vertical choice of the cut does not guarantee the physical requirement of localised {λ}>0𝜆0\Re\{\lambda\}>0roman_ℜ { italic_λ } > 0 modes. To avoid this, one must place the branch cut along the real line, as in Fig. 1b. Although this changes the continuation of the Bromwich contour to the negative {ω}<0𝜔0\Im\{\omega\}<0roman_ℑ { italic_ω } < 0 part of the plane, it should not affect the description of unstable modes, which is our main concern here.

With either definition in our hands, we proceed and write,

I,β(ζ)=(1)βζsβ[Z(sζ)]s=1,I_{\parallel,\beta}(\zeta)=\frac{(-1)^{\beta}}{\sqrt[*]{\zeta}}\partial_{s}^{% \beta}\left[Z\left(\sqrt{s}\sqrt[*]{\zeta}\right)\right]_{s=1},italic_I start_POSTSUBSCRIPT ∥ , italic_β end_POSTSUBSCRIPT ( italic_ζ ) = divide start_ARG ( - 1 ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT end_ARG start_ARG nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG end_ARG ∂ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT [ italic_Z ( square-root start_ARG italic_s end_ARG nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG ) ] start_POSTSUBSCRIPT italic_s = 1 end_POSTSUBSCRIPT , (23)

where Z()𝑍Z(\cdot)italic_Z ( ⋅ ) is the well known plasma dispersion function (Fried & Conte, 2015), analytical continuation of the integral

Z(a)=1πex2xadx,𝑍𝑎1𝜋superscriptsubscriptsuperscript𝑒superscript𝑥2𝑥𝑎differential-d𝑥Z(a)=\frac{1}{\sqrt{\pi}}\int_{-\infty}^{\infty}\frac{e^{-x^{2}}}{x-a}\mathrm{% d}x,italic_Z ( italic_a ) = divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_π end_ARG end_ARG ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_e start_POSTSUPERSCRIPT - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_x - italic_a end_ARG roman_d italic_x , (24)

beyond {a}>0𝑎0\Im\{a\}>0roman_ℑ { italic_a } > 0. The introduction of s𝑠sitalic_s as a dummy parameter in Eq. (23) follows from application of the Feynman trick (Woods, 1926, Ch. VI) to compute the integral in Eq. (22) in terms of Eq. (24).

We then have, after integration over xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT (see the details in Appendix C),

𝒟=1+τ+F0(b)[(1ωω+32ωTω)ζZ(ζ)ωTωζZ+]ωTωF2(b)ζZ(ζ)++ωω¯d4ω2{F2(b)[η(32+ζ)1+(1+2ζ+η(2ζ(ζ2)32))Z+]++F4(b)η[(1+2ζ)Z+1]}.𝒟1𝜏subscript𝐹0𝑏delimited-[]1subscript𝜔𝜔32superscriptsubscript𝜔𝑇𝜔𝜁𝑍𝜁superscriptsubscript𝜔𝑇𝜔𝜁subscript𝑍superscriptsubscript𝜔𝑇𝜔subscript𝐹2𝑏𝜁𝑍𝜁subscript𝜔subscript¯𝜔𝑑4superscript𝜔2subscript𝐹2𝑏delimited-[]𝜂32𝜁112𝜁𝜂2𝜁𝜁232subscript𝑍subscript𝐹4𝑏𝜂delimited-[]12𝜁subscript𝑍1\mathcal{D}=1+\tau+F_{0}(b)\left[\left(1-\frac{\omega_{\star}}{\omega}+\frac{3% }{2}\frac{\omega_{\star}^{T}}{\omega}\right)\sqrt[*]{\zeta}Z(\sqrt[*]{\zeta})-% \frac{\omega_{\star}^{T}}{\omega}\zeta Z_{+}\right]-\frac{\omega_{\star}^{T}}{% \omega}F_{2}(b)\sqrt[*]{\zeta}Z(\sqrt[*]{\zeta})+\\ +\frac{\omega_{\star}\bar{\omega}_{d}}{4\omega^{2}}\left\{F_{2}(b)\left[\eta% \left(\frac{3}{2}+\zeta\right)-1+\left(1+2\zeta+\eta\left(2\zeta(\zeta-2)-% \frac{3}{2}\right)\right)Z_{+}\right]+\right.\\ \left.+F_{4}(b)\eta\left[(1+2\zeta)Z_{+}-1\right]\right\}.start_ROW start_CELL caligraphic_D = 1 + italic_τ + italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_b ) [ ( 1 - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG + divide start_ARG 3 end_ARG start_ARG 2 end_ARG divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG ) nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG italic_Z ( nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG ) - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG italic_ζ italic_Z start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ] - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_b ) nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG italic_Z ( nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG ) + end_CELL end_ROW start_ROW start_CELL + divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG { italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_b ) [ italic_η ( divide start_ARG 3 end_ARG start_ARG 2 end_ARG + italic_ζ ) - 1 + ( 1 + 2 italic_ζ + italic_η ( 2 italic_ζ ( italic_ζ - 2 ) - divide start_ARG 3 end_ARG start_ARG 2 end_ARG ) ) italic_Z start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ] + end_CELL end_ROW start_ROW start_CELL + italic_F start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_b ) italic_η [ ( 1 + 2 italic_ζ ) italic_Z start_POSTSUBSCRIPT + end_POSTSUBSCRIPT - 1 ] } . end_CELL end_ROW (25)

where the Larmor radius functions are,

F0(b)=Γ0,subscript𝐹0𝑏subscriptΓ0\displaystyle F_{0}(b)=\Gamma_{0},italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_b ) = roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , (26)
F2(b)=(1b)Γ0+bΓ1,subscript𝐹2𝑏1𝑏subscriptΓ0𝑏subscriptΓ1\displaystyle F_{2}(b)=(1-b)\Gamma_{0}+b\Gamma_{1},italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_b ) = ( 1 - italic_b ) roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_b roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , (27)
F4(b)=2[(1b)2Γ0+(32b)bΓ1],subscript𝐹4𝑏2delimited-[]superscript1𝑏2subscriptΓ032𝑏𝑏subscriptΓ1\displaystyle F_{4}(b)=2\left[(1-b)^{2}\Gamma_{0}+\left(\frac{3}{2}-b\right)b% \Gamma_{1}\right],italic_F start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_b ) = 2 [ ( 1 - italic_b ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + ( divide start_ARG 3 end_ARG start_ARG 2 end_ARG - italic_b ) italic_b roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ] , (28)

with Γn=ebIn(b)subscriptΓ𝑛superscript𝑒𝑏subscript𝐼𝑛𝑏\Gamma_{n}=e^{-b}I_{n}(b)roman_Γ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT - italic_b end_POSTSUPERSCRIPT italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_b ), and Insubscript𝐼𝑛I_{n}italic_I start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the modified Bessel function of the first kind (Abramowitz & Stegun, 1968, §9.6). The shorthand notation Z+=Z/2=1+ζZ(ζ)subscript𝑍superscript𝑍21𝜁𝑍𝜁Z_{+}=-Z^{\prime}/2=1+\sqrt[*]{\zeta}Z(\sqrt[*]{\zeta})italic_Z start_POSTSUBSCRIPT + end_POSTSUBSCRIPT = - italic_Z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT / 2 = 1 + nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG italic_Z ( nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG ). In this form of the dispersion relation, we have presented all the terms that are, at least explicitly, order 1 or larger, taking ω/ωω/ω¯dsimilar-tosubscript𝜔𝜔𝜔subscript¯𝜔𝑑\omega_{\star}/\omega\sim\omega/\bar{\omega}_{d}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_ω ∼ italic_ω / over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT in ordering. Additional terms in the expansion could be included (see Tab. 1 in Appendix C), although these would stretch the original ordering in δ𝛿\deltaitalic_δ and ϵitalic-ϵ\epsilonitalic_ϵ and do not include any additional physics. This dispersion relation describes the behaviour of a linear localised ITG mode responding to a quadratic magnetic curvature with both a good and bad curvature. The dispersion is obtained through ordering of the localisation and the magnetic drift, which are taken to be strong and weak respectively. In addition, to reach such simple form, we consider what we refer to as a ‘pure’ mode, which simplifies the localisation response of the mode.

3.3 Computing the dispersion function and finding unstable modes

Some elements of the dispersion relation are reminiscent of the common local, slab ITG dispersion (Kadomtsev & Pogutse, 1970a) or local, short wavelength ITG (Smolyakov et al., 2002). We extend on these by a more careful consideration of mode localisation.

The information about the linear stability of our localised mode is encoded in the solutions to the dispersion relation 𝒟=0𝒟0\mathcal{D}=0caligraphic_D = 0. In general, this constitutes a complicated transcendental equation whose solutions need to be found numerically. The function 𝒟(ω)𝒟𝜔\mathcal{D}(\omega)caligraphic_D ( italic_ω ) can be evaluated in ω𝜔\omegaitalic_ω-space numerically using the definition of ζ𝜁\sqrt[*]{\zeta}nth-root start_ARG ∗ end_ARG start_ARG italic_ζ end_ARG above, and using one of the many efficient implementations of the plasma dispersion function (in this case, we use the function wofz (Johnson, 2024) in scipy.special). We accept as valid roots the values of ω𝜔\omegaitalic_ω for which |𝒟|=0𝒟0|\mathcal{D}|=0| caligraphic_D | = 0, which holds true whenever both its real and imaginary parts vanish. In Figure 2 we show some examples of what the dispersion function looks like in ω𝜔\omegaitalic_ω-space for some particularly simple cases in which no Larmor-radius effects are present.

Refer to caption
Figure 2: Dispersion function 𝒟𝒟\mathcal{D}caligraphic_D in ω𝜔\omegaitalic_ω space. The plots show |𝒟|𝒟|\mathcal{D}|| caligraphic_D | as a function of complex ω𝜔\omegaitalic_ω for different combinations of ω¯dsubscript¯𝜔𝑑\bar{\omega}_{d}over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT and ωTsuperscriptsubscript𝜔𝑇\omega_{\star}^{T}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT (all frequencies normalised to the transit frequency ωt=vTi/Λ2subscript𝜔𝑡subscript𝑣𝑇𝑖Λ2\omega_{t}=v_{Ti}/\Lambda\sqrt{2}italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT / roman_Λ square-root start_ARG 2 end_ARG). The set of three figures may be thus interpreted as the change in the instability due to an increase in ΛΛ\Lambdaroman_Λ, the width of the bad curvature region, where the positive {ω}𝜔\Im\{\omega\}roman_ℑ { italic_ω } part of the plane denotes instability. a) ωd/ωt=1×103subscript𝜔𝑑subscript𝜔𝑡1superscript103\omega_{d}/\omega_{t}=1\times 10^{-3}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = 1 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and ωT/ωt=1superscriptsubscript𝜔𝑇subscript𝜔𝑡1\omega_{\star}^{T}/\omega_{t}=-1italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = - 1, b) ωd/ωt=1×102subscript𝜔𝑑subscript𝜔𝑡1superscript102\omega_{d}/\omega_{t}=1\times 10^{-2}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = 1 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT and ωT/ωt=10superscriptsubscript𝜔𝑇subscript𝜔𝑡10\omega_{\star}^{T}/\omega_{t}=-10italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = - 10, c) ωd/ωt=1×101subscript𝜔𝑑subscript𝜔𝑡1superscript101\omega_{d}/\omega_{t}=1\times 10^{-1}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = 1 × 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and ωT/ωt=100superscriptsubscript𝜔𝑇subscript𝜔𝑡100\omega_{\star}^{T}/\omega_{t}=-100italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = - 100. In this case we have chosen b=0𝑏0b=0italic_b = 0, τ=1𝜏1\tau=1italic_τ = 1 and ω=0subscript𝜔0\omega_{\star}=0italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = 0 for simplicity. The red line represents one of the branch cuts; the vertical branch cut is not present in the domain plotted, as the unstable modes live near ω=0𝜔0\omega=0italic_ω = 0 as shown.

A single unstable mode with {ω}<0𝜔0\Re\{\omega\}<0roman_ℜ { italic_ω } < 0 is seen in the plots of Fig. 2. As the drift and diamagnetic frequencies are varied, the mode location evolves in ω𝜔\omegaitalic_ω-space. In fact, if we interpret the plots (a) to (c) as the change due to increasing ΛΛ\Lambdaroman_Λ, the width of the bad curvature region along the field line, we see that decreasing ΛΛ\Lambdaroman_Λ below a certain threshold appears to lead to the unstable mode eventually vanishing. The evolution of this instability with ΛΛ\Lambdaroman_Λ and the other parameters is what we are ultimately interested in, and shall be the main focus of our study in the following section. To automate the root-find of 𝒟𝒟\mathcal{D}caligraphic_D and be able to study the various interesting mode dependencies, we implement the following approach: i) we evaluate |𝒟|𝒟|\mathcal{D}|| caligraphic_D | in a coarse-grained ω𝜔\omegaitalic_ω-space roughly bounded by ωTsuperscriptsubscript𝜔𝑇\omega_{\star}^{T}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT, ii) we find the regions of lowest |𝒟|𝒟|\mathcal{D}|| caligraphic_D | residual, iii) we perform local least-squares minimisation around them, and iv) from the multiple roots found we select the most unstable one (see diagram in Fig. 3).

Refer to caption
Figure 3: Diagram sketching the procedure to find the most unstable mode. This algorithm is used when numerical roots of 𝒟𝒟\mathcal{D}caligraphic_D are required.

4 Physics and features of the localised kinetic ITG mode

In this section we investigate the behaviour of the localised ITG mode, including kinetic effects, by exploring Eq. (25). By inspection,the modes can possibly depend on the following parameters: the diamagnetic drive ωsubscript𝜔\omega_{\star}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT and ωTsuperscriptsubscript𝜔𝑇\omega_{\star}^{T}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT (density and temperature gradients respectively), the curvature drift magnitude ω¯dsubscript¯𝜔𝑑\bar{\omega}_{d}over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, the transit frequency ωt=vTi/Λ2subscript𝜔𝑡subscript𝑣𝑇𝑖Λ2\omega_{t}=v_{Ti}/\Lambda\sqrt{2}italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT / roman_Λ square-root start_ARG 2 end_ARG, the poloidal wavenumber kαsubscript𝑘𝛼k_{\alpha}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT and the ratio of electron to ion temperature τ𝜏\tauitalic_τ.

The frequencies and length-scales will be presented normalised to a reference transit frequency, ωt0subscript𝜔𝑡0\omega_{t0}italic_ω start_POSTSUBSCRIPT italic_t 0 end_POSTSUBSCRIPT. That is, the frequency characteristic of the motion of a thermal ion over a distance Λ0subscriptΛ0\Lambda_{0}roman_Λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT along the field line, ωt0=vTi/Λ02subscript𝜔𝑡0subscript𝑣𝑇𝑖subscriptΛ02\omega_{t0}=v_{Ti}/\Lambda_{0}\sqrt{2}italic_ω start_POSTSUBSCRIPT italic_t 0 end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT / roman_Λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT square-root start_ARG 2 end_ARG. We take this lengthscale Λ0subscriptΛ0\Lambda_{0}roman_Λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to normalise ΛΛ\Lambdaroman_Λ as well. The normalisation of the poloidal wavenumber is somewhat more complicated. Let us recall the definition of kαsubscript𝑘𝛼k_{\alpha}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT from the covariant form of the wavevector 𝐤=kαα+kψψsubscript𝐤perpendicular-tosubscript𝑘𝛼𝛼subscript𝑘𝜓𝜓\mathbf{k}_{\perp}=k_{\alpha}\nabla\alpha+k_{\psi}\nabla\psibold_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∇ italic_α + italic_k start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT ∇ italic_ψ. We took for simplicity kψ=0subscript𝑘𝜓0k_{\psi}=0italic_k start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT = 0, so that 𝐤=kααsubscript𝐤perpendicular-tosubscript𝑘𝛼𝛼\mathbf{k}_{\perp}=k_{\alpha}\nabla\alphabold_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∇ italic_α. The parameter kαsubscript𝑘𝛼k_{\alpha}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is dimensionless, leaving us with the uncomfortable situation of kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT having units of length. When we write kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT in what follows we adopt the convention of meaning kαρi¯=kαρi|α|¯subscript𝑘𝛼subscript𝜌𝑖subscript𝑘𝛼subscript𝜌𝑖𝛼\overline{k_{\alpha}\rho_{i}}=k_{\alpha}\rho_{i}|\nabla\alpha|over¯ start_ARG italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG = italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | ∇ italic_α | (we will often drop the overbar notation, but it should be clear when we do so). The Larmor radius parameter becomes b=(kαρi¯)2/2𝑏superscript¯subscript𝑘𝛼subscript𝜌𝑖22b=(\overline{k_{\alpha}\rho_{i}})^{2}/2italic_b = ( over¯ start_ARG italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2, which does not exactly match the convention in other works in which the minor radius scale a𝑎aitalic_a is chosen to normalise kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT.

We now construct the linear spectrum of the localised ITG mode as a function of kαρi¯¯subscript𝑘𝛼subscript𝜌𝑖\overline{k_{\alpha}\rho_{i}}over¯ start_ARG italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG, as the width of the bad curvature region is changed. Some examples are presented in Figures 4 and 5. For these cases we have chosen a strongly driven scenario, ωT/ω¯d=102superscriptsubscript𝜔𝑇subscript¯𝜔𝑑superscript102\omega_{\star}^{T}/\bar{\omega}_{d}=10^{2}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, in the presence of no density gradient.

Refer to caption
Figure 4: Evolution of the linear spectrum of unstable ITG with the bad curvature region size, ΛΛ\Lambdaroman_Λ. The plots show a) the growth rate and b) negative the real frequency of the most unstable mode as a function of kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and ΛΛ\Lambdaroman_Λ. The frequencies are normalised to ωt0=vTi/Λ0subscript𝜔𝑡0subscript𝑣𝑇𝑖subscriptΛ0\omega_{t0}=v_{Ti}/\Lambda_{0}italic_ω start_POSTSUBSCRIPT italic_t 0 end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT / roman_Λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, where Λ0subscriptΛ0\Lambda_{0}roman_Λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is some reference length. We only plot points when the mode satisfies the conditions γ>0𝛾0\gamma>0italic_γ > 0 and |ωd/ω|<1subscript𝜔𝑑𝜔1|\omega_{d}/\omega|<1| italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω | < 1. The plots are constructed for the choice ωT/ωt0=10superscriptsubscript𝜔𝑇subscript𝜔𝑡010\omega_{\star}^{T}/\omega_{t0}=-10italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t 0 end_POSTSUBSCRIPT = - 10, ωd/ωt0=0.1subscript𝜔𝑑subscript𝜔𝑡00.1\omega_{d}/\omega_{t0}=0.1italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t 0 end_POSTSUBSCRIPT = 0.1, ω/ωt0=0.0subscript𝜔subscript𝜔𝑡00.0\omega_{\star}/\omega_{t0}=0.0italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t 0 end_POSTSUBSCRIPT = 0.0 and τ=1𝜏1\tau=1italic_τ = 1.
Refer to caption
Figure 5: Properties of the unstable modes as a function of the bad curvature region size, ΛΛ\Lambdaroman_Λ, and kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for γ>0𝛾0\gamma>0italic_γ > 0 and |ωd/ω|<1subscript𝜔𝑑𝜔1|\omega_{d}/\omega|<1| italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω | < 1. The plots show, clockwise starting from the top left, the growth rate γ𝛾\gammaitalic_γ, the real frequency ωrsubscript𝜔𝑟-\omega_{r}- italic_ω start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, the Gaussian envelope scale |λ|𝜆|\lambda|| italic_λ |, the approximation scale {λ}|ω/ωd|𝜆𝜔subscript𝜔𝑑\Re\{\lambda\}|\omega/\omega_{d}|roman_ℜ { italic_λ } | italic_ω / italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT |, the kinetic measure |ζ|𝜁|\zeta|| italic_ζ | and the small scale |ωd/ω|subscript𝜔𝑑𝜔|\omega_{d}/\omega|| italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω |. The top left and middle plots can be interpreted as top views of Fig. 4. The red broken line in the bottom left plot is the estimate of the Landau threshold as detailed in Sec. 4.3. The blue region in the bottom right plot shows where we expect our localised mode approximation to break down. Here ωT/ωt0=10superscriptsubscript𝜔𝑇subscript𝜔𝑡010\omega_{\star}^{T}/\omega_{t0}=-10italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t 0 end_POSTSUBSCRIPT = - 10, ωd/ωt0=0.1subscript𝜔𝑑subscript𝜔𝑡00.1\omega_{d}/\omega_{t0}=0.1italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t 0 end_POSTSUBSCRIPT = 0.1, ω/ωt0=0.0subscript𝜔subscript𝜔𝑡00.0\omega_{\star}/\omega_{t0}=0.0italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t 0 end_POSTSUBSCRIPT = 0.0 and τ=1𝜏1\tau=1italic_τ = 1.

4.1 General features and assessment of the approximation

The linear spectra obtained have four distinctive features that are most clear in the limit of a wide bad curvature region. At small kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, as both the diamagnetic drive and the bad curvature diminish (given that they are linear functions of it), so does the instability, Fig. 4a. The mode remains co-rotating with the diamagnetic frequency, Fig. 4b, but eventually, for small enough frequency, it reaches what we may call the Landau threshold. This occurs when at some critical kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, the frequency matches the parallel transit frequency leading to Landau dam** that stabilises the localised mode. As a result, as this threshold is approached and the kinetic aspects of the problem grow (|ζ|1𝜁1|\zeta|\rightarrow 1| italic_ζ | → 1), the ITG tends to become increasingly stretched over the field line (|λ|𝜆|\lambda|| italic_λ | decreases). As it does so, the ever increasing good curvature of our model (which serves a role similar to that which the magnetic shear would play) dictates its behaviour, as the key approximation {λ}|ω/ωd|1/ϵ1similar-to𝜆𝜔subscript𝜔𝑑1italic-ϵmuch-greater-than1\Re\{\lambda\}|\omega/\omega_{d}|\sim 1/\epsilon\gg 1roman_ℜ { italic_λ } | italic_ω / italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT | ∼ 1 / italic_ϵ ≫ 1 fails. This naturally lends to the possibility of the behaviour in this limit being dominated by de-localised, slab-like ITG modes such as Floquet modes (Zocco et al., 2016). The precise value of kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT at which the threshold for the localised mode occurs cannot be expected to be exactly described by the model, but we may use its trends as an informative guide.

As we increase the poloidal wavenumber, the increase in the diamagnetic and curvature drifts drives the ITG more vigorously. The growth rate increases as the mode becomes more localised; the gains from becoming localised about the bad curvature energy source overcome the effects of increasing the effective ksubscript𝑘parallel-tok_{\parallel}italic_k start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT that invigorates Landau dam**. A larger kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT does however also enhance finite-Larmor-radius effects. The stabilisation effects of the geometry can suppress the ITG mode. This leads to the small-kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT peak (kαρi1less-than-or-similar-tosubscript𝑘𝛼subscript𝜌𝑖1k_{\alpha}\rho_{i}\lesssim 1italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≲ 1) in the spectrum, Fig. 4a. We may think of this as the standard ITG peak, and refer to the threshold (or dip) to its right as FLR stabilisation threshold.

Interestingly, increasing kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT does not continue to reinforce the stabilising effect of the field geometry (in our case what one can refer to as flux compression |α|1/|ψ|similar-to𝛼1𝜓|\nabla\alpha|\sim 1/|\nabla\psi|| ∇ italic_α | ∼ 1 / | ∇ italic_ψ |). There is some point at which the instability drive beats the FLR effects again, and the mode starts to become more and more unstable. This threshold we refer to as the FLR weakening threshold. As the mode keeps growing, it becomes increasingly localised near the bad curvature region, with its characteristic mode frequency rising sharply to settle to a roughly constant value, Fig. 4b. Eventually, the mode reaches a last critical kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT threshold. This corresponds to the situation in which the mode resonates with ωω¯dsimilar-to𝜔subscript¯𝜔𝑑\omega\sim-\bar{\omega}_{d}italic_ω ∼ - over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, Fig. 5. We call this the ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT threshold. Once again, as the kinetic effects grow and the threshold is approached, the fidelity of the model falters, in this case primarily because |ω¯d/ω|1similar-tosubscript¯𝜔𝑑𝜔1|\bar{\omega}_{d}/\omega|\sim 1| over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω | ∼ 1. This leads to a second ‘hump’ in the linear spectrum of the ITG instability. Such behaviour has been previously studied in the context of the so-called SWITG instability (Hirose et al., 2002; Smolyakov et al., 2002; Gao et al., 2003).

These features and their location evolve as the width of the bad curvature region changes. They do so in such a way that, as the bad curvature region narrows, the two instability peaks move towards each other, merge and eventually, below a certain critical ΛΛ\Lambdaroman_Λ, disappear. That is, the localised mode is stabilised by shortening the bad curvature region sufficiently. Unfortunately, and as it occurred near the kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT threshold, as ΛΛ\Lambdaroman_Λ becomes smaller, the mode becomes increasingly delocalised, and the model presented becomes a poor description. We may expect de-localised modes to gain prominence and persist in this limit.

4.2 Fluid limit

Let us start the more quantitative analysis by checking what we obtain in the fluid limit, that is, when all resonances can be neglected. This will prove useful in two ways. First, because it serves as a good check that the dispersion relation reproduces a correct asymptotic limit. Second, because many of the features of our linear spectra may be explained in the simplest of terms through the fluid perspective, as we shall see.

With this in mind, let us start by first stating what we mean in this context by the fluid limit: this is the limit of |ζ|𝜁|\zeta|\rightarrow\infty| italic_ζ | → ∞, i.e., values far from the kinetic Landau resonance. We therefore need to expand 𝒟𝒟\mathcal{D}caligraphic_D in Eq. (25) for large ζ𝜁\zetaitalic_ζ. Since (Fried & Conte, 2015)

Z(x)1x[1+12x2+34x4+],𝑍𝑥1𝑥delimited-[]112superscript𝑥234superscript𝑥4Z(x)\approx-\frac{1}{x}\left[1+\frac{1}{2x^{2}}+\frac{3}{4x^{4}}+\dots\right],italic_Z ( italic_x ) ≈ - divide start_ARG 1 end_ARG start_ARG italic_x end_ARG [ 1 + divide start_ARG 1 end_ARG start_ARG 2 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 3 end_ARG start_ARG 4 italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG + … ] , (29)

for large argument, then

𝒟1+τF0(b)[1ωω(1η)]+ωTωF2(b)12ζ[F0(b)(1ωω)ωTωF2(b)]+ωω¯d2ω2[(1η)F2(b)+ηF4(b)].\mathcal{D}\approx 1+\tau-F_{0}(b)\left[1-\frac{\omega_{\star}}{\omega}(1-\eta% )\right]+\frac{\omega_{\star}^{T}}{\omega}F_{2}(b)-\frac{1}{2\zeta}\left[F_{0}% (b)\left(1-\frac{\omega_{\star}}{\omega}\right)-\frac{\omega_{\star}^{T}}{% \omega}F_{2}(b)\right]+\\ -\frac{\omega_{\star}\bar{\omega}_{d}}{2\omega^{2}}\left[(1-\eta)F_{2}(b)+\eta F% _{4}(b)\right].start_ROW start_CELL caligraphic_D ≈ 1 + italic_τ - italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_b ) [ 1 - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ( 1 - italic_η ) ] + divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_b ) - divide start_ARG 1 end_ARG start_ARG 2 italic_ζ end_ARG [ italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_b ) ( 1 - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_b ) ] + end_CELL end_ROW start_ROW start_CELL - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG [ ( 1 - italic_η ) italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_b ) + italic_η italic_F start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_b ) ] . end_CELL end_ROW (30)

where,

12ζ=(ωtλω)2ω¯d2ω=ω¯d1/2ωtω3/2ω¯d2ω.12𝜁superscriptsubscript𝜔𝑡𝜆𝜔2subscript¯𝜔𝑑2𝜔superscriptsubscript¯𝜔𝑑12subscript𝜔𝑡superscript𝜔32subscript¯𝜔𝑑2𝜔\frac{1}{2\zeta}=\left(\frac{\omega_{t}\sqrt{\lambda}}{\omega}\right)^{2}-% \frac{\bar{\omega}_{d}}{2\omega}=\frac{\bar{\omega}_{d}^{1/2}\omega_{t}}{% \omega^{3/2}}-\frac{\bar{\omega}_{d}}{2\omega}.divide start_ARG 1 end_ARG start_ARG 2 italic_ζ end_ARG = ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT square-root start_ARG italic_λ end_ARG end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG = divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG . (31)

With this explicit form, we may then write the fluid form of the dispersion relation explicitly,

𝒟(1+τΓ0)+ωω[F0(b)(1η)+ηF2(b)]+ω¯dωω2[Γ0+b2(Γ0Γ1)η2F4(b)]++(ω¯dω)1/2ωtωω2[F0(b)+ηF2(b)],𝒟1𝜏subscriptΓ0subscript𝜔𝜔delimited-[]subscript𝐹0𝑏1𝜂𝜂subscript𝐹2𝑏subscript¯𝜔𝑑subscript𝜔superscript𝜔2delimited-[]subscriptΓ0𝑏2subscriptΓ0subscriptΓ1𝜂2subscript𝐹4𝑏superscriptsubscript¯𝜔𝑑𝜔12subscript𝜔𝑡subscript𝜔superscript𝜔2delimited-[]subscript𝐹0𝑏𝜂subscript𝐹2𝑏\mathcal{D}\approx(1+\tau-\Gamma_{0})+\frac{\omega_{\star}}{\omega}\left[F_{0}% (b)(1-\eta)+\eta F_{2}(b)\right]+\frac{\bar{\omega}_{d}\omega_{\star}}{\omega^% {2}}\left[-\Gamma_{0}+\frac{b}{2}(\Gamma_{0}-\Gamma_{1})-\frac{\eta}{2}F_{4}(b% )\right]+\\ +\left(\frac{\bar{\omega}_{d}}{\omega}\right)^{1/2}\frac{\omega_{t}\omega_{% \star}}{\omega^{2}}\left[F_{0}(b)+\eta F_{2}(b)\right],start_ROW start_CELL caligraphic_D ≈ ( 1 + italic_τ - roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG [ italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_b ) ( 1 - italic_η ) + italic_η italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_b ) ] + divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG [ - roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG italic_b end_ARG start_ARG 2 end_ARG ( roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) - divide start_ARG italic_η end_ARG start_ARG 2 end_ARG italic_F start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_b ) ] + end_CELL end_ROW start_ROW start_CELL + ( divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG [ italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_b ) + italic_η italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_b ) ] , end_CELL end_ROW (32)

where we have dropped terms that are order ω¯d/ωsubscript¯𝜔𝑑𝜔\bar{\omega}_{d}/\omegaover¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω. We have an algebraic equation whose roots one can straightforwardly find. Presented in this form of Eq. (32), the dispersion function may appear obscure, however it agrees with Connor et al. (1980) (see Appendix D). This evidences the particular significance of λ𝜆\lambdaitalic_λ, which provides our construction with the right fluid limit.

It is commonplace to consider the case of a vigorously temperature-gradient driven limit (namely, ωT/ω1much-greater-thansuperscriptsubscript𝜔𝑇𝜔1\omega_{\star}^{T}/\omega\gg 1italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / italic_ω ≫ 1), with small Larmor radius effects, b1much-less-than𝑏1b\ll 1italic_b ≪ 1. We may make a direct comparison in this limit to (Romanelli, 1989, Eq. (19))666More precisely, for a fair comparison, one should pick for the ordering ω/ωTω¯d/ω(ωt/ω)2bδsimilar-to𝜔superscriptsubscript𝜔𝑇subscript¯𝜔𝑑𝜔similar-tosuperscriptsubscript𝜔𝑡𝜔2similar-to𝑏similar-to𝛿\omega/\omega_{\star}^{T}\sim\bar{\omega}_{d}/\omega\sim(\omega_{t}/\omega)^{2% }\sim b\sim\deltaitalic_ω / italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ∼ over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω ∼ ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∼ italic_b ∼ italic_δ. These are the assumptions in Plunk et al. (2014). Note that there is a difference in the sign of ω¯dsubscript¯𝜔𝑑\bar{\omega}_{d}over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT between the two from their respective definition. In fact, this very same limit is achieved even if one ignores the ωω¯dsubscript𝜔subscript¯𝜔𝑑\omega_{\star}\bar{\omega}_{d}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT term in the dispersion function, Eq. (25), although one would not agree with the full Larmor radius expressions in Appendix D.,

τω3bωTω2ω¯dωTω+ωTωtωω¯d0.𝜏superscript𝜔3𝑏superscriptsubscript𝜔𝑇superscript𝜔2subscript¯𝜔𝑑superscriptsubscript𝜔𝑇𝜔superscriptsubscript𝜔𝑇subscript𝜔𝑡𝜔subscript¯𝜔𝑑0\tau\omega^{3}-b\omega_{\star}^{T}\omega^{2}-\bar{\omega}_{d}\omega_{\star}^{T% }\omega+\omega_{\star}^{T}\omega_{t}\sqrt{\omega\bar{\omega}_{d}}\approx 0.italic_τ italic_ω start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - italic_b italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_ω + italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT square-root start_ARG italic_ω over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG ≈ 0 . (33)

The fluid limit of the dispersion function in Eq. (25) is therefore correct. Given this connection, we may extend the common nomenclature distinguishing between the slab and toroidal ITG modes interpreted as follow (Wesson & Campbell, 2011; Zocco et al., 2016). Let the mode in which the streaming terms are ignorable777Specific orderings were given by Zocco et al. (2016). be referred to as the toroidal ITG mode; and let the slab ITG mode correspond to the reverse. Formally, this distinction is dictated by the relative importance of the last two terms in Eq. (33), which is simply |λ|𝜆|\lambda|| italic_λ |. Thus, when the mode has significant structure within the bad curvature well (|λ|>1𝜆1|\lambda|>1| italic_λ | > 1), we shall refer to the mode as toroidal (note that this mode is not necessarily strongly localised, which would be a statement about {λ}𝜆\Re\{\lambda\}roman_ℜ { italic_λ }). The slab modes will tend to be delocalised and thus care mostly about the larger |¯|¯|\bar{\ell}|| over¯ start_ARG roman_ℓ end_ARG | good curvature part of the problem.

4.3 The Landau threshold

We discussed qualitatively the presence of a cutoff at long wavelengths kαρi1much-less-thansubscript𝑘𝛼subscript𝜌𝑖1k_{\alpha}\rho_{i}\ll 1italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≪ 1 in the linear ITG mode spectrum. We argued that this threshold was indicative of a stabilisation of the mode by Landau dam**; that is, that the localised mode near the threshold becomes resonant with the parallel streaming frequency. Although very near the threshold the construction of the dispersion relation in Eq. (25) is not formally valid and de-localised modes may exist, we may nevertheless use the model to estimate where this threshold for the localised modes occurs, and how it is affected by the various properties of the field.

We thus need to find a real frequency solution to the equation 𝒟(ω)=0𝒟𝜔0\mathcal{D}(\omega)=0caligraphic_D ( italic_ω ) = 0 (i.e., the mode that as in Fig. 2 is just touching the ω𝜔\omegaitalic_ω-space real axis). Typically, such as in the standard local slab ITG, the real nature of the frequency makes the arguments of the plasma dispersion functions real, and separating the real and imaginary parts of the dispersion relation becomes straightforward, without the requirement of solving any transcendental equation. In the present case, though, the argument of the plasma dispersion function, Eq. (25), involves the square root of ω𝜔\omegaitalic_ω. As the mode is driven by the temperature gradient, ωTsuperscriptsubscript𝜔𝑇\omega_{\star}^{T}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT, ω<0𝜔0\omega<0italic_ω < 0 at the threshold, and thus Z(ζ)𝑍𝜁Z(\sqrt{\zeta})italic_Z ( square-root start_ARG italic_ζ end_ARG ) has both mixed real and imaginary parts. This makes solving a trascendental equation unavoidable.

We thus shall proceed by employing a physically motivated approach. As mentioned above, the Landau threshold corresponds to the resonance of the mode with the transit frequency, {ω}ωt.similar-to𝜔subscript𝜔𝑡\Re\{\omega\}\sim\omega_{t}.roman_ℜ { italic_ω } ∼ italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT . To be more precise, we have ζ1𝜁1\zeta\approx 1italic_ζ ≈ 1, and considering the mode to be slab-like in the region near the Landau threshold (i.e., rather elongated along the field line compared to the bad curvature region ΛΛ\Lambdaroman_Λ), we use, following Eq. (20a),

{ω}ωt2λ=vTiΛ/λ,𝜔subscript𝜔𝑡2𝜆subscript𝑣𝑇𝑖Λ𝜆\Re\{\omega\}\approx\omega_{t}\sqrt{2\lambda}=\frac{v_{Ti}}{\Lambda/\sqrt{% \lambda}},roman_ℜ { italic_ω } ≈ italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT square-root start_ARG 2 italic_λ end_ARG = divide start_ARG italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT end_ARG start_ARG roman_Λ / square-root start_ARG italic_λ end_ARG end_ARG , (34)

where Λ/λΛ𝜆\Lambda/\sqrt{\lambda}roman_Λ / square-root start_ARG italic_λ end_ARG is the characteristic longitudinal scale of the mode structure. The dependence of λ𝜆\lambdaitalic_λ on {ω}𝜔\Re\{\omega\}roman_ℜ { italic_ω } is known from Eq. (18). So what we ultimately need is some notion of the mode frequency, {ω}𝜔\Re\{\omega\}roman_ℜ { italic_ω }.

To estimate {ω}𝜔\Re\{\omega\}roman_ℜ { italic_ω }, we crudely assume that the kinetic effects mainly affect the stability of the mode, but leave the mode frequency to a large extent unaltered. Assuming a sufficiently large ΛΛ\Lambdaroman_Λ and ωT/ω¯dsuperscriptsubscript𝜔𝑇subscript¯𝜔𝑑\omega_{\star}^{T}/\bar{\omega}_{d}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, we use the slab branch of the fluid model to estimate {ω}𝜔\Re\{\omega\}roman_ℜ { italic_ω }, from Eq. (33),

{ω}(kαρi)3/5ω^d1/5(ωtω^Tτ)2/5cos(4π5),𝜔superscriptsubscript𝑘𝛼subscript𝜌𝑖35superscriptsubscript^𝜔𝑑15superscriptsubscript𝜔𝑡superscriptsubscript^𝜔𝑇𝜏254𝜋5\Re\{\omega\}\approx(k_{\alpha}\rho_{i})^{3/5}\hat{\omega}_{d}^{1/5}\left(% \frac{\omega_{t}\hat{\omega}_{\star}^{T}}{\tau}\right)^{2/5}\cos\left(\frac{4% \pi}{5}\right),roman_ℜ { italic_ω } ≈ ( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 / 5 end_POSTSUPERSCRIPT over^ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 5 end_POSTSUPERSCRIPT ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT over^ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_τ end_ARG ) start_POSTSUPERSCRIPT 2 / 5 end_POSTSUPERSCRIPT roman_cos ( divide start_ARG 4 italic_π end_ARG start_ARG 5 end_ARG ) , (35)

where the root with a negative real frequency must be chosen given the branch cut choice of the square root. The hat notation is meant to indicate that we have taken the kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT dependence of ω¯dsubscript¯𝜔𝑑\bar{\omega}_{d}over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT and ωTsuperscriptsubscript𝜔𝑇\omega_{\star}^{T}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT out explicitly. Although this might appear complicated, and the one fifth powers odd, the expression just found is precisely the frequency of a standard slab ITG mode with k=λ/Λsubscript𝑘parallel-to𝜆Λk_{\parallel}=\sqrt{\lambda}/\Lambdaitalic_k start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = square-root start_ARG italic_λ end_ARG / roman_Λ,888The precise critical threshold that one could estimate using the local slab ITG, using for the mode structure the expression for λ𝜆\lambdaitalic_λ in Eq. (18), is kαρi22(1+τ)(ωt/ω^T)τω¯d/ωTsubscript𝑘𝛼subscript𝜌𝑖221𝜏subscript𝜔𝑡superscriptsubscript^𝜔𝑇𝜏subscript¯𝜔𝑑superscriptsubscript𝜔𝑇k_{\alpha}\rho_{i}\approx 2\sqrt{2}(1+\tau)(\omega_{t}/\hat{\omega}_{\star}^{T% })\sqrt{\tau\bar{\omega}_{d}/\omega_{\star}^{T}}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≈ 2 square-root start_ARG 2 end_ARG ( 1 + italic_τ ) ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / over^ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ) square-root start_ARG italic_τ over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG. As emphasised in the text, this presents the same basic parameter dependence. the characteristic length scale of the mode. The main difference is that in our problem we have explicit field line dependence, while the pure slab ITG does not (other than an oscillating de-localised solution).

With this expression for the real frequency, we can reconsider Eq. (34), to write

(kαρi)Landau12.5τωtω^T(τω^dω^T)1/2.subscriptsubscript𝑘𝛼subscript𝜌𝑖Landau12.5𝜏subscript𝜔𝑡superscriptsubscript^𝜔𝑇superscript𝜏subscript^𝜔𝑑superscriptsubscript^𝜔𝑇12(k_{\alpha}\rho_{i})_{\mathrm{Landau}}\approx 12.5\tau\frac{\omega_{t}}{\hat{% \omega}_{\star}^{T}}\left(\frac{\tau\hat{\omega}_{d}}{\hat{\omega}_{\star}^{T}% }\right)^{1/2}.( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT roman_Landau end_POSTSUBSCRIPT ≈ 12.5 italic_τ divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG start_ARG over^ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG ( divide start_ARG italic_τ over^ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG over^ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT . (36)

As shown in Fig. 5, this is a fair estimate of the Landau threshold for the parameters considered, and most importantly, it shows the correct ΛΛ\Lambdaroman_Λ scaling. The threshold value increases as the bad curvature region becomes smaller, kαρi1/Λsimilar-tosubscript𝑘𝛼subscript𝜌𝑖1Λk_{\alpha}\rho_{i}\sim 1/\Lambdaitalic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∼ 1 / roman_Λ. Narrowing down the bad curvature well produces a narrower mode structure (the mode width goes like ΔΛ3/5similar-toΔsuperscriptΛ35\Delta\ell\sim\Lambda^{3/5}roman_Δ roman_ℓ ∼ roman_Λ start_POSTSUPERSCRIPT 3 / 5 end_POSTSUPERSCRIPT), Landau dam** becomes more effective, and thus the threshold increases. If the instability is driven more vigorously (i.e., we increase the temperature gradient drive, |ωT|superscriptsubscript𝜔𝑇|\omega_{\star}^{T}|| italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT |), then the parallel dynamics has a harder time stabilising the mode. The dependence on the temperature gradient of Eq. (36) is also in agreement with the numerical solution (see Fig. 6). Interestingly, increasing the curvature drift, and with it the magnitude of the bad curvature, does not worsen the threshold, but rather improve it. The reason is that the effect of ω¯dsubscript¯𝜔𝑑\bar{\omega}_{d}over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT in narrowing the mode is, in this limit, stronger than its direct drive of the instability. In this limit in which the mode is rather de-localised, both ω¯dsubscript¯𝜔𝑑\bar{\omega}_{d}over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT and ΛΛ\Lambdaroman_Λ should be interpreted to represent more ‘global’ properties of the geometry, as the (global) shear would.

Of course, near this threshold, the possibilities of other de-localised modes dominating the dynamics increases (Zocco et al., 2018, 2022; Podavini et al., 2023). In addition, the exact occurrence of the threshold will depend on the precise form of λ𝜆\lambdaitalic_λ, which we have already acknowledged the current model to only approximate. Thus the particular scaling of the critical kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and its value may change, but the main physical interpretation of its origin and dependence should remain.

4.4 The FLR stabilisation

The next noteworthy feature in the kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT spectrum is the stabilisation of the ITG mode as the finite Larmor radius becomes relevant. The result is the appearance of a characteristic peaked linear spectrum, with typical values of kαρi1less-than-or-similar-tosubscript𝑘𝛼subscript𝜌𝑖1k_{\alpha}\rho_{i}\lesssim 1italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≲ 1. This feature is not resonant-kinetic, and its presence in Figure 5 for large ΛΛ\Lambdaroman_Λ indicates it forms part of the fluid description of the instability.

In the fluid picture, the FLR stabilisation feature of the linear ITG spectrum results from the competition between the increased drive of the toroidal ITG and the increased stabilising efficiency of finite Larmor radius effects with increase poloidal wavenumber. Ignoring the streaming term in Eq. (33), the threshold can be shown to occur when,

(ωTb2τ)2+ωTω¯dτ0.superscriptsuperscriptsubscript𝜔𝑇𝑏2𝜏2superscriptsubscript𝜔𝑇subscript¯𝜔𝑑𝜏0\left(\frac{\omega_{\star}^{T}b}{2\tau}\right)^{2}+\frac{\omega_{\star}^{T}% \bar{\omega}_{d}}{\tau}\approx 0.( divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_b end_ARG start_ARG 2 italic_τ end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_τ end_ARG ≈ 0 . (37)

The critical poloidal wavenumber is then,

(kαρi)FLR2(τω¯dωT)1/4.subscriptsubscript𝑘𝛼subscript𝜌𝑖FLR2superscript𝜏subscript¯𝜔𝑑superscriptsubscript𝜔𝑇14(k_{\alpha}\rho_{i})_{\mathrm{FLR}}\approx 2\left(\frac{\tau\bar{\omega}_{d}}{% \omega_{\star}^{T}}\right)^{1/4}.( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT roman_FLR end_POSTSUBSCRIPT ≈ 2 ( divide start_ARG italic_τ over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 4 end_POSTSUPERSCRIPT . (38)

From the physical picture above, we expect this stabilisation threshold to have a value close to kαρi1similar-tosubscript𝑘𝛼subscript𝜌𝑖1k_{\alpha}\rho_{i}\sim 1italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∼ 1. A sign of this is the weak dependence on both the drift and the temperature gradient. In addition, this mechanism is not directly linked to the mode structure, and hence independent of ΛΛ\Lambdaroman_Λ. This latter resilience is apparent in Fig. 5. Only a small correction term may be observed within the fluid description of the mode by treating perturbatively the streaming contribution in Eq. (33), δ(kαρi)FLRτωt/2|ωT|𝛿subscriptsubscript𝑘𝛼subscript𝜌𝑖FLR𝜏subscript𝜔𝑡2superscriptsubscript𝜔𝑇\delta(k_{\alpha}\rho_{i})_{\mathrm{FLR}}\approx\tau\omega_{t}/2|\omega_{\star% }^{T}|italic_δ ( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT roman_FLR end_POSTSUBSCRIPT ≈ italic_τ italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / 2 | italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT |.

This resiliency of the FLR stabilisation threshold clashes with the rather fundamental dependence of the Landau threshold on ΛΛ\Lambdaroman_Λ. As a result, we expect the space in kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT available to the localised ITG mode to narrow down as the bad curvature region is squeezed. Eventually, the first peak in the linear spectrum would be eliminated, as in Fig. 4a, and thus at long wavelengths only extended ITG modes could remain present in the system. From the above, we may estimate the critical width of the bad curvature region, ΛΛ\Lambdaroman_Λ, at which this occurs. Extrapolating the behaviour of the threshold and the FLR stabilisation, Eqs. (36) and (38), the balance (kαρi)Landau(kαρi)FLRsimilar-tosubscriptsubscript𝑘𝛼subscript𝜌𝑖Landausubscriptsubscript𝑘𝛼subscript𝜌𝑖FLR(k_{\alpha}\rho_{i})_{\mathrm{Landau}}\sim(k_{\alpha}\rho_{i})_{\mathrm{FLR}}( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT roman_Landau end_POSTSUBSCRIPT ∼ ( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT roman_FLR end_POSTSUBSCRIPT yields

Λcrit6τvTiω^T(τω^dω^T)1/4.similar-tosubscriptΛcrit6𝜏subscript𝑣𝑇𝑖superscriptsubscript^𝜔𝑇superscript𝜏subscript^𝜔𝑑superscriptsubscript^𝜔𝑇14\Lambda_{\mathrm{crit}}\sim 6\tau\frac{v_{Ti}}{\hat{\omega}_{\star}^{T}}\left(% \frac{\tau\hat{\omega}_{d}}{\hat{\omega}_{\star}^{T}}\right)^{1/4}.roman_Λ start_POSTSUBSCRIPT roman_crit end_POSTSUBSCRIPT ∼ 6 italic_τ divide start_ARG italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT end_ARG start_ARG over^ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG ( divide start_ARG italic_τ over^ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG over^ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 4 end_POSTSUPERSCRIPT . (39)

This critical ΛΛ\Lambdaroman_Λ could also be rewritten in terms of a critical temperature gradient (ωT)crit4τ(ωt4ω¯d)1/5similar-tosubscriptsuperscriptsubscript𝜔𝑇crit4𝜏superscriptsuperscriptsubscript𝜔𝑡4subscript¯𝜔𝑑15(\omega_{\star}^{T})_{\mathrm{crit}}\sim 4\tau(\omega_{t}^{4}\bar{\omega}_{d})% ^{1/5}( italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT roman_crit end_POSTSUBSCRIPT ∼ 4 italic_τ ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 / 5 end_POSTSUPERSCRIPT. Given the preeminence of the Landau dam** physics, the critical temperature gradient threshold below which only extended modes are left at long wavelengths is particularly sensitive to the width of the bad curvature region. This emphasis in controlling the longitudinal spread of the mode aligns with some previous work on critical thresholds (Jenko et al., 2001; Roberg-Clark et al., 2022a).

4.5 The FLR weakening

The above considerations of the kinetic threshold and FLR stabilisation explain the presence of a peak in the linear spectrum of the mode in kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. However, as is clear from Figure 4, the ITG starts growing unstable once again at larger values of the poloidal wavenumber. This might appear surprising at first, because it is natural to think of FLR effects to increase monotonically with kαsubscript𝑘𝛼k_{\alpha}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, and thus after the FLR threshold, its stabilising effect to continue to exceed the increase in the turbulent drive. However, this picture is not correct.

The finite Larmor radius effects become weakened at large kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, when the relevant perpendicular scale starts to become significantly smaller than the Larmor radius. Larmor radius effects become inefficient, as they were for small kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Formally, one can ascribe this weakening of the FLR effects to the large argument behaviour of the Bessel functions introduced in the gyrokinetic equation by the gyroaveraging. The small b𝑏bitalic_b expansion of the Fn(b)subscript𝐹𝑛𝑏F_{n}(b)italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_b ) functions fails in this limit, and thus so does the fluid equation written in its typical form of Eq. (33). Retaining the appropriate behaviour leads to a critical FLR weakening threshold beyond which the mode grows back up999At this scale, the ion dynamics can more efficiently interact with electrons, which in principle should be retained kinetically. We will not explore these aspects in this work.. This type of ITG activity is referred to in the literature as short-wavelength ITG (SWITG) (Hirose et al., 2002; Smolyakov et al., 2002; Gao et al., 2003).

We shall then consider the large b𝑏bitalic_b limit of the fluid equation, Eq. (32) using the large argument asymptotics of the Bessel functions (Abramowitz & Stegun, 1968, Sec. 9.7),

2πb(1+τ)+ωω(1η2)34ωω¯dω2(1+η2)+(ω¯dω)1/2ωtωω2(1+η2)0.2𝜋𝑏1𝜏subscript𝜔𝜔1𝜂234subscript𝜔subscript¯𝜔𝑑superscript𝜔21𝜂2superscriptsubscript¯𝜔𝑑𝜔12subscript𝜔𝑡subscript𝜔superscript𝜔21𝜂20\sqrt{2\pi b}(1+\tau)+\frac{\omega_{\star}}{\omega}\left(1-\frac{\eta}{2}% \right)-\frac{3}{4}\frac{\omega_{\star}\bar{\omega}_{d}}{\omega^{2}}\left(1+% \frac{\eta}{2}\right)+\left(\frac{\bar{\omega}_{d}}{\omega}\right)^{1/2}\frac{% \omega_{t}\omega_{\star}}{\omega^{2}}\left(1+\frac{\eta}{2}\right)\approx 0.square-root start_ARG 2 italic_π italic_b end_ARG ( 1 + italic_τ ) + divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ( 1 - divide start_ARG italic_η end_ARG start_ARG 2 end_ARG ) - divide start_ARG 3 end_ARG start_ARG 4 end_ARG divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( 1 + divide start_ARG italic_η end_ARG start_ARG 2 end_ARG ) + ( divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( 1 + divide start_ARG italic_η end_ARG start_ARG 2 end_ARG ) ≈ 0 . (40)

In the limit of |λ|𝜆|\lambda|| italic_λ | being large (which is indeed the behaviour at large kαsubscript𝑘𝛼k_{\alpha}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, see Figure 5), we expect the streaming term in Eq. (40) to be small, and thus we are left with, once again, a quadratic in ω𝜔\omegaitalic_ω. The threshold then occurs when its discriminant vanishes, which gives

(kαρi)weak13π(1+τ)ωω¯d(1η/2)21+η/2.subscriptsubscript𝑘𝛼subscript𝜌𝑖weak13𝜋1𝜏subscript𝜔subscript¯𝜔𝑑superscript1𝜂221𝜂2(k_{\alpha}\rho_{i})_{\mathrm{weak}}\approx\frac{1}{3\sqrt{\pi}(1+\tau)}\frac{% \omega_{\star}}{\bar{\omega}_{d}}\frac{(1-\eta/2)^{2}}{1+\eta/2}.( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT roman_weak end_POSTSUBSCRIPT ≈ divide start_ARG 1 end_ARG start_ARG 3 square-root start_ARG italic_π end_ARG ( 1 + italic_τ ) end_ARG divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG divide start_ARG ( 1 - italic_η / 2 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 1 + italic_η / 2 end_ARG . (41)

A more refined version of this threshold kee** higher orders in 1/ζ1𝜁1/\zeta1 / italic_ζ yields for a flat density (kαρi)weak(0.6420.032ωT/ω¯d)/(1+τ)subscriptsubscript𝑘𝛼subscript𝜌𝑖weak0.6420.032superscriptsubscript𝜔𝑇subscript¯𝜔𝑑1𝜏(k_{\alpha}\rho_{i})_{\mathrm{weak}}\approx(0.642-0.032\omega_{\star}^{T}/\bar% {\omega}_{d})/(1+\tau)( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT roman_weak end_POSTSUBSCRIPT ≈ ( 0.642 - 0.032 italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) / ( 1 + italic_τ ). The exact numerical value here is however not important, especially given the ordering in δ𝛿\deltaitalic_δ and ϵitalic-ϵ\epsilonitalic_ϵ considered. The key is its dependence on the relative magnitude of the diamagnetic and curvature drift, which is quite strong compared to the FLR threshold (kαρi)FLR(ω¯d/ωT)1/4similar-tosubscriptsubscript𝑘𝛼subscript𝜌𝑖FLRsuperscriptsubscript¯𝜔𝑑superscriptsubscript𝜔𝑇14(k_{\alpha}\rho_{i})_{\mathrm{FLR}}\sim(\bar{\omega}_{d}/\omega_{\star}^{T})^{% 1/4}( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT roman_FLR end_POSTSUBSCRIPT ∼ ( over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 4 end_POSTSUPERSCRIPT, Eq. (38). Then, as the curvature of the field is increased or the driving temperature gradient reduced, the weakening FLR threshold will approach the FLR stabilisation threshold. Through a simple balance between (kαρi)FLR(kαρi)Weaksimilar-tosubscriptsubscript𝑘𝛼subscript𝜌𝑖FLRsubscriptsubscript𝑘𝛼subscript𝜌𝑖Weak(k_{\alpha}\rho_{i})_{\mathrm{FLR}}\sim(k_{\alpha}\rho_{i})_{\mathrm{Weak}}( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT roman_FLR end_POSTSUBSCRIPT ∼ ( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT roman_Weak end_POSTSUBSCRIPT, we expect to find the two regions merging (see Fig. 4) when |ωT|/ω¯d27(1+τ)4/5τ1/5similar-tosuperscriptsubscript𝜔𝑇subscript¯𝜔𝑑27superscript1𝜏45superscript𝜏15|\omega_{\star}^{T}|/\bar{\omega}_{d}\sim 27(1+\tau)^{4/5}\tau^{1/5}| italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT | / over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ∼ 27 ( 1 + italic_τ ) start_POSTSUPERSCRIPT 4 / 5 end_POSTSUPERSCRIPT italic_τ start_POSTSUPERSCRIPT 1 / 5 end_POSTSUPERSCRIPT. This is a rather large value of ωTsuperscriptsubscript𝜔𝑇\omega_{\star}^{T}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT, meaning that having two distinct peaks in the linear spectrum requires of a strongly driven regime (commpared to the drift).

Refer to caption
Figure 6: Properties of the unstable modes as a function of the temperature-gradient driven diamagnetic frequency, ωTsuperscriptsubscript𝜔𝑇\omega_{\star}^{T}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT, and kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The plots show, clockwise starting from the top left, the growth rate γ𝛾\gammaitalic_γ, the real frequency ωrsubscript𝜔𝑟-\omega_{r}- italic_ω start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, the Gaussian envelope scale |λ|𝜆|\lambda|| italic_λ |, the approximation scale {λ}|ω/ω¯d|𝜆𝜔subscript¯𝜔𝑑\Re\{\lambda\}|\omega/\bar{\omega}_{d}|roman_ℜ { italic_λ } | italic_ω / over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT |, the kinetic measure |ζ|𝜁|\zeta|| italic_ζ | and the small scale |ω¯d/ω|subscript¯𝜔𝑑𝜔|\bar{\omega}_{d}/\omega|| over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω |. The red broken line in the bottom left plot is the estimate of the Landau threshold as detailed in Sec. 4.3. The blue region in the bottom right plot shows where we expect our localised mode approximation to break down. This means that the precise instability threshold in ΛΛ\Lambdaroman_Λ cannot be fully trusted. We only plot points when the mode satisfies the conditions γ>0𝛾0\gamma>0italic_γ > 0 and |ω¯d/ω|<1subscript¯𝜔𝑑𝜔1|\bar{\omega}_{d}/\omega|<1| over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω | < 1. The plots are constructed for the choice ω¯d/ωt=1.0subscript¯𝜔𝑑subscript𝜔𝑡1.0\bar{\omega}_{d}/\omega_{t}=1.0over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = 1.0, ω/ωt0=0.0subscript𝜔subscript𝜔𝑡00.0\omega_{\star}/\omega_{t0}=0.0italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t 0 end_POSTSUBSCRIPT = 0.0 and τ=1𝜏1\tau=1italic_τ = 1.

In Figure 6 we illustrate some of these linear spectra dynamics as a a function of a changing temperature gradient.

4.6 The ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT threshold

At even larger kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT we clearly have another stabilisation effect that leads to the appearance of an instability threshold. The responsible mechanism is not captured in the fluid picture, and is kinetic in nature, as the value of |ζ|𝜁|\zeta|| italic_ζ | in Figure 5 suggests. It could be tempting from our previous discussion on the Landau threshold to suggest that a similar Landau dam** mechanism is present here as well. However, as the threshold is approached, the ITG mode does not seem to exhibit a clear tendency to relax its longitudinal structure. In fact, as |λ|1much-greater-than𝜆1|\lambda|\gg 1| italic_λ | ≫ 1 the mode retains a fine structure. What is then the mechanism in action? To answer the question it suffices to look back at the definition of ζ𝜁\zetaitalic_ζ in Eq. (20a), which in this limit gives ζω/ω¯d𝜁𝜔subscript¯𝜔𝑑\zeta\approx-\omega/\bar{\omega}_{d}italic_ζ ≈ - italic_ω / over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT. Thus, kinetic effects at large kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are dominated by the resonance of the mode rotation with the bad curvature of the field. Note that this resonance has been retained even in our expansion in ω¯d/ωδ1similar-tosubscript¯𝜔𝑑𝜔𝛿much-less-than1\bar{\omega}_{d}/\omega\sim\delta\ll 1over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω ∼ italic_δ ≪ 1, as can be recognised by inspection of our kinetic equation Eq. (17b).

With the dominant mechanism identified, we may estimate at what wavenumbers the mode frequency balances the drift frequency. The large kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT limit of the fluid equation Eq. (40) yields a roughly constant real frequency for the mode, which will eventually be matched by the magnetic drift, which grows with the wavenumber ω¯dkαρiproportional-tosubscript¯𝜔𝑑subscript𝑘𝛼subscript𝜌𝑖\bar{\omega}_{d}\propto k_{\alpha}\rho_{i}over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ∝ italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Using the higher order cubic version of Eq. (40), the balance {ω}ω¯dsimilar-to𝜔subscript¯𝜔𝑑\Re\{\omega\}\sim-\bar{\omega}_{d}roman_ℜ { italic_ω } ∼ - over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT gives

(kαρi)ω¯d1211+τ|ωTω¯d|.subscriptsubscript𝑘𝛼subscript𝜌𝑖subscript¯𝜔𝑑1211𝜏superscriptsubscript𝜔𝑇subscript¯𝜔𝑑(k_{\alpha}\rho_{i})_{\bar{\omega}_{d}}\approx\frac{1}{2}\frac{1}{1+\tau}\left% |\frac{\omega_{\star}^{T}}{\bar{\omega}_{d}}\right|.( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≈ divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG 1 end_ARG start_ARG 1 + italic_τ end_ARG | divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG | . (42)

The multiplicative factors in front depend on the details of the approximation of the model, but what is key is the dependence of the threshold on the ratio ωT/ω¯dsuperscriptsubscript𝜔𝑇subscript¯𝜔𝑑\omega_{\star}^{T}/\bar{\omega}_{d}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT (which correctly describes the behaviour in Fig. 6). The behaviour of this threshold suggests a narrowing of the linear spectrum that shall reach a critical narrow range (kαρi)critsuperscriptsubscript𝑘𝛼subscript𝜌𝑖crit(k_{\alpha}\rho_{i})^{\mathrm{crit}}( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT roman_crit end_POSTSUPERSCRIPT when

ωTω¯d|crit2(kαρi)crit(1+τ),evaluated-atsuperscriptsubscript𝜔𝑇subscript¯𝜔𝑑crit2superscriptsubscript𝑘𝛼subscript𝜌𝑖crit1𝜏\left.\frac{\omega_{\star}^{T}}{\bar{\omega}_{d}}\right|_{\mathrm{crit}}% \approx 2(k_{\alpha}\rho_{i})^{\mathrm{crit}}\left(1+\tau\right),divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG | start_POSTSUBSCRIPT roman_crit end_POSTSUBSCRIPT ≈ 2 ( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT roman_crit end_POSTSUPERSCRIPT ( 1 + italic_τ ) , (43)

or in tokamak notation R/LT𝑅subscript𝐿𝑇R/L_{T}italic_R / italic_L start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT, which is compatible with Romanelli (1989) and Guo & Romanelli (1993) results, with a visual estimate obtained from Fig. (2) of Biglari et al. (1989), with the Jenko-Dorland-Hammett formula (Jenko et al., 2001), and other critical gradient estimates (Roberg-Clark et al., 2022a).

In this consideration there appears not to be any direct involvement of the parallel streaming dynamics. However, as the curvature well is narrowed the two kinetic elements in the problem become mixed. Signs of this behaviour are seen in the apparent decorrelation between the ω/ω¯d𝜔subscript¯𝜔𝑑\omega/\bar{\omega}_{d}italic_ω / over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ratio and the threshold in Figure 5 when ΛΛ\Lambdaroman_Λ starts having an effect on the mode. In fact, in the above description of a critical temperature threshold, we considered some reference critical value of (kαρi)critsubscriptsubscript𝑘𝛼subscript𝜌𝑖crit(k_{\alpha}\rho_{i})_{\mathrm{crit}}( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT roman_crit end_POSTSUBSCRIPT. Following Figure 6 though, one can expect at some point the ω¯dsubscript¯𝜔𝑑\bar{\omega}_{d}over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT threshold to become close to the Landau threshold, and not just an arbitrarily chosen wavenumber. When this occurs, we may say that there will be no more localised ITG modes. In this case the threshold would scale as,

(ωT)crit(ω¯d3ωt2)1/5.similar-tosubscriptsuperscriptsubscript𝜔𝑇critsuperscriptsuperscriptsubscript¯𝜔𝑑3superscriptsubscript𝜔𝑡215(\omega_{\star}^{T})_{\mathrm{crit}}\sim(\bar{\omega}_{d}^{3}\omega_{t}^{2})^{% 1/5}.( italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT roman_crit end_POSTSUBSCRIPT ∼ ( over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 5 end_POSTSUPERSCRIPT . (44)

Thus, the threshold changes because we can affect it not only by making the resonance with ω¯dsubscript¯𝜔𝑑\bar{\omega}_{d}over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT appear at longer wavelengths, but also by amplifying the effects involved in the Landau threshold. As a result, increasing the bad curvature or the parallel scale both will have a positive effect on reducing the ITG, although extended modes may persist. The particular scaling obtained by balancing the two physics ingredients goes beyond Biglari et al. (1989) and R/LTO(1)similar-to𝑅subscript𝐿𝑇𝑂1R/L_{T}\sim O(1)italic_R / italic_L start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ∼ italic_O ( 1 ), where R𝑅Ritalic_R is the major radius. In our case, the balance yields R/LT1/q0.6similar-to𝑅subscript𝐿𝑇1superscript𝑞0.6R/L_{T}\sim 1/q^{0.6}italic_R / italic_L start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ∼ 1 / italic_q start_POSTSUPERSCRIPT 0.6 end_POSTSUPERSCRIPT, involving the safety factor q𝑞qitalic_q, in relation to the connection length. The importance of the parallel physics in determining the critical gradients has been recognised by many authors (Hahm & Tang, 1988; Jenko et al., 2001; Roberg-Clark et al., 2022a), and here we see is involved quite explicitly. Often parallel dynamics are associated to the global shear, which our model does not explicitly treat. However, as discussed with the Landau threshold, the dependence on ωtsubscript𝜔𝑡\omega_{t}italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT can be related to the role played by global shear when the modes become rather delocalised. The exact form of the scaling with q𝑞qitalic_q will depend on the exact behaviour of the mode that is becoming de-localised and thus should be taken with a grain of salt (especially given the weak spot of the model determining λ𝜆\lambdaitalic_λ). We shall not forget that although increasing the Landau threshold helps in this endeavour of erasing localised modes, one could of course leave behind de-localised modes that could also be deleterious (Zocco et al., 2018, 2022).

5 The role of higher harmonics

The kinetic considerations above have focused on the behaviour of a localised mode whose shape is in its simplest form described by a Gaussian envelope. That is, by the ‘pure’ 0-th order in our Taylor-Gauss expansion. Shapes of the modes are seldom so simple, and in the way that we expect multiple modes as solutions to a Schrödinger equation (and in fact also the fluid equation (Hahm & Tang, 1988)), we may also expect to find possible solutions to our problem in which the mode has larger n𝑛nitalic_n.

A dominant ‘pure’ M𝑀Mitalic_M mode can be described in a way rather analogous to that considered above for n=0𝑛0n=0italic_n = 0. It may be shown, see Appendix B, that the resulting equations are identical to the n=0𝑛0n=0italic_n = 0 case with the only modification,

ζn=M=12[λ(2M+1)(ωtω)2ω¯d2ω]1.subscript𝜁𝑛𝑀12superscriptdelimited-[]𝜆2𝑀1superscriptsubscript𝜔𝑡𝜔2subscript¯𝜔𝑑2𝜔1\zeta_{n=M}=\frac{1}{2}\left[\lambda(2M+1)\left(\frac{\omega_{t}}{\omega}% \right)^{2}-\frac{\bar{\omega}_{d}}{2\omega}\right]^{-1}.italic_ζ start_POSTSUBSCRIPT italic_n = italic_M end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG [ italic_λ ( 2 italic_M + 1 ) ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT . (45)

The only change is the different contribution of the streaming term, formally equivalent to an increased ωt(2M+1)ωtsubscript𝜔𝑡2𝑀1subscript𝜔𝑡\omega_{t}\rightarrow(2M+1)\omega_{t}italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT → ( 2 italic_M + 1 ) italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT. How can such a scaling be physically interpreted? Let us picture the change to the mode structure as one increases mode number. The mode looks like a pair of peaks pushed against the Gaussian envelope. The larger M𝑀Mitalic_M, the harder they are squeezed, which leads to the width of the peaks to roughly go like Δ¯1/Msimilar-toΔ¯1𝑀\Delta\bar{\ell}\sim 1/\sqrt{M}roman_Δ over¯ start_ARG roman_ℓ end_ARG ∼ 1 / square-root start_ARG italic_M end_ARG, as can be explicitly shown by computing the standard deviation. This linear scaling is what leads to the form in Eq. (45).

The main effect of M𝑀Mitalic_M is thus to change the parallel structure of the mode, and thus any of the phenomena that is directly linked to this feature of the instability. The Landau threshold will be most readily affected, but also the particulars of critical thresholds where ζ1similar-to𝜁1\zeta\sim 1italic_ζ ∼ 1 and |λ|𝜆|\lambda|| italic_λ | is not exceedingly large. With the effective ωtsubscript𝜔𝑡\omega_{t}italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT scaling, we may write (kαρi)Landau2M+1similar-tosubscriptsubscript𝑘𝛼subscript𝜌𝑖Landau2𝑀1(k_{\alpha}\rho_{i})_{\mathrm{Landau}}\sim 2M+1( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT roman_Landau end_POSTSUBSCRIPT ∼ 2 italic_M + 1, showing that Landau dam** becomes more prominent for the higher modes. This means that we expect to see the lowest modes excited first, as well as those to be most sensitive to stabilisation through squeezing of the connection length, Λcrit2M+1similar-tosubscriptΛcrit2𝑀1\Lambda_{\mathrm{crit}}\sim 2M+1roman_Λ start_POSTSUBSCRIPT roman_crit end_POSTSUBSCRIPT ∼ 2 italic_M + 1. The change in the ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT threshold with M𝑀Mitalic_M may appear shocking, given that this we argued is rather insensitive to parallel scales. However, as M𝑀Mitalic_M is increased, the ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT threshold mixes with the Landau threshold, as follows directly from ζ𝜁\zetaitalic_ζ, Eq. (20a). Similar behaviour to what we find here, Fig. 7, may be interpreted from the work of Gao et al. (2005). Thus we expect close to marginality the first localised mode to go unstable to be the n=0𝑛0n=0italic_n = 0 mode.

Refer to caption
Figure 7: Growth and frequency of ITG mode for different structure. Plots showing the growth rate and real frequency of the ITG mode for the mode numbers n=0,1,2𝑛012n=0,~{}1,~{}2italic_n = 0 , 1 , 2 using the approximate generalisation of ζ𝜁\zetaitalic_ζ. The plots are computed using Λ/Λ0=10ΛsubscriptΛ010\Lambda/\Lambda_{0}=10roman_Λ / roman_Λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 10, ω¯d/ωt=0.1subscript¯𝜔𝑑subscript𝜔𝑡0.1\bar{\omega}_{d}/\omega_{t}=0.1over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = 0.1, ωT/ωt=30superscriptsubscript𝜔𝑇subscript𝜔𝑡30\omega_{\star}^{T}/\omega_{t}=-30italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = - 30, ω=0subscript𝜔0\omega_{\star}=0italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = 0 and τ=1𝜏1\tau=1italic_τ = 1.

To study the changes on the magnitude of the growth rate, we write the M𝑀Mitalic_M-th mode generalisation of Eq. (33), which can be shown to match other treatments of the fluid equation (Hahm & Tang, 1988; Plunk et al., 2014). In this fluid case, it is clear that increasing mode number enhances the growth rate of our ITG, which is increasingly of a more slab character, γslab(2M+1)2/5similar-tosubscript𝛾slabsuperscript2𝑀125\gamma_{\mathrm{slab}}\sim(2M+1)^{2/5}italic_γ start_POSTSUBSCRIPT roman_slab end_POSTSUBSCRIPT ∼ ( 2 italic_M + 1 ) start_POSTSUPERSCRIPT 2 / 5 end_POSTSUPERSCRIPT. The fluid picture is unbound! Only through the regularising role played by the kinetic effects is the hierarchy regulated. There is a competition then between larger modes tending towards larger growth rates (in the fluid limit), but also becoming more effectively stabilised by Landau dam**. The results of such competition are presented for some example parameters in Figure 7.

These modifications that occur in the model provide us with a flavour of the kind of changes that one would expect when our ‘pure’ mode assumption is relaxed and the true value of λ𝜆\lambdaitalic_λ is different. While the main qualitative features shall remain, we expect quantitative differences to exist. Overall scalings can be understood as in Figure 7, but other more sophisticated dependence of λ𝜆\lambdaitalic_λ on kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT could also lead to additional features in the linear spectrum. The model should help us distinguish between these as well.

6 Qualitative comparison to simulations

In this Section we consider as a way of example the linear spectrum of ITG modes in a realistic but nonetheless simple stellarator geometry. We use this as way of illustration of how the lessons learnt from our model can be applied in practice to understand behaviour in more complex situations, but also its limitations. The example presented is the linear ITG mode spectrum along a flux tube of the HSX (Anderson et al., 1995) stellarator, a quasisymmetric stellarator (Nührenberg & Zille, 1988; Boozer, 1983; Rodriguez et al., 2020) with helical symmetry. The latter means that |𝐁|𝐁|\mathbf{B}|| bold_B | has a direction of symmetry to prevent fast loss of trapped particles, which implies a particularly simple, quasi-periodic curvature along field-lines. This simplicity, together with a reduced global magnetic shear, makes this example suitable for the comparison. The electrostatic mode in the gyrokinetic simulations is driven by an ion temperature gradient a/LT=2.5𝑎subscript𝐿𝑇2.5a/L_{T}=2.5italic_a / italic_L start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 2.5, kee** the density flat and treating electrons adiabatically. The linear gyrokinetic simulations are conducted with the stella code (Barnes et al., 2019), of which the details are provided in the supplementary material. The resulting linear spectrum is presented in Figure 8, where the poloidal wavenumber and frequencies have been normalised following the notation in this paper. For comparison to this HSX simulation, an equivalent simulation has also been carried out with a modified geometry in which all field quantities are constant along the field line except the curvature, which is modelled like a single quadratic well (modelled with parameter ω¯dsubscript¯𝜔𝑑\bar{\omega}_{d}over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT corresponding to the value of bad curvature at the bottom of the central well, and ΛΛ\Lambdaroman_Λ as the distance from the centre of the well to the first zero curvature crossing point), truncated at a finite good curvature to prevent spurious modes as discussed later.

Refer to caption
Figure 8: Linear mode spectrum for HSX gyrokinetic simulations. The growth rates (left) and real frequencies (right) are shown for linear HSX gyrokinetic simulations with a/LT=2.5𝑎subscript𝐿𝑇2.5a/L_{T}=2.5italic_a / italic_L start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 2.5 and a/Ln=0𝑎subscript𝐿𝑛0a/L_{n}=0italic_a / italic_L start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 0, and adiabatic electrons. The dashed line corresponds to the simulations performed with the same gradients but a modified geometry in which the only spatial dependence in the problem is ω¯dsubscript¯𝜔𝑑\bar{\omega}_{d}over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, and this is modelled as a truncated quadratic well. The coloured shade show the variation in the later half of the simulation of the mode frequency, giving a sense of trust of the spectra (blue and red for the HSX and model geometries respectively). The vertical lines correspond to the predicted FLR stabilisation threshold and the ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT stabilising resonance.

The first noticeable conclusion from the comparison between these two simulations (solid and dashed lines in Fig. 8) is that the simplified geometry appears to capture the behaviour of the ITG mode exceptionally well. This strongly backs the approach and geometric approximations considered in this paper, and supports our view on the key role played by the curvature in localising the ITG mode. Thus, we are allowed to describe realistic linear stellarator spectra from our theoretical perspective, even though we notice that the ‘pure’ mode assumption (and perhaps others such as a vigorous drive) does not provide a good quantitative description, in particular requiring somewhat larger temperature gradients in order to match magnitudes of mode frequency and growth rate of simulation results. We show an example of this lack of agreement in Figure 9, where we include the analytical model prediction for the parameters characterising HSX, and one with a 27%percent2727\%27 % larger temperature gradient.

Refer to caption
Figure 9: Comparison of simulation to analytic model showing quantitative discrepancies. The plot shows a comparison between the growth rate (top) and frequency (bottom) of the dominant linear ITG mode in the simulation of Figure 8 (solid line) and the analytic model developed in the paper. The dashed line corresponds to the model prediction for ω¯d/ωt=1.95subscript¯𝜔𝑑subscript𝜔𝑡1.95\bar{\omega}_{d}/\omega_{t}=1.95over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = 1.95 and ωT/ωt=18superscriptsubscript𝜔𝑇subscript𝜔𝑡18\omega_{\star}^{T}/\omega_{t}=-18italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = - 18, parameters obtained from the main well of the HSX geometry. The dot-dash line corresponds to the analytical prediction with a 27%percent2727\%27 % larger temperature gradient, with the shade representing ±5%plus-or-minuspercent5\pm 5\%± 5 % variation. This shows that the model suffers as a quantitative predictor. This suggests consideration of the model mainly as a physical qualitative framework to interpret linear spectra behaviour.

We can still investigate the key physical principles we have explored through our model to interpret the linear spectrum observed. Let us start from the rightmost part of the spectrum, where we expect to find our ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT threshold. In fact, and as shown in Fig. 8 and predicted, this stabilisation point does occur approximately at {ω}ω¯dsimilar-to𝜔subscript¯𝜔𝑑\Re\{\omega\}\sim-\bar{\omega}_{d}roman_ℜ { italic_ω } ∼ - over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT (numerically within 6%percent66\%6 %). Our interpretation of the nature of this threshold enables us to understand how the spectrum should change as the bad curvature of the field or the temperature gradient are varied. The spectrum would narrow as the curvature is increased or the gradients decreased. Such a simple perspective can explain the narrowing of spectra observed in recent efforts to optimise stellarators for improved temperature gradient thresholds (Roberg-Clark et al., 2022b). In the comparison between the full geometry simulation and that of the reduced geometry, we see that the latter seems to exhibit an additional unstable branch at smaller poloidal wavelengths. This branch corresponds to an anti-ionic temperature driven instability (i.e., rotating in the electronic direction, but not due to kinetic electrons, since their response is adiabatic) that exists in the region of good-curvature of the modified geometry. This behaviour is an interesting case study for the future: if anti-ionic modes localise in good curvature regions, attempting to stabilise the ITG mode by increasing the good curvature in the field could be problematic.

The region beyond the ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT threshold in the HSX case considered here does not exhibit other significant instabilities. However, less symmetric geometries such as those in quasi-isodynamic stellarators (Podavini et al., 2023), often exhibit linear spectra with additional structure beyond this point. How can this be framed within our description? That is: how can the ITG mode escape stabilisation by the ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT resonance? It is not the localisation of the mode that is suppressing the mode, as it may occur at small kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, thus delocalising the mode is not a solution.101010Close to marginality, where the kinetic difference between the ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT and Landau dam** as explored in previous sections becomes less definite, delocalisation may also become a relevant mechanism even on this side of the spectrum. The alternative left is for the mode to localise in a different well, a hop** mode. Moving to a well that is a priori less unstable because it has better curvature, or milder FLR effects, may however be beneficial for the mode because it may be sufficient to make {ω}ω¯d𝜔subscript¯𝜔𝑑\Re\{\omega\}\neq-\bar{\omega}_{d}roman_ℜ { italic_ω } ≠ - over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT and avoid stabilisation through the ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT resonance. Such a jump of the localised mode could occur more than once and we hypothesise, could lead to additional growth rate peaks in the spectrum. These changes on localisation are reminiscent of changes that occur to modes with finite kψsubscript𝑘𝜓k_{\psi}italic_k start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT (Parisi et al., 2022). The global behaviour of the localised mode including the possibility of living at different wells could be understood by constructing one spectrum like that in Fig. 8 for each well, with the corresponding change in the model parameters; more precisely, the magnitude of the drift, the well width and the FLR effect (with the latter strongly affecting the scale of kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT). The study of this is left for future work, but is an attractive way forward to understanding the behaviour of kinetic ion instabilities.

Our physics interpretation of the occurrence of the dip in the linear spectrum of Fig. 8 from FLR stabilisation and then weakening appears to be also correct. The dotted line in the figure corresponds in fact to the simple expression derived above in Eq. (38). This relative stabilisation of Larmor radius effects can also be observed in the response of the mode eigenfunction.

Refer to caption
Figure 10: Mode structure for HSX gyrokinetic simulations. The plot shows the parameters λrsubscript𝜆𝑟\lambda_{r}italic_λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT describing the localisation and oscillation of the modes in Fig. 8. λrsubscript𝜆𝑟\lambda_{r}italic_λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT is obtained by fitting a Gaussian exp[λr¯2/2]subscript𝜆𝑟superscript¯22\exp[-\lambda_{r}\bar{\ell}^{2}/2]roman_exp [ - italic_λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 ] to the absolute value of the electrostatic potential ϕitalic-ϕ\phiitalic_ϕ. The mode is not a pure exponential with generally wider tails. Three examples of the modes in the simplified geometry are provided in the right panel. The λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT parameter is obtained by reading the main oscillation frequency of {ϕ}/|ϕ|italic-ϕitalic-ϕ\Re\{\phi\}/|\phi|roman_ℜ { italic_ϕ } / | italic_ϕ |. Both these measures are inspired by the Gaussian basis used in this paper. The qualitative behaviour observed is fully consistent with the behaviour of our model. That is, λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT monotonically increases with the poloidal wavenumber while the localisation decreases near the Landau threshold (kαρi0similar-tosubscript𝑘𝛼subscript𝜌𝑖0k_{\alpha}\rho_{i}\sim 0italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∼ 0), the FLR stabilisation threshold and the drift resonance.

We present some example structures from the simulations in Figure 10. We use the structure from the stella simulations of the simplified geometry for better clarity. The figure also presents a measure of the localisation of the mode, λrsubscript𝜆𝑟\lambda_{r}italic_λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, and its structure, λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Numerically, we compute λrsubscript𝜆𝑟\lambda_{r}italic_λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT by fitting a Gaussian exp[λr¯2/2]subscript𝜆𝑟superscript¯22\exp[-\lambda_{r}\bar{\ell}^{2}/2]roman_exp [ - italic_λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 ] to |ϕ|italic-ϕ|\phi|| italic_ϕ |. Note that the wavefunctions are not really pure Gaussians; in fact, they appear to fit closer to a Gaussian in the centre and a weaker exponential decay further away. Of course the precise behaviour depends on the details of the geometry. Larger λrsubscript𝜆𝑟\lambda_{r}italic_λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT denotes a more localised mode. If one takes {ϕ}/|ϕ|italic-ϕitalic-ϕ\Re\{\phi\}/|\phi|roman_ℜ { italic_ϕ } / | italic_ϕ |, the mode then exhibits a clear periodicity (other than in the middle region). We define λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT to be the main frequency of the that oscillatory behaviour in ¯¯\bar{\ell}over¯ start_ARG roman_ℓ end_ARG. As it is clear in Fig. 10, as predicted by our treatment, as FLR stabilises the ITG, the mode becomes less localised, while kee** the scale of its mode structure (i.e., λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT unchanged). Note also that the behaviour near the ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT resonance gives us some intuition that the ‘pure’ mode form of λ𝜆\lambdaitalic_λ does not fully hold there, as the mode only weakly de-localises.

We are then left with the behaviour at the longest poloidal wavelength. As predicted by our theory, the mode becomes increasingly delocalised at lower kαρisubscript𝑘𝛼subscript𝜌𝑖k_{\alpha}\rho_{i}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. In terms of our Landau threshold picture, this occurs while Landau stabilisation becomes less and less effective. A causality relation between the lack of localisation and growth rate maximisation could perhaps be established via an energetic argument. Of course, at some point the model ceases to be valid, and the linear spectrum becomes dominated by completely delocalised modes. These very extended modes we may refer to as Floquet or slab-like modes (Zocco et al., 2018, 2022; Podavini et al., 2023). The difference in growth rate between the two gyrokinetic simulations in Fig. 8 at long wavelengths can be explained to be due to the preeminence of these modes. The presence of such elongated structures make the simulations rather challenging, as long flux tubes are necessary. Their behaviour lies outside the realm of the present model.

We thus understand the very distinct nature of the ITG mode for large and small wavelengths, where the kinetic stabilising mechanisms are different, and thus so is the mode response. This enables one to elucidate the meaning of the linear spectrum presented, and offers, as given, a way forward to interpreting linear spectra and their physical meaning in more complex geometries.

7 Conclusions

In this work, we have proposed a theory of the kinetic Ion-Temperature-Gradient driven mode which features resonant kinetic effects, with a localised mode structure induced by the field-line-dependent geometry. We focused on the localising action of the magnetic drift, allowing for general conclusions that apply both in a tokamak and stellarator context.

The magnetic drift spatial dependence models good and bad curvature regions, introduced with a local well quadratic model. The mathematical description of the problem is based on a power series expansion of the eigenfunctions in the field-following co-ordinate, mitigated by a Gaussian envelope. This generates a hierarchy of coupled eigenvalue problems which for small magnetic drift frequency, strong drive and localised modes, can be truncated. A relatively simple dispersion relation is constructed by considering the simplifying assumption of a singly dominant mode, which we refer to as ‘pure’. The resulting description features long-wavelength Landau dam**, arbitrary Larmor radius effects, and a regularising resonant action of the magnetic drift for short wavelength modes. All these salient physical features are demonstrated to be numerically observed in realistic stellarator geometry, although the simplicity of the model limits quantitative comparisons with numerical results. Venues for improving the model are also proposed for future work.

The model is also used to provide a prediction for the resonant stabilisation of the toroidal branch of the ITG mode without tacitly assuming constant (along the field line, and thus unrealistic) eigenfunctions. The result is insight into how to tailor the field-line dependence of the magnetic drift in order to suppress the ITG instability. By enforcing that all scales are either Landau damped (at long wavelengths) or kinetically suppressed by the magnetic drift resonance (at short wavelengths) one obtains a critical gradient for ITG destabilisation that scales with a/LT1/(q0.6R).proportional-to𝑎subscript𝐿𝑇1superscript𝑞0.6𝑅a/L_{T}\propto 1/(q^{0.6}R).italic_a / italic_L start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ∝ 1 / ( italic_q start_POSTSUPERSCRIPT 0.6 end_POSTSUPERSCRIPT italic_R ) . This explains why large inverse aspect-ratio devices feature small critical thresholds (as is well known), but also indicates a beneficial effects in having a small safety factor q𝑞qitalic_q, or more generally short connection length. This final aspect is of particular interest, since it is synergistic with the field-line-bending stabilisation of magnetohydrodynamic (MHD) instabilities (Bernstein et al., 1958; Mercier, 1962; Greene & Johnson, 1962; Correa-Restrepo, 1978; Connor et al., 1978). These expectations, scalings, synergies and behaviour alongside MHD stability will be the subject of careful analysis and simulation in a future paper (Rodriguez & Zocco, 2024).

Data availability

The data that support the findings of this study are openly available at the Zenodo repository with DOI/URL 10.5281/zenodo.11388974.

Acknowledgements

We gratefully acknowledge fruitful discussion with R. Nies, P. Costello, G. Plunk, G. Roberg-Clark, L. Podavini and F. Parra.

Funding

E. R. was supported by a grant by Alexander-von-Humboldt-Stiftung, Bonn, Germany, through a postdoctoral research fellowship. Part of this work was conceived during the Simons Collaboration on Hidden Symmetries meetings, to which we are grateful for its support.

Declaration of interest

The authors report no conflict of interest.

Appendix A Taylor-Gauss expansion of GK equation

In this Appendix we detail the Taylor-Gauss expansion of the GK equation. Our starting point is the GK equation, Eq. (9), which we write as,

2(ωtxω)2¯2g+(1ω~dω(¯21))g+4ω~d(ωtxω)2¯¯g=qiTiF0iJ0(1ω~ω)ϕ,2superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2superscriptsubscript¯2𝑔1subscript~𝜔𝑑𝜔superscript¯21𝑔4subscript~𝜔𝑑superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2¯subscript¯𝑔subscript𝑞𝑖subscript𝑇𝑖subscript𝐹0𝑖subscript𝐽01subscript~𝜔𝜔italic-ϕ2\left(\frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}\partial_{\bar{\ell}}^% {2}g+\left(1-\frac{\tilde{\omega}_{d}}{\omega}(\bar{\ell}^{2}-1)\right)g+4% \tilde{\omega}_{d}\left(\frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}\bar{% \ell}\partial_{\bar{\ell}}g=\frac{q_{i}}{T_{i}}F_{0i}J_{0}\left(1-\frac{\tilde% {\omega}_{\star}}{\omega}\right)\phi,2 ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∂ start_POSTSUBSCRIPT over¯ start_ARG roman_ℓ end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_g + ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ( over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) ) italic_g + 4 over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over¯ start_ARG roman_ℓ end_ARG ∂ start_POSTSUBSCRIPT over¯ start_ARG roman_ℓ end_ARG end_POSTSUBSCRIPT italic_g = divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) italic_ϕ , (9)

To obtain our Taylor-Gauss resolution of the equation we then substitute,

g(¯,𝐯)=n=0gn(𝐯){λ}nn!¯neλ¯2/2,g(\bar{\ell},\mathbf{v})=\sum_{n=0}^{\infty}g_{n}(\mathbf{v})\sqrt{\frac{\Re\{% \lambda\}^{n}}{n!}}\bar{\ell}^{n}e^{-\lambda\bar{\ell}^{2}/2},italic_g ( over¯ start_ARG roman_ℓ end_ARG , bold_v ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_v ) square-root start_ARG divide start_ARG roman_ℜ { italic_λ } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG end_ARG over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_λ over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 end_POSTSUPERSCRIPT , (11)

and likewise for ϕitalic-ϕ\phiitalic_ϕ, into the equation. Note that the normalisation is chosen here in such a way that the magnitude of the basis (i.e., the terms multiplying gnsubscript𝑔𝑛g_{n}italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT) are considered roughly of order one.

The action of the second derivative on this basis yields,

¯2g(¯,𝐯)=superscriptsubscript¯2𝑔¯𝐯absent\displaystyle\partial_{\bar{\ell}}^{2}g(\bar{\ell},\mathbf{v})=∂ start_POSTSUBSCRIPT over¯ start_ARG roman_ℓ end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_g ( over¯ start_ARG roman_ℓ end_ARG , bold_v ) = n=0gneλ¯2/2{λ}nn![n(n1)¯n2(2n+1)λ¯n+λ2¯n+2]\displaystyle\sum_{n=0}^{\infty}g_{n}e^{-\lambda\bar{\ell}^{2}/2}\sqrt{\frac{% \Re\{\lambda\}^{n}}{n!}}\left[n(n-1)\bar{\ell}^{n-2}-(2n+1)\lambda\bar{\ell}^{% n}+\lambda^{2}\bar{\ell}^{n+2}\right]∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_λ over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 end_POSTSUPERSCRIPT square-root start_ARG divide start_ARG roman_ℜ { italic_λ } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG end_ARG [ italic_n ( italic_n - 1 ) over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT italic_n - 2 end_POSTSUPERSCRIPT - ( 2 italic_n + 1 ) italic_λ over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT + italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT italic_n + 2 end_POSTSUPERSCRIPT ]
=\displaystyle== n=0λ¯nN(n)eλ¯2/2[(n+2)(n+1){λ}λgn+2(2n+1)gn+n(n1)λ{λ}gn2].superscriptsubscript𝑛0𝜆superscript¯𝑛𝑁𝑛superscript𝑒𝜆superscript¯22delimited-[]𝑛2𝑛1𝜆𝜆subscript𝑔𝑛22𝑛1subscript𝑔𝑛𝑛𝑛1𝜆𝜆subscript𝑔𝑛2\displaystyle\sum_{n=0}^{\infty}\frac{\lambda\bar{\ell}^{n}}{N(n)}e^{-\lambda% \bar{\ell}^{2}/2}\left[\sqrt{(n+2)(n+1)}\frac{\Re\{\lambda\}}{\lambda}g_{n+2}-% (2n+1)g_{n}+\sqrt{n(n-1)}\frac{\lambda}{\Re\{\lambda\}}g_{n-2}\right].∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_λ over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_N ( italic_n ) end_ARG italic_e start_POSTSUPERSCRIPT - italic_λ over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 end_POSTSUPERSCRIPT [ square-root start_ARG ( italic_n + 2 ) ( italic_n + 1 ) end_ARG divide start_ARG roman_ℜ { italic_λ } end_ARG start_ARG italic_λ end_ARG italic_g start_POSTSUBSCRIPT italic_n + 2 end_POSTSUBSCRIPT - ( 2 italic_n + 1 ) italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + square-root start_ARG italic_n ( italic_n - 1 ) end_ARG divide start_ARG italic_λ end_ARG start_ARG roman_ℜ { italic_λ } end_ARG italic_g start_POSTSUBSCRIPT italic_n - 2 end_POSTSUBSCRIPT ] . (46)

In this notation it should be interpreted that gn=0subscript𝑔𝑛0g_{n}=0italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 0 for n<0𝑛0n<0italic_n < 0, and likewise any negative power of ¯¯\bar{\ell}over¯ start_ARG roman_ℓ end_ARG in the summation should be taken to be zero.

For the terms that present a product with the drift frequency, and thus with a second power of ¯2superscript¯2\bar{\ell}^{2}over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, we simply have,

¯2g(¯,𝐯)=n=0¯neλ¯2/2{λ}nn!1{λ}n(n+1)gn2,\bar{\ell}^{2}g(\bar{\ell},\mathbf{v})=\sum_{n=0}^{\infty}\bar{\ell}^{n}e^{-% \lambda\bar{\ell}^{2}/2}\sqrt{\frac{\Re\{\lambda\}^{n}}{n!}}\frac{1}{\Re\{% \lambda\}}\sqrt{n(n+1)}g_{n-2},over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_g ( over¯ start_ARG roman_ℓ end_ARG , bold_v ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_λ over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 end_POSTSUPERSCRIPT square-root start_ARG divide start_ARG roman_ℜ { italic_λ } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG end_ARG divide start_ARG 1 end_ARG start_ARG roman_ℜ { italic_λ } end_ARG square-root start_ARG italic_n ( italic_n + 1 ) end_ARG italic_g start_POSTSUBSCRIPT italic_n - 2 end_POSTSUBSCRIPT , (47)

and,

¯¯g=n=0¯neλ¯2/2{λ}nn!(ngnλ{λ}n(n1)gn2).\bar{\ell}\partial_{\bar{\ell}}g=\sum_{n=0}^{\infty}\bar{\ell}^{n}e^{-\lambda% \bar{\ell}^{2}/2}\sqrt{\frac{\Re\{\lambda\}^{n}}{n!}}\left(ng_{n}-\frac{% \lambda}{\Re\{\lambda\}}\sqrt{n(n-1)}g_{n-2}\right).over¯ start_ARG roman_ℓ end_ARG ∂ start_POSTSUBSCRIPT over¯ start_ARG roman_ℓ end_ARG end_POSTSUBSCRIPT italic_g = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_λ over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 end_POSTSUPERSCRIPT square-root start_ARG divide start_ARG roman_ℜ { italic_λ } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG end_ARG ( italic_n italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - divide start_ARG italic_λ end_ARG start_ARG roman_ℜ { italic_λ } end_ARG square-root start_ARG italic_n ( italic_n - 1 ) end_ARG italic_g start_POSTSUBSCRIPT italic_n - 2 end_POSTSUBSCRIPT ) . (48)

Similar expressions would be obtained if a Hermite basis was used instead. With these expressions and collecting terms, we get the general equation

En=2{λ}(ωtxω)2(n+2)(n+1)gn+2++[1+ω~dω2λ(2n+1)(ωtxω)2+4nω~dω(ωtxω)2]gn++2{λ}[λ2(ωtxω)2ω~d2ω2λω~dω(ωtxω)2]n(n1)gn2qiTiF0iJ0(1ω~ω)ϕn.subscript𝐸𝑛2𝜆superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2𝑛2𝑛1subscript𝑔𝑛2delimited-[]1subscript~𝜔𝑑𝜔2𝜆2𝑛1superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔24𝑛subscript~𝜔𝑑𝜔superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2subscript𝑔𝑛2𝜆delimited-[]superscript𝜆2superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2subscript~𝜔𝑑2𝜔2𝜆subscript~𝜔𝑑𝜔superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2𝑛𝑛1subscript𝑔𝑛2subscript𝑞𝑖subscript𝑇𝑖subscript𝐹0𝑖subscript𝐽01subscript~𝜔𝜔subscriptitalic-ϕ𝑛E_{n}=2\Re\{\lambda\}\left(\frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}% \sqrt{(n+2)(n+1)}g_{n+2}+\\ +\left[1+\frac{\tilde{\omega}_{d}}{\omega}-2\lambda(2n+1)\left(\frac{\omega_{t% }x_{\parallel}}{\omega}\right)^{2}+4n\frac{\tilde{\omega}_{d}}{\omega}\left(% \frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}\right]g_{n}+\\ +\frac{2}{\Re\{\lambda\}}\left[\lambda^{2}\left(\frac{\omega_{t}x_{\parallel}}% {\omega}\right)^{2}-\frac{\tilde{\omega}_{d}}{2\omega}-2\lambda\frac{\tilde{% \omega}_{d}}{\omega}\left(\frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}% \right]\sqrt{n(n-1)}g_{n-2}-\\ -\frac{q_{i}}{T_{i}}F_{0i}J_{0}\left(1-\frac{\tilde{\omega}_{\star}}{\omega}% \right)\phi_{n}.start_ROW start_CELL italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 2 roman_ℜ { italic_λ } ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT square-root start_ARG ( italic_n + 2 ) ( italic_n + 1 ) end_ARG italic_g start_POSTSUBSCRIPT italic_n + 2 end_POSTSUBSCRIPT + end_CELL end_ROW start_ROW start_CELL + [ 1 + divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG - 2 italic_λ ( 2 italic_n + 1 ) ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 italic_n divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + end_CELL end_ROW start_ROW start_CELL + divide start_ARG 2 end_ARG start_ARG roman_ℜ { italic_λ } end_ARG [ italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG - 2 italic_λ divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] square-root start_ARG italic_n ( italic_n - 1 ) end_ARG italic_g start_POSTSUBSCRIPT italic_n - 2 end_POSTSUBSCRIPT - end_CELL end_ROW start_ROW start_CELL - divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT . end_CELL end_ROW (48)

presented in the main text, and to be interpreted as,

n=0En¯neλ¯2/2{λ}nn!=0.\sum_{n=0}^{\infty}E_{n}\bar{\ell}^{n}e^{-\lambda\bar{\ell}^{2}/2}\sqrt{\frac{% \Re\{\lambda\}^{n}}{n!}}=0.∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_λ over¯ start_ARG roman_ℓ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 end_POSTSUPERSCRIPT square-root start_ARG divide start_ARG roman_ℜ { italic_λ } start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG end_ARG = 0 . (49)

It is clear that for an exact solution to the system we need to satisfy {En=0}subscript𝐸𝑛0\{E_{n}=0\}{ italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 0 } for all n𝑛nitalic_n, if the equation is to be satisfied for all ¯¯\bar{\ell}over¯ start_ARG roman_ℓ end_ARG.

Appendix B Details on dispersion construction and higher order modes

In this Appendix we show the details on how the construction of the dispersion function in the text is done, including the generalisation to other modes other than the simplest Gaussian n=0𝑛0n=0italic_n = 0.

B.1 General structure of the problem

To obtain the dispersion relation of whichever mode we are interested in investigating, it is necessary to construct the matrix 𝔻𝔻\mathbb{D}blackboard_D in Eq. (13). That is, we must make the appropriate combinations of the coefficients in Eq. (12b) to bring the set of equations to the form considered in Eq. (13). Because we are dealing with a system of equations with dimension N𝑁Nitalic_N, our truncation number, we would like some systematic way in which to construct the relevant entries in the matrix. In particular, we would like to exploit our ordering in δ𝛿\deltaitalic_δ and ϵitalic-ϵ\epsilonitalic_ϵ to simplify the procedure before explicitly constructing the equations.

To that end, let us write the system of equations defined by the first N/2+1𝑁21N/2+1italic_N / 2 + 1 equations in matrix form,

\mathsfbiGij𝐠j=𝚽ijϕj,\mathsfbisubscript𝐺𝑖𝑗subscript𝐠𝑗subscript𝚽𝑖𝑗subscriptbold-italic-ϕ𝑗\mathsfbi{G}_{ij}\mathbf{g}_{j}=\boldsymbol{\Phi}_{ij}\boldsymbol{\phi}_{j},italic_G start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT bold_g start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = bold_Φ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT bold_italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (50)

where the vectors 𝐠𝐠\mathbf{g}bold_g and ϕbold-italic-ϕ\boldsymbol{\phi}bold_italic_ϕ contain the modes from n=0𝑛0n=0italic_n = 0 to N𝑁Nitalic_N, and the matrices can be constructed following the general expression for Ensubscript𝐸𝑛E_{n}italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT in Eq. (12b). Because the system considers separately the even and odd orders, we shall restrict these matrices to the even or odd parts of the problem separately. The arguments to follow are independent of which class we consider (with minor changes), but we choose the even part of the solution (just because one must be chosen).

Because we are about to invoke some ordering considerations to simplify the system of equations, we shall fairly represent the order of each mode. We shall treat first each gnsubscript𝑔𝑛g_{n}italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT in an unordered fashion, and thus only consider the ordering of the factors we have explicitly in Eq. (12b). For simplicity, and in hindsight of what will end up later being considered, we shall take the following consistent scaling for λδ(ω/ωt)similar-to𝜆𝛿𝜔subscript𝜔𝑡\lambda\sim\sqrt{\delta}(\omega/\omega_{t})italic_λ ∼ square-root start_ARG italic_δ end_ARG ( italic_ω / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ). By itself it is not ordered in any particular way (it simply cannot be too small), but λ(ωt/ω)2O(ϵ)similar-to𝜆superscriptsubscript𝜔𝑡𝜔2𝑂italic-ϵ\lambda(\omega_{t}/\omega)^{2}\sim O(\epsilon)italic_λ ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∼ italic_O ( italic_ϵ ) and λ2(ωt/ω)2O(δ)similar-tosuperscript𝜆2superscriptsubscript𝜔𝑡𝜔2𝑂𝛿\lambda^{2}(\omega_{t}/\omega)^{2}\sim O(\delta)italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∼ italic_O ( italic_δ ). With this in mind it is straightforward to picture the ordering of each element in the matrices,

\mathsfbiG=(1ϵϵ1ϵϵ1ϵϵ1ϵϵ1),𝚽=(11111).formulae-sequence\mathsfbi𝐺matrix1italic-ϵmissing-subexpressionmissing-subexpressionmissing-subexpressionitalic-ϵ1italic-ϵmissing-subexpressionmissing-subexpressionmissing-subexpressionitalic-ϵ1italic-ϵmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionitalic-ϵ1italic-ϵmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionitalic-ϵ1𝚽matrix1missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression1missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression1missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression1missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression1\mathsfbi{G}=\begin{pmatrix}1&\epsilon&&&\\ \epsilon&1&\epsilon&&\\ &\epsilon&1&\epsilon&\\ &&\ddots&\ddots&\ddots\\ &&&\epsilon&1&\epsilon\\ &&&&\epsilon&1\\ \end{pmatrix},~{}~{}~{}~{}\boldsymbol{\Phi}=\begin{pmatrix}1&&&&\\ &1&&&\\ &&1&&\\ &&&\ddots&\\ &&&&1&\\ &&&&&1\\ \end{pmatrix}.italic_G = ( start_ARG start_ROW start_CELL 1 end_CELL start_CELL italic_ϵ end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_ϵ end_CELL start_CELL 1 end_CELL start_CELL italic_ϵ end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL italic_ϵ end_CELL start_CELL 1 end_CELL start_CELL italic_ϵ end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL italic_ϵ end_CELL start_CELL 1 end_CELL start_CELL italic_ϵ end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL italic_ϵ end_CELL start_CELL 1 end_CELL end_ROW end_ARG ) , bold_Φ = ( start_ARG start_ROW start_CELL 1 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL 1 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL 1 end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL ⋱ end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 1 end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 1 end_CELL end_ROW end_ARG ) . (51)

We note that \mathsfbiG\mathsfbi𝐺\mathsfbi{G}italic_G is a tridiagonal matrix while 𝚽𝚽\boldsymbol{\Phi}bold_Φ only has non-zero entries in the main and lower diagonals. And importantly, hop** off-diagonal terms are ordered like ϵitalic-ϵ\epsilonitalic_ϵ, which shall allow us to simplify the problem significantly. To order these matrix elements the way we have done, we must note that the hop** terms as considered do not only bring ϵitalic-ϵ\epsilonitalic_ϵ, but are also proportional to the mode number n𝑛nitalic_n. Thus, to preserve the ordering ϵitalic-ϵ\epsilonitalic_ϵ of the off-diagonals we shall limit the truncation NNϵ=1/ϵmuch-less-than𝑁subscript𝑁italic-ϵ1italic-ϵN\ll N_{\epsilon}=1/\epsilonitalic_N ≪ italic_N start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT = 1 / italic_ϵ.

With these matrices set up, we may now attempt the approximate construction of 𝔻=\mathsfbiG1𝚽𝔻\mathsfbisuperscript𝐺1𝚽\mathbb{D}=\mathsfbi{G}^{-1}\boldsymbol{\Phi}blackboard_D = italic_G start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_Φ to the right order O(ϵ2)𝑂superscriptitalic-ϵ2O(\epsilon^{2})italic_O ( italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). Fortunately, the inversion of a tridiagonal matrix can be expressed succinctly in the following form by Usmani (1994). Defining the elements along the main diagonal as aisubscript𝑎𝑖a_{i}italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT from i=1,,N/2𝑖1𝑁2i=1,\dots,N/2italic_i = 1 , … , italic_N / 2,111111If we were considering the truncated system for the odd parts, then we would end at (N+1)/2𝑁12(N+1)/2( italic_N + 1 ) / 2. and the upper and lower diagonals as bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and cisubscript𝑐𝑖c_{i}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT respectively (also starting at n=1𝑛1n=1italic_n = 1) for matrix \mathsfbiG\mathsfbi𝐺\mathsfbi{G}italic_G, the inverse may be written as,

(\mathsfbiG1)ij={(1)i+jbibj1θi1ϕj+1/θn(ij)(1)i+jcjci1θj1ϕi+1/θn(i>j),subscript\mathsfbisuperscript𝐺1𝑖𝑗casessuperscript1𝑖𝑗subscript𝑏𝑖subscript𝑏𝑗1subscript𝜃𝑖1subscriptitalic-ϕ𝑗1subscript𝜃𝑛𝑖𝑗superscript1𝑖𝑗subscript𝑐𝑗subscript𝑐𝑖1subscript𝜃𝑗1subscriptitalic-ϕ𝑖1subscript𝜃𝑛𝑖𝑗(\mathsfbi{G}^{-1})_{ij}=\begin{cases}(-1)^{i+j}b_{i}\dots b_{j-1}\theta_{i-1}% \phi_{j+1}/\theta_{n}&(i\leq j)\\ (-1)^{i+j}c_{j}\dots c_{i-1}\theta_{j-1}\phi_{i+1}/\theta_{n}&(i>j),\end{cases}( italic_G start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = { start_ROW start_CELL ( - 1 ) start_POSTSUPERSCRIPT italic_i + italic_j end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT … italic_b start_POSTSUBSCRIPT italic_j - 1 end_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT / italic_θ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_CELL start_CELL ( italic_i ≤ italic_j ) end_CELL end_ROW start_ROW start_CELL ( - 1 ) start_POSTSUPERSCRIPT italic_i + italic_j end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT … italic_c start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT italic_j - 1 end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT / italic_θ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_CELL start_CELL ( italic_i > italic_j ) , end_CELL end_ROW (52)

where θi=aiθi1bi1ci1θi1subscript𝜃𝑖subscript𝑎𝑖subscript𝜃𝑖1subscript𝑏𝑖1subscript𝑐𝑖1subscript𝜃𝑖1\theta_{i}=a_{i}\theta_{i-1}-b_{i-1}c_{i-1}\theta_{i-1}italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT - italic_b start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT with θ0=1subscript𝜃01\theta_{0}=1italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 and θ1=a1subscript𝜃1subscript𝑎1\theta_{1}=a_{1}italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and ϕi=aiϕi+1biciϕi+2subscriptitalic-ϕ𝑖subscript𝑎𝑖subscriptitalic-ϕ𝑖1subscript𝑏𝑖subscript𝑐𝑖subscriptitalic-ϕ𝑖2\phi_{i}=a_{i}\phi_{i+1}-b_{i}c_{i}\phi_{i+2}italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT - italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_i + 2 end_POSTSUBSCRIPT with ϕN/2+1=1subscriptitalic-ϕ𝑁211\phi_{N/2+1}=1italic_ϕ start_POSTSUBSCRIPT italic_N / 2 + 1 end_POSTSUBSCRIPT = 1 and ϕN/2=aN/2subscriptitalic-ϕ𝑁2subscript𝑎𝑁2\phi_{N/2}=a_{N/2}italic_ϕ start_POSTSUBSCRIPT italic_N / 2 end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT italic_N / 2 end_POSTSUBSCRIPT. Now, recall we are interested in the construction to order O(ϵ)𝑂italic-ϵO(\epsilon)italic_O ( italic_ϵ ), and thus, we may drop any term that is higher order. In particular, this implies drop** in the iteration expressions the terms involving products of b𝑏bitalic_b and c𝑐citalic_c, which as we indicated above, are each order ϵitalic-ϵ\epsilonitalic_ϵ. As a result, θi=aia1subscript𝜃𝑖subscript𝑎𝑖subscript𝑎1\theta_{i}=a_{i}\dots a_{1}italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT … italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and ϕi=aN/2aisubscriptitalic-ϕ𝑖subscript𝑎𝑁2subscript𝑎𝑖\phi_{i}=a_{N/2}\dots a_{i}italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT italic_N / 2 end_POSTSUBSCRIPT … italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Finally, restricting the products of b𝑏bitalic_b and c𝑐citalic_c involved in Eq. (52) not to surpass the right order, we may write the inverse succinctly as follows,

(\mathsfbiG1)ij=1aiδijbiaiai+1δi,j1ci1aiai1δi,j+1.subscript\mathsfbisuperscript𝐺1𝑖𝑗1subscript𝑎𝑖subscript𝛿𝑖𝑗subscript𝑏𝑖subscript𝑎𝑖subscript𝑎𝑖1subscript𝛿𝑖𝑗1subscript𝑐𝑖1subscript𝑎𝑖subscript𝑎𝑖1subscript𝛿𝑖𝑗1(\mathsfbi{G}^{-1})_{ij}=\frac{1}{a_{i}}\delta_{ij}-\frac{b_{i}}{a_{i}a_{i+1}}% \delta_{i,j-1}-\frac{c_{i-1}}{a_{i}a_{i-1}}\delta_{i,j+1}.( italic_G start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT - divide start_ARG italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT end_ARG italic_δ start_POSTSUBSCRIPT italic_i , italic_j - 1 end_POSTSUBSCRIPT - divide start_ARG italic_c start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT end_ARG italic_δ start_POSTSUBSCRIPT italic_i , italic_j + 1 end_POSTSUBSCRIPT . (53)

Define now the diagonal elements of 𝚽ij=pjδijsubscript𝚽𝑖𝑗subscript𝑝𝑗subscript𝛿𝑖𝑗\boldsymbol{\Phi}_{ij}=p_{j}\delta_{ij}bold_Φ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT; the matrix product is then, to leading order,

𝔻2i,2j=(\mathsfbiG1𝚽)ij=piaiδijcjpjaiajδi1,jbipjaiajδi+1,j.subscript𝔻2𝑖2𝑗subscript\mathsfbisuperscript𝐺1𝚽𝑖𝑗subscript𝑝𝑖subscript𝑎𝑖subscript𝛿𝑖𝑗subscript𝑐𝑗subscript𝑝𝑗subscript𝑎𝑖subscript𝑎𝑗subscript𝛿𝑖1𝑗subscript𝑏𝑖subscript𝑝𝑗subscript𝑎𝑖subscript𝑎𝑗subscript𝛿𝑖1𝑗\mathbb{D}_{2i,2j}=\left(\mathsfbi{G}^{-1}\boldsymbol{\Phi}\right)_{ij}=\frac{% p_{i}}{a_{i}}\delta_{ij}-\frac{c_{j}p_{j}}{a_{i}a_{j}}\delta_{i-1,j}-\frac{b_{% i}p_{j}}{a_{i}a_{j}}\delta_{i+1,j}.blackboard_D start_POSTSUBSCRIPT 2 italic_i , 2 italic_j end_POSTSUBSCRIPT = ( italic_G start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_Φ ) start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = divide start_ARG italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT - divide start_ARG italic_c start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG italic_δ start_POSTSUBSCRIPT italic_i - 1 , italic_j end_POSTSUBSCRIPT - divide start_ARG italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG italic_δ start_POSTSUBSCRIPT italic_i + 1 , italic_j end_POSTSUBSCRIPT . (54)

To further simplify, let us be more explicit on the various coefficients of \mathsfbiG\mathsfbi𝐺\mathsfbi{G}italic_G and 𝚽𝚽\boldsymbol{\Phi}bold_Φ. Reading the expressions off Eq. (12b) to the correct order,

pn/2subscript𝑝𝑛2\displaystyle p_{n/2}italic_p start_POSTSUBSCRIPT italic_n / 2 end_POSTSUBSCRIPT =qiTiF0iJ0(1ω~ω),absentsubscript𝑞𝑖subscript𝑇𝑖subscript𝐹0𝑖subscript𝐽01subscript~𝜔𝜔\displaystyle=\frac{q_{i}}{T_{i}}F_{0i}J_{0}\left(1-\frac{\tilde{\omega}_{% \star}}{\omega}\right),= divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) , (55a)
an/2subscript𝑎𝑛2\displaystyle a_{n/2}italic_a start_POSTSUBSCRIPT italic_n / 2 end_POSTSUBSCRIPT 1+ω~dω2λ(2n+1)(ωtxω)2,absent1subscript~𝜔𝑑𝜔2𝜆2𝑛1superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2\displaystyle\approx 1+\frac{\tilde{\omega}_{d}}{\omega}-2\lambda(2n+1)\left(% \frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2},≈ 1 + divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG - 2 italic_λ ( 2 italic_n + 1 ) ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (55b)
bn/2subscript𝑏𝑛2\displaystyle b_{n/2}italic_b start_POSTSUBSCRIPT italic_n / 2 end_POSTSUBSCRIPT =2{λ}(ωtxω)2(n+2)(n+1),absent2𝜆superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2𝑛2𝑛1\displaystyle=2\Re\{\lambda\}\left(\frac{\omega_{t}x_{\parallel}}{\omega}% \right)^{2}\sqrt{(n+2)(n+1)},= 2 roman_ℜ { italic_λ } ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT square-root start_ARG ( italic_n + 2 ) ( italic_n + 1 ) end_ARG , (55c)
cn/21subscript𝑐𝑛21\displaystyle c_{n/2-1}italic_c start_POSTSUBSCRIPT italic_n / 2 - 1 end_POSTSUBSCRIPT =2{λ}n(n1)[λ2(ωtxω)2ω~d2ω].absent2𝜆𝑛𝑛1delimited-[]superscript𝜆2superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2subscript~𝜔𝑑2𝜔\displaystyle=\frac{2}{\Re\{\lambda\}}\sqrt{n(n-1)}\left[\lambda^{2}\left(% \frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}-\frac{\tilde{\omega}_{d}}{2% \omega}\right].= divide start_ARG 2 end_ARG start_ARG roman_ℜ { italic_λ } end_ARG square-root start_ARG italic_n ( italic_n - 1 ) end_ARG [ italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG ] . (55d)

With these, we may write, at once,

𝔻nnsubscript𝔻𝑛𝑛\displaystyle\mathbb{D}_{nn}blackboard_D start_POSTSUBSCRIPT italic_n italic_n end_POSTSUBSCRIPT =qiTiF0iJ0(1ω~ω)11+ω~dω2λ(2n+1)(ωtxω)2,absentsubscript𝑞𝑖subscript𝑇𝑖subscript𝐹0𝑖subscript𝐽01subscript~𝜔𝜔11subscript~𝜔𝑑𝜔2𝜆2𝑛1superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2\displaystyle=\frac{q_{i}}{T_{i}}F_{0i}J_{0}\left(1-\frac{\tilde{\omega}_{% \star}}{\omega}\right)\frac{1}{1+\frac{\tilde{\omega}_{d}}{\omega}-2\lambda(2n% +1)\left(\frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}},= divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) divide start_ARG 1 end_ARG start_ARG 1 + divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG - 2 italic_λ ( 2 italic_n + 1 ) ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (56a)
𝔻n,n2subscript𝔻𝑛𝑛2\displaystyle\mathbb{D}_{n,n-2}blackboard_D start_POSTSUBSCRIPT italic_n , italic_n - 2 end_POSTSUBSCRIPT =qiTiF0iJ0(1ω~ω)2{λ}n(n1)[λ2(ωtxω)2ω~d2ω],absentsubscript𝑞𝑖subscript𝑇𝑖subscript𝐹0𝑖subscript𝐽01subscript~𝜔𝜔2𝜆𝑛𝑛1delimited-[]superscript𝜆2superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2subscript~𝜔𝑑2𝜔\displaystyle=-\frac{q_{i}}{T_{i}}F_{0i}J_{0}\left(1-\frac{\tilde{\omega}_{% \star}}{\omega}\right)\frac{2}{\Re\{\lambda\}}\sqrt{n(n-1)}\left[\lambda^{2}% \left(\frac{\omega_{t}x_{\parallel}}{\omega}\right)^{2}-\frac{\tilde{\omega}_{% d}}{2\omega}\right],= - divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) divide start_ARG 2 end_ARG start_ARG roman_ℜ { italic_λ } end_ARG square-root start_ARG italic_n ( italic_n - 1 ) end_ARG [ italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG ] , (56b)
𝔻n,n+2subscript𝔻𝑛𝑛2\displaystyle\mathbb{D}_{n,n+2}blackboard_D start_POSTSUBSCRIPT italic_n , italic_n + 2 end_POSTSUBSCRIPT =qiTiF0iJ0(1ω~ω)2{λ}(n+1)(n+2)(ωtxω)2,absentsubscript𝑞𝑖subscript𝑇𝑖subscript𝐹0𝑖subscript𝐽01subscript~𝜔𝜔2𝜆𝑛1𝑛2superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2\displaystyle=-\frac{q_{i}}{T_{i}}F_{0i}J_{0}\left(1-\frac{\tilde{\omega}_{% \star}}{\omega}\right)2\Re\{\lambda\}\sqrt{(n+1)(n+2)}\left(\frac{\omega_{t}x_% {\parallel}}{\omega}\right)^{2},= - divide start_ARG italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) 2 roman_ℜ { italic_λ } square-root start_ARG ( italic_n + 1 ) ( italic_n + 2 ) end_ARG ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (56c)

as the only relevant matrix elements to order ϵitalic-ϵ\epsilonitalic_ϵ.

With this matrix in place, we are in a position to apply quasineutrality to construct the system of equations on ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT in Eq. (15a). Let us be explicit in the construction of this system by writing,

j=0N𝕄ijϕj=0,superscriptsubscript𝑗0𝑁subscript𝕄𝑖𝑗subscriptitalic-ϕ𝑗0\sum_{j=0}^{N}\mathbb{M}_{ij}\phi_{j}=0,∑ start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT blackboard_M start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 0 , (15a)

and write with 𝒟nm=(1/n¯)d3𝐯J0𝔻nmsubscript𝒟𝑛𝑚1¯𝑛superscriptd3𝐯subscript𝐽0subscript𝔻𝑛𝑚\mathcal{D}_{nm}=(1/\bar{n})\int\mathrm{d}^{3}\mathbf{v}J_{0}\mathbb{D}_{nm}caligraphic_D start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT = ( 1 / over¯ start_ARG italic_n end_ARG ) ∫ roman_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_v italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT blackboard_D start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT,

𝕄nm=(1+τ𝒟nn)δnm𝒟n,n2δn2,m𝒟n,n+2δn+2,m.subscript𝕄𝑛𝑚1𝜏subscript𝒟𝑛𝑛subscript𝛿𝑛𝑚subscript𝒟𝑛𝑛2subscript𝛿𝑛2𝑚subscript𝒟𝑛𝑛2subscript𝛿𝑛2𝑚\mathbb{M}_{nm}=(1+\tau-\mathcal{D}_{nn})\delta_{nm}-\mathcal{D}_{n,n-2}\delta% _{n-2,m}-\mathcal{D}_{n,n+2}\delta_{n+2,m}.blackboard_M start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT = ( 1 + italic_τ - caligraphic_D start_POSTSUBSCRIPT italic_n italic_n end_POSTSUBSCRIPT ) italic_δ start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT - caligraphic_D start_POSTSUBSCRIPT italic_n , italic_n - 2 end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT italic_n - 2 , italic_m end_POSTSUBSCRIPT - caligraphic_D start_POSTSUBSCRIPT italic_n , italic_n + 2 end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT italic_n + 2 , italic_m end_POSTSUBSCRIPT . (57)

B.2 Solving for a ‘pure’ M𝑀Mitalic_M-th mode

Let us now focus on the problem when there is a dominant M𝑀Mitalic_M-th mode, with MNmuch-less-than𝑀𝑁M\ll Nitalic_M ≪ italic_N. That is, let us take ϕMO(1)similar-tosubscriptitalic-ϕ𝑀𝑂1\phi_{M}\sim O(1)italic_ϕ start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT ∼ italic_O ( 1 ). In addition we make the choice ϕn=0subscriptitalic-ϕ𝑛0\phi_{n}=0italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 0 for n>M𝑛𝑀n>Mitalic_n > italic_M, with which we deem the description of a ‘pure’ mode (more comments on this to follow later). We may then write the equations one by one starting from the M𝑀Mitalic_M-th mode down,

𝒟M+2,MϕMsubscript𝒟𝑀2𝑀subscriptitalic-ϕ𝑀\displaystyle-\mathcal{D}_{M+2,M}\phi_{M}- caligraphic_D start_POSTSUBSCRIPT italic_M + 2 , italic_M end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT =0,absent0\displaystyle=0,= 0 ,
(1+τ𝒟MM)ϕM𝒟M,M2ϕM21𝜏subscript𝒟𝑀𝑀subscriptitalic-ϕ𝑀subscript𝒟𝑀𝑀2subscriptitalic-ϕ𝑀2\displaystyle(1+\tau-\mathcal{D}_{MM})\phi_{M}-\mathcal{D}_{M,M-2}\phi_{M-2}( 1 + italic_τ - caligraphic_D start_POSTSUBSCRIPT italic_M italic_M end_POSTSUBSCRIPT ) italic_ϕ start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT - caligraphic_D start_POSTSUBSCRIPT italic_M , italic_M - 2 end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_M - 2 end_POSTSUBSCRIPT =0,absent0\displaystyle=0,= 0 ,
(1+τ𝒟M2,M2)ϕM2𝒟M2,M4ϕM4𝒟M2,MϕM1𝜏subscript𝒟𝑀2𝑀2subscriptitalic-ϕ𝑀2subscript𝒟𝑀2𝑀4subscriptitalic-ϕ𝑀4subscript𝒟𝑀2𝑀subscriptitalic-ϕ𝑀\displaystyle(1+\tau-\mathcal{D}_{M-2,M-2})\phi_{M-2}-\mathcal{D}_{M-2,M-4}% \phi_{M-4}-\mathcal{D}_{M-2,M}\phi_{M}( 1 + italic_τ - caligraphic_D start_POSTSUBSCRIPT italic_M - 2 , italic_M - 2 end_POSTSUBSCRIPT ) italic_ϕ start_POSTSUBSCRIPT italic_M - 2 end_POSTSUBSCRIPT - caligraphic_D start_POSTSUBSCRIPT italic_M - 2 , italic_M - 4 end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_M - 4 end_POSTSUBSCRIPT - caligraphic_D start_POSTSUBSCRIPT italic_M - 2 , italic_M end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT =0,absent0\displaystyle=0,= 0 ,
\displaystyle\vdots

This system of equations can be straightforwardly solved to O(ϵ)𝑂italic-ϵO(\epsilon)italic_O ( italic_ϵ ) by taking the following ordering for the various modes of ϕitalic-ϕ\phiitalic_ϕ: ϕMO(1)similar-tosubscriptitalic-ϕ𝑀𝑂1\phi_{M}\sim O(1)italic_ϕ start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT ∼ italic_O ( 1 ), ϕM2O(ϵ)similar-tosubscriptitalic-ϕ𝑀2𝑂italic-ϵ\phi_{M-2}\sim O(\epsilon)italic_ϕ start_POSTSUBSCRIPT italic_M - 2 end_POSTSUBSCRIPT ∼ italic_O ( italic_ϵ ) and ϕM2kO(ϵk)similar-tosubscriptitalic-ϕ𝑀2𝑘𝑂superscriptitalic-ϵ𝑘\phi_{M-2k}\sim O(\epsilon^{k})italic_ϕ start_POSTSUBSCRIPT italic_M - 2 italic_k end_POSTSUBSCRIPT ∼ italic_O ( italic_ϵ start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ). In that case, the consistent solution to the problem is the following dispersion relation condition on ω𝜔\omegaitalic_ω and λ𝜆\lambdaitalic_λ,

1+τ𝒟MM=0,1𝜏subscript𝒟𝑀𝑀0\displaystyle 1+\tau-\mathcal{D}_{MM}=0,1 + italic_τ - caligraphic_D start_POSTSUBSCRIPT italic_M italic_M end_POSTSUBSCRIPT = 0 , (58a)
𝒟M+2,M=0,subscript𝒟𝑀2𝑀0\displaystyle\mathcal{D}_{M+2,M}=0,caligraphic_D start_POSTSUBSCRIPT italic_M + 2 , italic_M end_POSTSUBSCRIPT = 0 , (58b)

together with

ϕM2=𝒟M2,M1+τ𝒟M2,M2,subscriptitalic-ϕ𝑀2subscript𝒟𝑀2𝑀1𝜏subscript𝒟𝑀2𝑀2\phi_{M-2}=\frac{\mathcal{D}_{M-2,M}}{1+\tau-\mathcal{D}_{M-2,M-2}},italic_ϕ start_POSTSUBSCRIPT italic_M - 2 end_POSTSUBSCRIPT = divide start_ARG caligraphic_D start_POSTSUBSCRIPT italic_M - 2 , italic_M end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_τ - caligraphic_D start_POSTSUBSCRIPT italic_M - 2 , italic_M - 2 end_POSTSUBSCRIPT end_ARG , (59)

which is order ϵitalic-ϵ\epsilonitalic_ϵ. Thus for studying the M𝑀Mitalic_M-th mode, we must solve Eqs. (58).

Let us start first by investigating the second condition, namely 𝒟M+2,M=0subscript𝒟𝑀2𝑀0\mathcal{D}_{M+2,M}=0caligraphic_D start_POSTSUBSCRIPT italic_M + 2 , italic_M end_POSTSUBSCRIPT = 0. Drop** unimportant factors from Eq. (56b), we are left with the integral condition,

d3𝐯J02ev2/vTi2(1ω~ω)[λ2(ωtxω)2ω~d2ω]=0.superscriptd3𝐯superscriptsubscript𝐽02superscript𝑒superscript𝑣2superscriptsubscript𝑣𝑇𝑖21subscript~𝜔𝜔delimited-[]superscript𝜆2superscriptsubscript𝜔𝑡subscript𝑥parallel-to𝜔2subscript~𝜔𝑑2𝜔0\int\mathrm{d}^{3}\mathbf{v}J_{0}^{2}e^{-v^{2}/v_{Ti}^{2}}\left(1-\frac{\tilde% {\omega}_{\star}}{\omega}\right)\left[\lambda^{2}\left(\frac{\omega_{t}x_{% \parallel}}{\omega}\right)^{2}-\frac{\tilde{\omega}_{d}}{2\omega}\right]=0.∫ roman_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_v italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) [ italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG ] = 0 . (60)

To perform the integrals over velocity space, we use ω~=ω[1+η(x2+x23/2)]subscript~𝜔subscript𝜔delimited-[]1𝜂superscriptsubscript𝑥parallel-to2superscriptsubscript𝑥perpendicular-to232\tilde{\omega}_{\star}=\omega_{\star}\left[1+\eta(x_{\parallel}^{2}+x_{\perp}^% {2}-3/2)\right]over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT [ 1 + italic_η ( italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 3 / 2 ) ] and ω~d=ω¯d(x2+x2/2)subscript~𝜔𝑑subscript¯𝜔𝑑superscriptsubscript𝑥parallel-to2superscriptsubscript𝑥perpendicular-to22\tilde{\omega}_{d}=\bar{\omega}_{d}(x_{\parallel}^{2}+x_{\perp}^{2}/2)over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 ), and for simplicity, drop the FLR contributions to the integral, so that,

λ=ωω¯dωt2.𝜆𝜔subscript¯𝜔𝑑superscriptsubscript𝜔𝑡2\lambda=\sqrt{\frac{\omega\bar{\omega}_{d}}{\omega_{t}^{2}}}.italic_λ = square-root start_ARG divide start_ARG italic_ω over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG . (61)

This is the characteristic localisation of the M𝑀Mitalic_M-th mode. Note that it is actually M𝑀Mitalic_M-independent, meaning that the differences between modes arise from the other expression in Eq. (58). We shall refer to this equation as the M𝑀Mitalic_M-th mode dispersion relation, and we may explicitly write it as,

𝒟=1+τ+ζn¯J02(1ω~ω)f0x2ζ(1+ω¯d2ωx2)d3𝐯,𝒟1𝜏𝜁¯𝑛superscriptsubscript𝐽021subscript~𝜔𝜔subscript𝑓0superscriptsubscript𝑥parallel-to2𝜁1subscript¯𝜔𝑑2𝜔superscriptsubscript𝑥perpendicular-to2superscriptd3𝐯\mathcal{D}=1+\tau+\frac{\zeta}{\bar{n}}\int J_{0}^{2}\left(1-\frac{\tilde{% \omega}_{\star}}{\omega}\right)\frac{f_{0}}{x_{\parallel}^{2}-\zeta\left(1+% \frac{\bar{\omega}_{d}}{2\omega}x_{\perp}^{2}\right)}\mathrm{d}^{3}\mathbf{v},caligraphic_D = 1 + italic_τ + divide start_ARG italic_ζ end_ARG start_ARG over¯ start_ARG italic_n end_ARG end_ARG ∫ italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) divide start_ARG italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ζ ( 1 + divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG roman_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_v , (19)

where,

ζ=12[λ(2M+1)(ωtω)2ω¯d2ω]1.𝜁12superscriptdelimited-[]𝜆2𝑀1superscriptsubscript𝜔𝑡𝜔2subscript¯𝜔𝑑2𝜔1\zeta=\frac{1}{2}\left[\lambda(2M+1)\left(\frac{\omega_{t}}{\omega}\right)^{2}% -\frac{\bar{\omega}_{d}}{2\omega}\right]^{-1}.italic_ζ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG [ italic_λ ( 2 italic_M + 1 ) ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT . (62)

for any 0MN0𝑀much-less-than𝑁0\leq M\ll N0 ≤ italic_M ≪ italic_N, and both for even and odd M𝑀Mitalic_M without distinction. The mode number solely enters the problem through the equations in the kinetic parameter ζ𝜁\zetaitalic_ζ, and in particular, in its streaming contribution. The larger the mode number, the larger the contribution from the streaming term.

Note that one could argue that this particular ‘pure’ solution to the problem is not the only one. In particular, and following the ordering argument of ϕMO(1)similar-tosubscriptitalic-ϕ𝑀𝑂1\phi_{M}\sim O(1)italic_ϕ start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT ∼ italic_O ( 1 ) and ϕM2O(ϵ)similar-tosubscriptitalic-ϕ𝑀2𝑂italic-ϵ\phi_{M-2}\sim O(\epsilon)italic_ϕ start_POSTSUBSCRIPT italic_M - 2 end_POSTSUBSCRIPT ∼ italic_O ( italic_ϵ ), one could consider constructing a solution in a more symmetric way around the M𝑀Mitalic_M-th mode, where the ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT for n>M𝑛𝑀n>Mitalic_n > italic_M are not exactly zero, but ordered like O(ϵ(nM)/2)𝑂superscriptitalic-ϵ𝑛𝑀2O(\epsilon^{(n-M)/2})italic_O ( italic_ϵ start_POSTSUPERSCRIPT ( italic_n - italic_M ) / 2 end_POSTSUPERSCRIPT ). In that instance, the description to order ϵitalic-ϵ\epsilonitalic_ϵ would leave us with a single dispersion equation, namely Eq. (58a). Thus, even if λ𝜆\lambdaitalic_λ should at least have an ordering like that in the ‘pure’ mode, its precise form would not be constrained. In what sense is then the ‘pure’ mode an illustrative choice? There are two important arguments to defend the preeminence of the ‘pure’ mode, albeit not fully conclusive. The first, is that by making the choice of λ𝜆\lambdaitalic_λ above, and as explicitly shown, we make, to order ϵitalic-ϵ\epsilonitalic_ϵ, the lower off-diagonal terms of 𝒟𝒟\mathcal{D}caligraphic_D vanish. As such, the connection of modes to higher harmonics is broken, in a sort of closure scheme, preventing problems with factors of increasing mode numbers and isolating the solution from the truncation point. Second of all, this choice of λ𝜆\lambdaitalic_λ leads to an agreement of our system with the fluid limit in the limit of the latter. Other choices of λ𝜆\lambdaitalic_λ would not do so. And finally, it is the simplest choice to make. All this invites us to study the ‘pure’ modes in this paper. However, in doing so we expect to find discrepancies with the full problem, which likely involves a more subtle involvement of λ𝜆\lambdaitalic_λ. Evidence of this is shown in the numerical comparison to simulations. Although many of the qualitative features of the spectra can be well captured and explained, predicting the exact form and dependence a priori is out of reach. Although this treatment of λ𝜆\lambdaitalic_λ is the weakest point of the treatment, understanding the ‘pure’ mode behaviour is highly insightful. Future work may be devoted to improving the model by perhaps allowing λ𝜆\lambdaitalic_λ as a free parameter to optimise to extremise the growth rate of the ITG (much like a ballooning parameter).

B.2.1 FLR corrections to λ𝜆\lambdaitalic_λ

To give a flavour of variations that λ𝜆\lambdaitalic_λ may be subject to even within the ‘pure’ mode framework, let us show how to include the FLR corrections in Eq. (60). The full-FLR form of the localisation parameter λ𝜆\lambdaitalic_λ takes the form,

λ=ωω¯dωt2[1(b)],𝜆𝜔subscript¯𝜔𝑑superscriptsubscript𝜔𝑡2delimited-[]1𝑏\lambda=\sqrt{\frac{\omega\bar{\omega}_{d}}{\omega_{t}^{2}}\left[1-\mathcal{F}% (b)\right]},italic_λ = square-root start_ARG divide start_ARG italic_ω over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG [ 1 - caligraphic_F ( italic_b ) ] end_ARG , (63)

where

(b)=b2(Γ0Γ1)(1ωω+2ωTω(b1))ωTωΓ1Γ0(1ωω+ωTω(b1))bωTωΓ1.𝑏𝑏2subscriptΓ0subscriptΓ11subscript𝜔𝜔2superscriptsubscript𝜔𝑇𝜔𝑏1superscriptsubscript𝜔𝑇𝜔subscriptΓ1subscriptΓ01subscript𝜔𝜔superscriptsubscript𝜔𝑇𝜔𝑏1𝑏superscriptsubscript𝜔𝑇𝜔subscriptΓ1\mathcal{F}(b)=\frac{b}{2}\frac{(\Gamma_{0}-\Gamma_{1})\left(1-\frac{\omega_{% \star}}{\omega}+2\frac{\omega_{\star}^{T}}{\omega}(b-1)\right)-\frac{\omega_{% \star}^{T}}{\omega}\Gamma_{1}}{\Gamma_{0}\left(1-\frac{\omega_{\star}}{\omega}% +\frac{\omega_{\star}^{T}}{\omega}(b-1)\right)-b\frac{\omega_{\star}^{T}}{% \omega}\Gamma_{1}}.caligraphic_F ( italic_b ) = divide start_ARG italic_b end_ARG start_ARG 2 end_ARG divide start_ARG ( roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ( 1 - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG + 2 divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG ( italic_b - 1 ) ) - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG + divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG ( italic_b - 1 ) ) - italic_b divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG . (64)

It follows from this form that, indeed, in the limit of a small FLR, 00\mathcal{F}\rightarrow 0caligraphic_F → 0, and thus we recover the simple limit λωω¯d/ωt2similar-to𝜆𝜔subscript¯𝜔𝑑superscriptsubscript𝜔𝑡2\lambda\sim\sqrt{\omega\bar{\omega}_{d}/\omega_{t}^{2}}italic_λ ∼ square-root start_ARG italic_ω over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. In the opposite limit, we find that λ(3/2)λb=0𝜆32subscript𝜆𝑏0\lambda\approx(\sqrt{3}/2)\lambda_{b=0}italic_λ ≈ ( square-root start_ARG 3 end_ARG / 2 ) italic_λ start_POSTSUBSCRIPT italic_b = 0 end_POSTSUBSCRIPT, which is roughly a 13%similar-toabsentpercent13\sim 13\%∼ 13 % reduction; i.e., the longitudinal scale of the mode will increase by roughly a 7%similar-toabsentpercent7\sim 7\%∼ 7 % with respect to the no-FLR expectation. The maximum deviation tends to occur when b1similar-to𝑏1b\sim 1italic_b ∼ 1, corresponding to the point where the FLR effects are most efficient. At moderate values for ωT/ωsuperscriptsubscript𝜔𝑇𝜔\omega_{\star}^{T}/\omegaitalic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / italic_ω note that the amplification can be significant following the potentially resonant denominator (see Figure 11). In the strong drive limit the simple ‘pure’ mode should nevertheless be an appropriate qualitative description even if we do not include the variation of λ𝜆\lambdaitalic_λ. There would be no point in kee** a significant added complication to a description through the ‘pure’ mode which is already a convenience choice.

Refer to caption
Figure 11: Finite Larmor radius corrections to λ𝜆\lambdaitalic_λ. Correction factor to λ𝜆\lambdaitalic_λ due to finite Larmor radius effects for a number of different temperature gradient drives, ωT/ωsuperscriptsubscript𝜔𝑇𝜔\omega_{\star}^{T}/\omegaitalic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT / italic_ω, for an ionic unstable mode. The effects of FLR on λ𝜆\lambdaitalic_λ are moderate at large diamagnetic drive, but become very significant near b1similar-to𝑏1b\sim 1italic_b ∼ 1 at lower drives.

Appendix C Expressing the dispersion in terms of the dispersion function

In this appendix we detail the construction of the final form of the dispersion function 𝒟𝒟\mathcal{D}caligraphic_D. We start from,

𝒟=1+τ+ζn¯J02(1ω~ω)f0x2ζ(1+ω¯d2ωx2)d3𝐯,𝒟1𝜏𝜁¯𝑛superscriptsubscript𝐽021subscript~𝜔𝜔subscript𝑓0superscriptsubscript𝑥parallel-to2𝜁1subscript¯𝜔𝑑2𝜔superscriptsubscript𝑥perpendicular-to2superscriptd3𝐯\mathcal{D}=1+\tau+\frac{\zeta}{\bar{n}}\int J_{0}^{2}\left(1-\frac{\tilde{% \omega}_{\star}}{\omega}\right)\frac{f_{0}}{x_{\parallel}^{2}-\zeta\left(1+% \frac{\bar{\omega}_{d}}{2\omega}x_{\perp}^{2}\right)}\mathrm{d}^{3}\mathbf{v},caligraphic_D = 1 + italic_τ + divide start_ARG italic_ζ end_ARG start_ARG over¯ start_ARG italic_n end_ARG end_ARG ∫ italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) divide start_ARG italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ζ ( 1 + divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG roman_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_v , (19)

and recall from the main text that we first considered the integral over xsubscript𝑥parallel-tox_{\parallel}italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT. In doing such an integral, we can write the problem in terms of plasma dispersion functions. However, given the form of I,βI_{\parallel,\beta}italic_I start_POSTSUBSCRIPT ∥ , italic_β end_POSTSUBSCRIPT, Eq. (22), we need to spell out the powers of xsubscript𝑥parallel-tox_{\parallel}italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT in the problem explicitly. To do so we must recall the definitions of ω~dsubscript~𝜔𝑑\tilde{\omega}_{d}over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT and ω~subscript~𝜔\tilde{\omega}_{\star}over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT. With that, the numerator of the integrand,

(1ω~ω)=1subscript~𝜔𝜔absent\displaystyle\left(1-\frac{\tilde{\omega}_{\star}}{\omega}\right)=( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) = [1ωω+ωTω(32x2)]ωTωx2delimited-[]1subscript𝜔𝜔superscriptsubscript𝜔𝑇𝜔32superscriptsubscript𝑥perpendicular-to2superscriptsubscript𝜔𝑇𝜔superscriptsubscript𝑥parallel-to2\displaystyle\left[1-\frac{\omega_{\star}}{\omega}+\frac{\omega_{\star}^{T}}{% \omega}\left(\frac{3}{2}-x_{\perp}^{2}\right)\right]-\frac{\omega_{\star}^{T}}% {\omega}x_{\parallel}^{2}[ 1 - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG + divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG ( divide start_ARG 3 end_ARG start_ARG 2 end_ARG - italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ] - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
=𝒜+x2.absent𝒜superscriptsubscript𝑥parallel-to2\displaystyle=\mathcal{A}+\mathcal{B}x_{\parallel}^{2}.= caligraphic_A + caligraphic_B italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (65)

With this xsubscript𝑥parallel-tox_{\parallel}italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT dependence made explicit, we may perform the first integral over xsubscript𝑥parallel-tox_{\parallel}italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT in Eq. (19). Writing the velocity space measure explicitly as,

d3𝐯=2π(vT)3xdxdx,superscriptd3𝐯2𝜋superscriptsubscript𝑣𝑇3subscript𝑥perpendicular-todsubscript𝑥perpendicular-todsubscript𝑥parallel-to\mathrm{d}^{3}\mathbf{v}=2\pi(v_{T})^{3}x_{\perp}\mathrm{d}x_{\perp}\mathrm{d}% x_{\parallel},roman_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_v = 2 italic_π ( italic_v start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT roman_d italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT roman_d italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT , (66)

we have,

I𝐼\displaystyle Iitalic_I =1n¯f0J02(1ω~ω)x2ζ(1+ω¯d2ωx2)d3𝐯=absent1¯𝑛subscript𝑓0superscriptsubscript𝐽021subscript~𝜔𝜔superscriptsubscript𝑥parallel-to2𝜁1subscript¯𝜔𝑑2𝜔superscriptsubscript𝑥perpendicular-to2superscriptd3𝐯absent\displaystyle=\frac{1}{\bar{n}}\int f_{0}J_{0}^{2}\frac{\left(1-\frac{\tilde{% \omega}_{\star}}{\omega}\right)}{x_{\parallel}^{2}-\zeta\left(1+\frac{\bar{% \omega}_{d}}{2\omega}x_{\perp}^{2}\right)}\mathrm{d}^{3}\mathbf{v}== divide start_ARG 1 end_ARG start_ARG over¯ start_ARG italic_n end_ARG end_ARG ∫ italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG ( 1 - divide start_ARG over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ) end_ARG start_ARG italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ζ ( 1 + divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG roman_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bold_v = (67)
=20J0(x2b)2xex2[𝒜Z0(ζ¯)+Z1(ζ¯)]absent2superscriptsubscript0subscript𝐽0superscriptsubscript𝑥perpendicular-to2𝑏2subscript𝑥perpendicular-tosuperscript𝑒superscriptsubscript𝑥perpendicular-to2delimited-[]𝒜subscript𝑍0¯𝜁subscript𝑍1¯𝜁\displaystyle=2\int_{0}^{\infty}J_{0}\left(x_{\perp}\sqrt{2b}\right)^{2}x_{% \perp}e^{-x_{\perp}^{2}}\left[\mathcal{A}Z_{0}(\bar{\zeta})+\mathcal{B}Z_{1}(% \bar{\zeta})\right]= 2 ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT square-root start_ARG 2 italic_b end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT [ caligraphic_A italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( over¯ start_ARG italic_ζ end_ARG ) + caligraphic_B italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( over¯ start_ARG italic_ζ end_ARG ) ] (68)

where,

ζ¯=ζ(1+ω¯d2ωx2),¯𝜁𝜁1subscript¯𝜔𝑑2𝜔superscriptsubscript𝑥perpendicular-to2\displaystyle\bar{\zeta}=\zeta\left(1+\frac{\bar{\omega}_{d}}{2\omega}x_{\perp% }^{2}\right),over¯ start_ARG italic_ζ end_ARG = italic_ζ ( 1 + divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω end_ARG italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) , (69)
Z0(x)=Z(x)x,subscript𝑍0𝑥𝑍𝑥𝑥\displaystyle Z_{0}(x)=\frac{Z(\sqrt[*]{x})}{\sqrt[*]{x}},italic_Z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) = divide start_ARG italic_Z ( nth-root start_ARG ∗ end_ARG start_ARG italic_x end_ARG ) end_ARG start_ARG nth-root start_ARG ∗ end_ARG start_ARG italic_x end_ARG end_ARG , (70)
Z1(x)=(1+xZ(x)),subscript𝑍1𝑥1𝑥𝑍𝑥\displaystyle Z_{1}(x)=\left(1+\sqrt[*]{x}Z(\sqrt[*]{x})\right),italic_Z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x ) = ( 1 + nth-root start_ARG ∗ end_ARG start_ARG italic_x end_ARG italic_Z ( nth-root start_ARG ∗ end_ARG start_ARG italic_x end_ARG ) ) , (71)
Z2(x)=12[1+2x(1+xZ(x))].subscript𝑍2𝑥12delimited-[]12𝑥1𝑥𝑍𝑥\displaystyle Z_{2}(x)=\frac{1}{2}\left[1+2x(1+\sqrt[*]{x}Z(\sqrt[*]{x}))% \right].italic_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG [ 1 + 2 italic_x ( 1 + nth-root start_ARG ∗ end_ARG start_ARG italic_x end_ARG italic_Z ( nth-root start_ARG ∗ end_ARG start_ARG italic_x end_ARG ) ) ] . (72)

It is now the time of performing the integral over xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT. The main difficulty here is that xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT appears in the arguments of the plasma dispersion functions over which we need to be integrating. Integrals of this form for small FLR effects have been recently found in exact form by Ivanov & Adkins (2023), but full FLR effects are sought here. To that end, we proceed by exploiting the ω¯d/ω1much-less-thansubscript¯𝜔𝑑𝜔1\bar{\omega}_{d}/\omega\ll 1over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω ≪ 1 assumption, together with the exponential exp[x2]superscriptsubscript𝑥perpendicular-to2\exp[-x_{\perp}^{2}]roman_exp [ - italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] that limits the values of xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT from becoming much larger than one, and expand the functions in the integrand. Application of Taylor expansion and the chain rule will yield different powers of xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT in the integrand. This procedure is straightforward, and may be efficiently calculated with the aid of computer algebra. Note that by performing this expansion, we are losing the ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT resonance in xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT while we have kept its xsubscript𝑥parallel-tox_{\parallel}italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT part. The ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT resonance effects are thus only partially captured.

Integrals over powers of xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT with the other factors in the integrand of Eq. (67) are well known (Kadomtsev & Pogutse, 1970a)(Gradshteyn & Ryzhik, 2014, Eq. 6.615). Thus, all we need to do is collect powers of xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT and collect terms. Terms corresponding to a particular power of xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT will be multiplied by the appropriate FLR factor that results from the integral. The relevant xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT integrals are,

F0(b)=20J0(x2b)2xex2=Γ0(b),subscript𝐹0𝑏2superscriptsubscript0subscript𝐽0superscriptsubscript𝑥perpendicular-to2𝑏2subscript𝑥perpendicular-tosuperscript𝑒superscriptsubscript𝑥perpendicular-to2subscriptΓ0𝑏\displaystyle F_{0}(b)=2\int_{0}^{\infty}J_{0}\left(x_{\perp}\sqrt{2b}\right)^% {2}x_{\perp}e^{-x_{\perp}^{2}}=\Gamma_{0}(b),italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_b ) = 2 ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT square-root start_ARG 2 italic_b end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT = roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_b ) , (73a)
F2(b)=20J0(x2b)2x3ex2=(1b)Γ0(b)+bΓ1(b),subscript𝐹2𝑏2superscriptsubscript0subscript𝐽0superscriptsubscript𝑥perpendicular-to2𝑏2superscriptsubscript𝑥perpendicular-to3superscript𝑒superscriptsubscript𝑥perpendicular-to21𝑏subscriptΓ0𝑏𝑏subscriptΓ1𝑏\displaystyle F_{2}(b)=2\int_{0}^{\infty}J_{0}\left(x_{\perp}\sqrt{2b}\right)^% {2}x_{\perp}^{3}e^{-x_{\perp}^{2}}=(1-b)\Gamma_{0}(b)+b\Gamma_{1}(b),italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_b ) = 2 ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT square-root start_ARG 2 italic_b end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT = ( 1 - italic_b ) roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_b ) + italic_b roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_b ) , (73b)
F4(b)=20J0(x2b)2x5ex2=2[(1b)2Γ0(b)+(32b)bΓ1(b)],subscript𝐹4𝑏2superscriptsubscript0subscript𝐽0superscriptsubscript𝑥perpendicular-to2𝑏2superscriptsubscript𝑥perpendicular-to5superscript𝑒superscriptsubscript𝑥perpendicular-to22delimited-[]superscript1𝑏2subscriptΓ0𝑏32𝑏𝑏subscriptΓ1𝑏\displaystyle F_{4}(b)=2\int_{0}^{\infty}J_{0}\left(x_{\perp}\sqrt{2b}\right)^% {2}x_{\perp}^{5}e^{-x_{\perp}^{2}}=2\left[(1-b)^{2}\Gamma_{0}(b)+\left(\frac{3% }{2}-b\right)b\Gamma_{1}(b)\right],italic_F start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_b ) = 2 ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT square-root start_ARG 2 italic_b end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT = 2 [ ( 1 - italic_b ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_b ) + ( divide start_ARG 3 end_ARG start_ARG 2 end_ARG - italic_b ) italic_b roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_b ) ] , (73c)

The resulting 𝒟𝒟\mathcal{D}caligraphic_D can be written as,

𝒟=1+τ+𝒯(α,β,γ)ω¯dαωβωα+βFγ(b).𝒟1𝜏subscript𝒯𝛼𝛽𝛾superscriptsubscript¯𝜔𝑑𝛼superscriptsubscript𝜔𝛽superscript𝜔𝛼𝛽subscript𝐹𝛾𝑏\mathcal{D}=1+\tau+\sum\mathcal{T}_{(\alpha,\beta,\gamma)}\frac{\bar{\omega}_{% d}^{\alpha}\omega_{\star}^{\beta}}{\omega^{\alpha+\beta}}F_{\gamma}(b).caligraphic_D = 1 + italic_τ + ∑ caligraphic_T start_POSTSUBSCRIPT ( italic_α , italic_β , italic_γ ) end_POSTSUBSCRIPT divide start_ARG over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT italic_α + italic_β end_POSTSUPERSCRIPT end_ARG italic_F start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ( italic_b ) . (74)

Some of the leading order 𝒯(α,β,γ)subscript𝒯𝛼𝛽𝛾\mathcal{T}_{(\alpha,\beta,\gamma)}caligraphic_T start_POSTSUBSCRIPT ( italic_α , italic_β , italic_γ ) end_POSTSUBSCRIPT terms are shown in Table 1 for reference. Many of the terms included are not necessary. They are not with regards to explaining the physical behaviour of the mode, and are in addition higher order in δ𝛿\deltaitalic_δ and ϵitalic-ϵ\epsilonitalic_ϵ than originally devised for. Nevertheless, it may be helpful in analysing the behaviour of the dispersion equation.

α𝛼\alphaitalic_α β𝛽\betaitalic_β γ𝛾\gammaitalic_γ 𝒯(α,β,γ)subscript𝒯𝛼𝛽𝛾\mathcal{T}_{(\alpha,\beta,\gamma)}caligraphic_T start_POSTSUBSCRIPT ( italic_α , italic_β , italic_γ ) end_POSTSUBSCRIPT
0 0 0 ζZ(ζ)𝜁𝑍𝜁\sqrt{\zeta}Z(\sqrt{\zeta})square-root start_ARG italic_ζ end_ARG italic_Z ( square-root start_ARG italic_ζ end_ARG )
0 1 0 (13η2)ζZ(ζ)ηζZ+(ζ)13𝜂2𝜁𝑍𝜁𝜂𝜁subscript𝑍𝜁~{}~{}-\left(1-\frac{3\eta}{2}\right)\sqrt{\zeta}Z(\sqrt{\zeta})-\eta\zeta Z_{% +}(\sqrt{\zeta})~{}~{}- ( 1 - divide start_ARG 3 italic_η end_ARG start_ARG 2 end_ARG ) square-root start_ARG italic_ζ end_ARG italic_Z ( square-root start_ARG italic_ζ end_ARG ) - italic_η italic_ζ italic_Z start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( square-root start_ARG italic_ζ end_ARG )
0 1 2 ηζZ(ζ)𝜂𝜁𝑍𝜁~{}~{}-\eta\sqrt{\zeta}Z(\sqrt{\zeta})~{}~{}- italic_η square-root start_ARG italic_ζ end_ARG italic_Z ( square-root start_ARG italic_ζ end_ARG )
1 0 2 14[1(1+2ζ)Z+(ζ)]14delimited-[]112𝜁subscript𝑍𝜁~{}~{}\frac{1}{4}\left[1-(1+2\zeta)Z_{+}(\sqrt{\zeta})\right]~{}~{}divide start_ARG 1 end_ARG start_ARG 4 end_ARG [ 1 - ( 1 + 2 italic_ζ ) italic_Z start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( square-root start_ARG italic_ζ end_ARG ) ]
1 1 2 14{1+η(32+ζ)+[1+2ζ+η(32+2ζ(ζ2))]Z+(ζ)}141𝜂32𝜁delimited-[]12𝜁𝜂322𝜁𝜁2subscript𝑍𝜁~{}~{}\frac{1}{4}\left\{-1+\eta\left(\frac{3}{2}+\zeta\right)+\left[1+2\zeta+% \eta\left(-\frac{3}{2}+2\zeta(\zeta-2)\right)\right]Z_{+}(\sqrt{\zeta})\right% \}~{}~{}divide start_ARG 1 end_ARG start_ARG 4 end_ARG { - 1 + italic_η ( divide start_ARG 3 end_ARG start_ARG 2 end_ARG + italic_ζ ) + [ 1 + 2 italic_ζ + italic_η ( - divide start_ARG 3 end_ARG start_ARG 2 end_ARG + 2 italic_ζ ( italic_ζ - 2 ) ) ] italic_Z start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( square-root start_ARG italic_ζ end_ARG ) }
1 1 4 η4[1+(12ζ)Z+(ζ)]𝜂4delimited-[]112𝜁subscript𝑍𝜁~{}~{}\frac{\eta}{4}\left[-1+(1-2\zeta)Z_{+}(\sqrt{\zeta})\right]~{}~{}divide start_ARG italic_η end_ARG start_ARG 4 end_ARG [ - 1 + ( 1 - 2 italic_ζ ) italic_Z start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( square-root start_ARG italic_ζ end_ARG ) ]
2 0 4 116[32+ζ+(32+2ζ(1+ζ))Z+(ζ)]116delimited-[]32𝜁322𝜁1𝜁subscript𝑍𝜁~{}~{}\frac{1}{16}\left[-\frac{3}{2}+\zeta+\left(\frac{3}{2}+2\zeta(1+\zeta)% \right)Z_{+}(\sqrt{\zeta})\right]~{}~{}divide start_ARG 1 end_ARG start_ARG 16 end_ARG [ - divide start_ARG 3 end_ARG start_ARG 2 end_ARG + italic_ζ + ( divide start_ARG 3 end_ARG start_ARG 2 end_ARG + 2 italic_ζ ( 1 + italic_ζ ) ) italic_Z start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( square-root start_ARG italic_ζ end_ARG ) ]
2 1 4 132{32ζη(92+2ζ(ζ1))+[34ζ(1+ζ)+η(92+ζ(7+2ζ(52ζ)))]Z+(ζ)}13232𝜁𝜂922𝜁𝜁1delimited-[]34𝜁1𝜁𝜂92𝜁72𝜁52𝜁subscript𝑍𝜁~{}~{}\frac{1}{32}\left\{3-2\zeta-\eta\left(\frac{9}{2}+2\zeta(\zeta-1)\right)% +\left[-3-4\zeta(1+\zeta)+\eta\left(\frac{9}{2}+\zeta\left(7+2\zeta(5-2\zeta)% \right)\right)\right]Z_{+}(\sqrt{\zeta})\right\}~{}~{}divide start_ARG 1 end_ARG start_ARG 32 end_ARG { 3 - 2 italic_ζ - italic_η ( divide start_ARG 9 end_ARG start_ARG 2 end_ARG + 2 italic_ζ ( italic_ζ - 1 ) ) + [ - 3 - 4 italic_ζ ( 1 + italic_ζ ) + italic_η ( divide start_ARG 9 end_ARG start_ARG 2 end_ARG + italic_ζ ( 7 + 2 italic_ζ ( 5 - 2 italic_ζ ) ) ) ] italic_Z start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( square-root start_ARG italic_ζ end_ARG ) }
2 1 6 η16[32ζ(32+2ζ(1+ζ))Z+(ζ)]𝜂16delimited-[]32𝜁322𝜁1𝜁subscript𝑍𝜁~{}~{}\frac{\eta}{16}\left[\frac{3}{2}-\zeta-\left(\frac{3}{2}+2\zeta(1+\zeta)% \right)Z_{+}(\sqrt{\zeta})\right]~{}~{}divide start_ARG italic_η end_ARG start_ARG 16 end_ARG [ divide start_ARG 3 end_ARG start_ARG 2 end_ARG - italic_ζ - ( divide start_ARG 3 end_ARG start_ARG 2 end_ARG + 2 italic_ζ ( 1 + italic_ζ ) ) italic_Z start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( square-root start_ARG italic_ζ end_ARG ) ]
Table 1: Contributions to the dispersion relation. Different term contributions to the dispersion relation 𝒟𝒟\mathcal{D}caligraphic_D which may or may not be included depending on the required approximation. The columns α𝛼\alphaitalic_α and β𝛽\betaitalic_β denote the powers of ω¯d/ωsubscript¯𝜔𝑑𝜔\bar{\omega}_{d}/\omegaover¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω and ω/ωsubscript𝜔𝜔\omega_{\star}/\omegaitalic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT / italic_ω respectively that these terms are multiplied by, Column γ𝛾\gammaitalic_γ labels the FLR function that is multiplied by these terms, related directly to the number of powers of xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT prior to integrating over xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT. The notation Z+(ζ)=1+ζZ(ζ)subscript𝑍𝜁1𝜁𝑍𝜁Z_{+}(\sqrt{\zeta})=1+\sqrt{\zeta}Z(\sqrt{\zeta})italic_Z start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( square-root start_ARG italic_ζ end_ARG ) = 1 + square-root start_ARG italic_ζ end_ARG italic_Z ( square-root start_ARG italic_ζ end_ARG ) for brevity.

Appendix D Full Larmor radius form of the ITG fluid equation

In this Appendix we sketch the derivation of the full-Larmor-radius form of the ITG fluid equation. We follow closely the work of Connor et al. (1980), and where possible we shall simply quote this work. Let us remind ourselves about the set-up of the fluid ITG problem. We start by assuming that the transit time is long compared to the characteristic time of the instability (ωt/ωϵ1similar-tosubscript𝜔𝑡𝜔italic-ϵmuch-less-than1\omega_{t}/\omega\sim\epsilon\ll 1italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / italic_ω ∼ italic_ϵ ≪ 1), so that we may consider expanding our solution to the GK equation ignoring any kinetic resonance there.

The general solution to the GK equation in Eq. (1) for passing particles can be written using an integrating factor (Connor et al., 1980, Eqs. (15)-(16)),

gp=iσ(ωω~)f0σJ0ϕ|v|eiσM(,),subscript𝑔𝑝𝑖𝜎𝜔subscript~𝜔subscript𝑓0superscriptsubscript𝜎subscript𝐽0italic-ϕsubscript𝑣parallel-tosuperscript𝑒𝑖𝜎𝑀superscriptg_{p}=-i\sigma(\omega-\tilde{\omega}_{\star})f_{0}\int_{-\sigma\infty}^{\ell}% \frac{J_{0}\phi}{|v_{\parallel}|}e^{i\sigma M(\ell^{\prime},\ell)},italic_g start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = - italic_i italic_σ ( italic_ω - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ) italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT - italic_σ ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ℓ end_POSTSUPERSCRIPT divide start_ARG italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ϕ end_ARG start_ARG | italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT | end_ARG italic_e start_POSTSUPERSCRIPT italic_i italic_σ italic_M ( roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , roman_ℓ ) end_POSTSUPERSCRIPT , (75)

and,

M(a,b)=abωω~d|v|d,𝑀𝑎𝑏superscriptsubscript𝑎𝑏𝜔subscript~𝜔𝑑subscript𝑣parallel-todifferential-dM(a,b)=\int_{a}^{b}\frac{\omega-\tilde{\omega}_{d}}{|v_{\parallel}|}\mathrm{d}\ell,italic_M ( italic_a , italic_b ) = ∫ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_b end_POSTSUPERSCRIPT divide start_ARG italic_ω - over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG | italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT | end_ARG roman_d roman_ℓ , (76)

with σ𝜎\sigmaitalic_σ the sign of vsubscript𝑣parallel-tov_{\parallel}italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT. In writing the solution for gpsubscript𝑔𝑝g_{p}italic_g start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT we implemented the usual vanishing conditions for {ω}>0𝜔0\Im\{\omega\}>0roman_ℑ { italic_ω } > 0. We will be considering the contribution from passing ions, treating the electron response to be adiabatic.

As the function M(,)𝑀superscriptM(\ell^{\prime},\ell)italic_M ( roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , roman_ℓ ) in the exponent of the integrand scales like ω/ωtsimilar-toabsent𝜔subscript𝜔𝑡\sim\omega/\omega_{t}∼ italic_ω / italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, the exponent is large. We may then approximate the whole integral integrating by parts (Bender & Orszag, 2013), and kee** terms up to order O(ϵ)𝑂italic-ϵO(\epsilon)italic_O ( italic_ϵ ). In doing so we assume that ϕϕ/ϵsimilar-tosubscriptitalic-ϕitalic-ϕitalic-ϵ\partial_{\ell}\phi\sim\phi/\epsilon∂ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT italic_ϕ ∼ italic_ϕ / italic_ϵ and ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT to be ordered like ω𝜔\omegaitalic_ω or smaller.

Once expanded, we must then integrate gpsubscript𝑔𝑝g_{p}italic_g start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT over velocity space and apply the quasineutrality condition (Connor et al., 1980, Eq. (19a)). Upon careful explicit evaluation of the integrals over vsubscript𝑣parallel-tov_{\parallel}italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT we obtain plasma dispersion functions (resonance coming purely from the velocity dependence of ω~dsubscript~𝜔𝑑\tilde{\omega}_{d}over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT), and we may write the resulting quasineutrality condition as (Connor et al., 1980, Eqs.(35)-(37)) after some algebra.

To derive the much simpler looking eigenvalue equation in the fluid limit of papers such as Plunk et al. (2014); Zocco et al. (2016), one must, in addition to assuming the smallness of the streaming contribution to the mode, also assume the smallness of the drift frequency ωd/ω1much-less-thansubscript𝜔𝑑𝜔1\omega_{d}/\omega\ll 1italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT / italic_ω ≪ 1. This particular ordering allows us to ignore the drift frequency resonance that led to the appearence of plasma dispersion functions, as in the large argument limit, ignoring this resonance simply incurs in an exponentially small error. With this expansion, the integral over xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT left to be done simply become standard, and we shall use the same notation as in the text, Eqs. (LABEL:eqn:weber_integrals), for these integrals, Fn(b)subscript𝐹𝑛𝑏F_{n}(b)italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_b ). Kee** the leading order terms it can be shown then that one has,

(1+ταT)ϕB[βBTϕ]=0,1𝜏𝛼𝑇italic-ϕ𝐵subscriptdelimited-[]𝛽𝐵𝑇subscriptitalic-ϕ0(1+\tau-\frac{\alpha}{T})\phi-B\partial_{\ell}\left[\frac{\beta}{BT}\partial_{% \ell}\phi\right]=0,( 1 + italic_τ - divide start_ARG italic_α end_ARG start_ARG italic_T end_ARG ) italic_ϕ - italic_B ∂ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT [ divide start_ARG italic_β end_ARG start_ARG italic_B italic_T end_ARG ∂ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT italic_ϕ ] = 0 , (77a)
αT=Γ0[1ωω(1η)ωdω2ω2]F2(b)(ωTω+ωωd2ω2)ωTωd2ω2F4(b),𝛼𝑇subscriptΓ0delimited-[]1subscript𝜔𝜔1𝜂subscript𝜔𝑑subscript𝜔2superscript𝜔2subscript𝐹2𝑏superscriptsubscript𝜔𝑇𝜔subscript𝜔subscript𝜔𝑑2superscript𝜔2superscriptsubscript𝜔𝑇subscript𝜔𝑑2superscript𝜔2subscript𝐹4𝑏\frac{\alpha}{T}=\Gamma_{0}\left[1-\frac{\omega_{\star}}{\omega}(1-\eta)-\frac% {\omega_{d}\omega_{\star}}{2\omega^{2}}\right]-F_{2}(b)\left(\frac{\omega_{% \star}^{T}}{\omega}+\frac{\omega_{\star}\omega_{d}}{2\omega^{2}}\right)-\frac{% \omega_{\star}^{T}\omega_{d}}{2\omega^{2}}F_{4}(b),divide start_ARG italic_α end_ARG start_ARG italic_T end_ARG = roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT [ 1 - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_ω end_ARG ( 1 - italic_η ) - divide start_ARG italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] - italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_b ) ( divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG + divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_b ) , (77b)
βT=vTi2ωTω3F2(b).𝛽𝑇superscriptsubscript𝑣𝑇𝑖2superscriptsubscript𝜔𝑇superscript𝜔3subscript𝐹2𝑏\frac{\beta}{T}=\frac{v_{Ti}^{2}\omega_{\star}^{T}}{\omega^{3}}F_{2}(b).divide start_ARG italic_β end_ARG start_ARG italic_T end_ARG = divide start_ARG italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_b ) . (77c)

As a note of caution, here we took m=1𝑚1m=1italic_m = 1 for the mass. Putting these all together in a more succinct way, and taking for simplicity ω=0subscript𝜔0\omega_{\star}=0italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = 0 but ωT0superscriptsubscript𝜔𝑇0\omega_{\star}^{T}\neq 0italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ≠ 0, for our problem in which B𝐵Bitalic_B is constant, and the only spatial dependence in the field is in the drift frequency, we may write

[(1+τΓ0)ωTωb(Γ0Γ1)+ωTωd2ω2F4(b)]ϕωt2ωT2ω3¯2ϕ=0,delimited-[]1𝜏subscriptΓ0superscriptsubscript𝜔𝑇𝜔𝑏subscriptΓ0subscriptΓ1superscriptsubscript𝜔𝑇subscript𝜔𝑑2superscript𝜔2subscript𝐹4𝑏italic-ϕsuperscriptsubscript𝜔𝑡2superscriptsubscript𝜔𝑇2superscript𝜔3superscriptsubscript¯2italic-ϕ0\left[(1+\tau-\Gamma_{0})-\frac{\omega_{\star}^{T}}{\omega}b(\Gamma_{0}-\Gamma% _{1})+\frac{\omega_{\star}^{T}\omega_{d}}{2\omega^{2}}F_{4}(b)\right]\phi-% \frac{\omega_{t}^{2}\omega_{\star}^{T}}{2\omega^{3}}\partial_{\bar{\ell}}^{2}% \phi=0,[ ( 1 + italic_τ - roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) - divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω end_ARG italic_b ( roman_Γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) + divide start_ARG italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_F start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_b ) ] italic_ϕ - divide start_ARG italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_ω start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ∂ start_POSTSUBSCRIPT over¯ start_ARG roman_ℓ end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϕ = 0 , (78)

where we have defined the transit frequency ωtsubscript𝜔𝑡\omega_{t}italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT using the thermal velocity of the ions (which this equation attempts to describe) and some reference parallel length scale. This full-FLR form of the fluid equation is consistent with the fluid asymptotic limit of the dispersion function, Eq. (32), discussed in this paper. The small b𝑏bitalic_b limit is precisely (upon relaxing ω=0subscript𝜔0\omega_{\star}=0italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = 0) of the form employed in Plunk et al. (2014); Zocco et al. (2016).

Appendix E Notation glossary

In Table 2 we summarise the notation employed in the paper for reference. Many of the variables used are standard in gyrokinetics.

\ellroman_ℓ Length along the field line ¯¯\bar{\ell}over¯ start_ARG roman_ℓ end_ARG Normalised length along the field line
α𝛼\alphaitalic_α Straight field line field line label ψ𝜓\psiitalic_ψ 2π2𝜋2\pi2 italic_π times the toroidal magnetic flux (surface label)
qisubscript𝑞𝑖q_{i}italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT Ion charge Tisubscript𝑇𝑖T_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT Ion temperature
n¯¯𝑛\bar{n}over¯ start_ARG italic_n end_ARG Density ρisubscript𝜌𝑖\rho_{i}italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT Ion Larmor radius
vTisubscript𝑣𝑇𝑖v_{Ti}italic_v start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT Thermal speed of ions ω𝜔\omegaitalic_ω Mode frequency
γ𝛾\gammaitalic_γ Mode growth rate ω¯dsubscript¯𝜔𝑑\bar{\omega}_{d}over¯ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT Negative of the drift frequency at bottom of bad curvature well (6)
ωsubscript𝜔\omega_{\star}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT Density gradient driven diamagnetic drift (1) ωTsuperscriptsubscript𝜔𝑇\omega_{\star}^{T}italic_ω start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT Temperature gradient driven diamagnetic drift
η𝜂\etaitalic_η Ratio of temperature to density gradient ω~dsubscript~𝜔𝑑\tilde{\omega}_{d}over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT Drift frequency (with velocity dependence)
ω~subscript~𝜔\tilde{\omega}_{\star}over~ start_ARG italic_ω end_ARG start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT Diamagnetic frequency (with velocity dependence) ωtsubscript𝜔𝑡\omega_{t}italic_ω start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT Transit frequency (10)
ΛΛ\Lambdaroman_Λ Half width of bad curvature region (6) vsubscript𝑣parallel-tov_{\parallel}italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT Parallel velocity
xsubscript𝑥parallel-tox_{\parallel}italic_x start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT Normalised parallel velocity σ𝜎\sigmaitalic_σ Sign of vsubscript𝑣parallel-tov_{\parallel}italic_v start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT
vsubscript𝑣perpendicular-tov_{\perp}italic_v start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT Perpendicular velocity xsubscript𝑥perpendicular-tox_{\perp}italic_x start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT Normalised perpendicular velocity
kαsubscript𝑘𝛼k_{\alpha}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT Poloidal wavenumber (1) kψsubscript𝑘𝜓k_{\psi}italic_k start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT Normal wavenumber (1)
𝐤subscript𝐤perpendicular-to\mathbf{k}_{\perp}bold_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT Perpendicular wavevector b𝑏bitalic_b Finite Larmor radius parameter (1)
F0isubscript𝐹0𝑖F_{0i}italic_F start_POSTSUBSCRIPT 0 italic_i end_POSTSUBSCRIPT Leading order ion Maxwellian distribution g𝑔gitalic_g Non-adiabatic perturbed distribution function
gnsubscript𝑔𝑛g_{n}italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT n𝑛nitalic_n-th Taylor-Gauss mode of g𝑔gitalic_g (11) ϕitalic-ϕ\phiitalic_ϕ Electrostatic potential
ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT n𝑛nitalic_n-th Taylor-Gauss mode of ϕitalic-ϕ\phiitalic_ϕ (11) τ𝜏\tauitalic_τ Ratio of ion to electron temperature (5)
λ𝜆\lambdaitalic_λ Exponential envelope parameter (11) ζ𝜁\zetaitalic_ζ Kinetic parameter (20a)
δ𝛿\deltaitalic_δ Small drift ordering parameter ϵitalic-ϵ\epsilonitalic_ϵ Mode localisation ordering parameter (8)
Table 2: Glossary of notation. Table including the symbols employed throughout the paper and their informal meaning.

References

  • Abramowitz & Stegun (1968) Abramowitz, Milton & Stegun, Irene A 1968 Handbook of mathematical functions with formulas, graphs, and mathematical tables, , vol. 55. US Government printing office.
  • Anderson et al. (1995) Anderson, F S.B., Almagri, A F, Anderson, D T, Matthews, P G, Talmadge, J N & Shohet, J L 1995 Helically symmetric experiment, (hsx) goals, design and status. Fusion Technology 27.
  • Antonsen Jr & Lane (1980) Antonsen Jr, Thomas M & Lane, Barton 1980 Kinetic equations for low frequency instabilities in inhomogeneous plasmas. The Physics of Fluids 23 (6), 1205–1214.
  • Barnes et al. (2019) Barnes, Michael, Parra, Felix I & Landreman, Matt 2019 stella: An operator-split, implicit–explicit δ𝛿\deltaitalic_δf-gyrokinetic code for general magnetic field configurations. Journal of Computational Physics 391, 365–380.
  • Bender & Orszag (2013) Bender, Carl M & Orszag, Steven A 2013 Advanced mathematical methods for scientists and engineers I: Asymptotic methods and perturbation theory. Springer Science & Business Media.
  • Bernstein et al. (1958) Bernstein, Ira B, Frieman, EA, Kruskal, Martin David & Kulsrud, RM91737 1958 An energy principle for hydromagnetic stability problems. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences 244 (1236), 17–40.
  • Biglari et al. (1989) Biglari, H, Diamond, PH & Rosenbluth, MN 1989 Toroidal ion-pressure-gradient-driven drift instabilities and transport revisited. Physics of Fluids B: Plasma Physics 1 (1), 109–118.
  • Boozer (1983) Boozer, Allen H 1983 Transport and isomorphic equilibria. The Physics of Fluids 26 (2), 496–499.
  • Connor et al. (1980) Connor, JW, Hastie, RJ & Taylor, JB 1980 Stability of general plasma equilibria. iii. Plasma Physics 22 (7), 757.
  • Connor et al. (1978) Connor, J W, Hastie, R J & Taylor, J B 1978 Phys. Rev. Lett. 40 (6), 396.
  • Coppi et al. (1967) Coppi, B., Rosenbluth, M. N. & Sagdeev, R. Z. 1967 Phys. Fluids 10 (3), 582.
  • Correa-Restrepo (1978) Correa-Restrepo, D 1978 Ballooning modes in three-dimensional mhd equilibria with shear. Zeitschrift für Naturforschung A 33 (7), 789–791.
  • Cowley et al. (1991) Cowley, S. C., Kulsrud, R. M. & Sudan, R. 1991 Considerations of ion-temperature-gradient-driven turbulence. Physics of Fluids B: Plasma Physics 3 (10), 2767–2782.
  • D’haeseleer et al. (2012) D’haeseleer, W. D., Hitchon, W. N. G., Callen, J. D. & Shohet, J. L. 2012 Flux coordinates and magnetic field structure: a guide to a fundamental tool of plasma theory. Springer Science & Business Media.
  • Fried & Conte (2015) Fried, Burton D & Conte, Samuel D 2015 The plasma dispersion function: the Hilbert transform of the Gaussian. Academic press.
  • Gao et al. (2003) Gao, Zhe, Sanuki, H, Itoh, K & Dong, JQ 2003 Temperature gradient driven short wavelength modes in sheared slab plasmas. Physics of Plasmas 10 (7), 2831–2839.
  • Gao et al. (2005) Gao, Zhe, Sanuki, H, Itoh, K & Dong, JQ 2005 Short wavelength ion temperature gradient instability in toroidal plasmas. Physics of plasmas 12 (2).
  • Gradshteyn & Ryzhik (2014) Gradshteyn, Izrail Solomonovich & Ryzhik, Iosif Moiseevich 2014 Table of integrals, series, and products. Academic press.
  • Greene & Johnson (1962) Greene, John M & Johnson, John L 1962 Stability criterion for arbitrary hydromagnetic equilibria. The Physics of Fluids 5 (5), 510–517.
  • Guo & Romanelli (1993) Guo, S. C. & Romanelli, F. 1993 Physics of Fluids B: Plasma Physics 5 (2), 520–533.
  • Hahm & Tang (1988) Hahm, TS & Tang, WM 1988 Properties of ion temperature gradient drift instabilities in h-mode plasmas. Tech. Rep.. Princeton Plasma Physics Lab.(PPPL), Princeton, NJ (United States).
  • Hastie (2013) Hastie, R J 2013 private communication.
  • Hirose et al. (2002) Hirose, A, Elia, M, Smolyakov, AI & Yagi, M 2002 Short wavelength temperature gradient driven modes in tokamaks. Physics of Plasmas 9 (5), 1659–1666.
  • Horton Jr et al. (1981) Horton Jr, Wendell, Choi, Duk-In & Tang, WM 1981 Toroidal drift modes driven by ion pressure gradients. The Physics of Fluids 24 (6), 1077–1085.
  • Ivanov & Adkins (2023) Ivanov, PG & Adkins, T 2023 An analytical form of the dispersion function for local linear gyrokinetics in a curved magnetic field. Journal of Plasma Physics 89 (2), 905890213.
  • Jenko et al. (2001) Jenko, F, Dorland, W & Hammett, GW 2001 Critical gradient formula for toroidal electron temperature gradient modes. Physics of Plasmas 8 (9), 4096–4104.
  • Johnson (2024) Johnson, Steven G. 2024 Faddeeva w function implementation. http://ab-initio.mit.edu/Faddeeva.
  • Kadomtsev & Pogutse (1970a) Kadomtsev, BB & Pogutse, OP 1970a Turbulence in toroidal systems. In Reviews of Plasma Physics: Volume 5, pp. 249–400. Springer.
  • Kadomtsev & Pogutse (1970b) Kadomtsev, B. B. & Pogutse, O. P. 1970b Reviews of Plasma Physics/Voprosy Teorii Plazmy. Consultants Bureau.
  • Kuroda et al. (1998) Kuroda, Tohru, Sugama, Hideo, Kanno, Ryutaro, Okamoto, Masao & Horton, Wendell 1998 Initial value problem of the toroidal ion temperature gradient mode. Journal of the Physical Society of Japan 67 (11), 3787–3787.
  • Mandell et al. (2018) Mandell, NR, Dorland, W & Landreman, M 2018 Laguerre–hermite pseudo-spectral velocity formulation of gyrokinetics. Journal of Plasma Physics 84 (1), 905840108.
  • Mercier (1962) Mercier, C 1962 Critère de stabilité d’un système toroïdal hydromagnétique en pression scalaire. Nucl. Fusion Suppl 2, 801.
  • Nührenberg & Zille (1988) Nührenberg, J. & Zille, R. 1988 Quasi-helically symmetric toroidal stellarators. Physics Letters A 129 (2), 113 – 117.
  • Parisi et al. (2022) Parisi, Jason F, Parra, FI, Roach, Colin M, Hardman, Michael R, Schekochihin, AA, Abel, IG, Aiba, Nobuyuki, Ball, Justin, Barnes, Michael, Chapman-Oplopoiou, Benjamin & others 2022 Three-dimensional inhomogeneity of electron-temperature-gradient turbulence in the edge of tokamak plasmas. Nuclear Fusion 62 (8), 086045.
  • Plunk et al. (2014) Plunk, GG, Helander, P, Xanthopoulos, P & Connor, JW 2014 Collisionless microinstabilities in stellarators. iii. the ion-temperature-gradient mode. Physics of Plasmas 21 (3), 032112.
  • Podavini et al. (2023) Podavini, L., Zocco, A., García-Regaña, J. M., Barnes, M., Parra, F. I., Mishchenko, A. & Helander, P. 2023 Electrostatic microinstabilities and turbulence in Wendelstein 7-X close to the stability threshold, arXiv: 2311.04342.
  • Qiu et al. (2018) Qiu, Zhiyong, Chen, Liu & Zonca, Fulvio 2018 Kinetic theory of geodesic acoustic modes in toroidal plasmas: a brief review. Plasma Science and Technology 20 (9), 094004.
  • Roberg-Clark et al. (2022a) Roberg-Clark, GT, Plunk, GG & Xanthopoulos, P 2022a Coarse-grained gyrokinetics for the critical ion temperature gradient in stellarators. Physical Review Research 4 (3), L032028.
  • Roberg-Clark et al. (2022b) Roberg-Clark, GT, Xanthopoulos, P & Plunk, GG 2022b Reduction of electrostatic turbulence in a quasi-helically symmetric stellarator via critical gradient optimization. arXiv preprint arXiv:2210.16030 .
  • Rodriguez et al. (2020) Rodriguez, E., Helander, P. & Bhattacharjee, A. 2020 Necessary and sufficient conditions for quasisymmetry. Physics of Plasmas 27 (6), 062501.
  • Rodriguez & Zocco (2024) Rodriguez, E. & Zocco, A. 2024 .
  • Romanelli (1989) Romanelli, F. 1989 Ion temperature-gradient-driven modes and anomalous ion transport in tokamaks. Phys. Fluids B 1 (5), 1018.
  • Rudakov & Sagdeev (1961) Rudakov, L. I. & Sagdeev, R. Z. 1961 O neustoychivosti neodnorodnoy razrezhennoy plazmyi v silnom magnitnom pole. Dokl. Akad. Nauk CCCP 138.
  • Smolyakov et al. (2002) Smolyakov, AI, Yagi, M & Kishimoto, Y 2002 Short wavelength temperature gradient driven modes in tokamak plasmas. Physical review letters 89 (12), 125005.
  • Sugama (1999) Sugama, Hideo 1999 Dam** of toroidal ion temperature gradient modes. Physics of Plasmas 6 (9), 3527–3535.
  • Sugama & Watanabe (2006a) Sugama, H. & Watanabe, T.-H. 2006a Collisionless dam** of geodesic acoustic modes. Journal of Plasma Physics 72 (6), 825–828.
  • Sugama & Watanabe (2006b) Sugama, H. & Watanabe, T.-H. 2006b Collisionless dam** of zonal flows in helical systems. Physics of Plasmas 13 (1), 012501.
  • Tang et al. (1980) Tang, W.M., Connor, J.W. & Hastie, R.J. 1980 Nucl. Fusion 20 (11), 1439.
  • Taylor (1976) Taylor, J. B. 1976 Plasma Phys. and Control. Nuclear Fusion Research 2.
  • Terry et al. (1982) Terry, P., Anderson, W. & Horton, W. 1982 Kinetic effects on the toroidal ion pressure gradient drift mode. Nucl. Fusion 22 (4), 487.
  • Tricomi (1985) Tricomi, F 1985 Integral Equations. Dover.
  • Usmani (1994) Usmani, RA 1994 Inversion of jacobi’s tridiagonal matrix. Computers & Mathematics with Applications 27 (8), 59–66.
  • Wesson & Campbell (2011) Wesson, John & Campbell, David J 2011 Tokamaks, , vol. 149. Oxford university press.
  • Woods (1926) Woods, Frederick Shenstone 1926 Advanced calculus: a course arranged with special reference to the needs of students of applied mathematics. Ginn.
  • Zocco et al. (2016) Zocco, A, Plunk, GG, Xanthopoulos, P & Helander, P 2016 Geometric stabilization of the electrostatic ion-temperature-gradient driven instability. i. nearly axisymmetric systems. Physics of Plasmas 23 (8), 082516.
  • Zocco et al. (2022) Zocco, Alessandro, Podavini, Linda, Garcìa-Regaña, José Manuel, Barnes, Michael, Parra, Felix I., Mishchenko, A. & Helander, Per 2022 Gyrokinetic electrostatic turbulence close to marginality in the wendelstein 7-x stellarator. Phys. Rev. E 106, L013202.
  • Zocco et al. (2018) Zocco, A., Xanthopoulos, P., Doerk, H., Connor, J. W. & Helander, P. 2018 Journal of Plasma Physics 84 (1), 715840101.
  • Zonca et al. (1996) Zonca, Fulvio, Chen, Liu & Santoro, Robert A 1996 Kinetic theory of low-frequency alfvén modes in tokamaks. Plasma Physics and Controlled Fusion 38 (11), 2011.