Hostname: page-component-745bb68f8f-grxwn Total loading time: 0 Render date: 2025-02-06T08:49:49.234Z Has data issue: false hasContentIssue false

Lower hybrid current drive in a tokamak for correlated passes through resonance

Published online by Cambridge University Press:  10 June 2021

Peter J. Catto*
Affiliation:
Plasma Science and Fusion Center, Massachusetts Institute of Technology, Cambridge, MA, USA
*
 Email address for correspondence: catto@psfc.mit.edu
Rights & Permissions [Opens in a new window]

Abstract

Standard quasilinear descriptions are based on the constant magnetic field form of the quasilinear operator so improperly treat the trapped electron modifications associated with tokamak geometry. Moreover, successive poloidal transits of the Landau resonance during lower hybrid current drive in a tokamak are well correlated, and these geometrical details must be properly retained to account for the presence of trapped electrons that do not contribute to the driven current. The recently derived quasilinear operator in tokamak geometry accounts for these features and finds that the quasilinear diffusivity is proportional to a delta function with a transit or bounce averaged argument (rather than a local Landau resonance condition). The new quasilinear operator is combined with the Cordey (Nucl. Fusion, vol. 16, 1976, pp. 499–507) eigenfunctions to properly derive a rather simple and compact analytic expression for the trapped electron modifications to the driven lower hybrid current and the efficiency of the current drive.

Type
Research Article
Copyright
Copyright © The Author(s), 2021. Published by Cambridge University Press

1. Introduction

The original proposals to use radio frequency waves (rf) (Fisch Reference Fisch1978; Karney & Fisch Reference Karney and Fisch1979) and early estimates (Fisch & Boozer Reference Fisch and Boozer1980; Fisch & Karney Reference Fisch and Karney1981) and analytic evaluations (Antonsen & Chu Reference Antonsen and Chu1982; Cordey, Edlington & Start Reference Cordey, Edlington and Start1982; Taguchi Reference Taguchi1983) relied on the constant magnetic field treatment of quasilinear theory (Kennel & Engelmann Reference Kennel and Engelmann1966). Indeed, all subsequent analytic treatments (Karney & Fisch Reference Karney and Fisch1985; Yoshioka & Antonsen Reference Yoshioka and Antonsen1986; Cohen Reference Cohen1987; Giruzzi Reference Giruzzi1987; Chiu et al. Reference Chiu, Chan, Harvey and Porkolab1989; Ehst & Karney Reference Ehst and Karney1991) of the amount of parallel electron current that can be driven with lower hybrid waves in a tokamak continued to employ the Kennel--Engelmann quasilinear (QL) operator. More recent work, starting in the late 1980s, focused on numerical simulations of lower hybrid current drive (LHCD) and is the subject of an extensive review by Bonoli (Reference Bonoli2014). These treatments work quite well and also lead to sensible estimates of the LHCD efficiency even though tokamak geometry is not properly retained. Recently, a QL description that properly accounts for the correlated nature of successive poloidal passes through the Landau damping resonance in a tokamak has been derived (Catto & Tolman Reference Catto and Tolman2021). In this new formulation the wave--particle resonance condition is a transit or bounce averaged resonance condition in velocity space, rather than a resonance at a local poloidal angle in the torus. This improved QL operator is used herein to discover the changes that arise because of this subtle difference, and to demonstrate how only minor changes are required in standard treatments to properly account for toroidal geometry and the presence of trapped electrons.

In the next section, the collision operators employed are given and the adjoint procedure of Antonsen & Chu (Reference Antonsen and Chu1982) slightly extended. Section 3 presents a solution of the adjoint kinetic equation for the electrons by using the Sturm--Liouville eigenfunctions originally introduced by Cordey (Reference Cordey1976) -- a procedure that avoids the limitations of employing model collision operators that only retain pitch angle scattering collisions (Taguchi Reference Taguchi1983) or a Legendre equation approximation to the pitch angle scattering operator when energy diffusion is retained (Cordey et al. Reference Cordey, Edlington and Start1982; Karney & Fisch Reference Karney and Fisch1985; Cohen Reference Cohen1987). The final solution is rather simple and compact in form because only a single Cordey eigenfunction is required. It is used in § 4 to evaluate the parallel current that can be driven by lower hybrid waves. The resulting expression properly retains for the first time the electron trapping modifications that reduce the amount of driven current. In § 5 the power needed to drive the current is used to evaluate the current drive efficiency with these modifications due to the presence of trapped electrons as well as poloidal and radial mode structure effects retained. The discussion in § 6 also gives an estimate of the density at which nonlinear effects are expected to enter as well as a summary.

2. Adjoint method and collision operators

Using Gaussian cgs units, and the drift kinetic variables of spatial location $\boldsymbol{r}$, total energy $E = {v^2}/2 - e\Phi /{m_e}$, magnetic moment $\mu = v_ \bot ^2/2B$ and gyrophase $\varphi$, such that the velocity is

(2.1)\begin{equation}\boldsymbol{v} = {\boldsymbol{v}_ \bot } + {v_{||}}\boldsymbol{n} = {v_ \bot }[{\boldsymbol{e}_1}(\boldsymbol{r})\cos \varphi + {\boldsymbol{e}_2}(\boldsymbol{r})\sin \varphi ] + {v_{||}}\boldsymbol{n}(\boldsymbol{r}),\end{equation}

with $v_{||}^2 = {v^2}{\kern 1pt} - 2\mu B$, the QL equation for the unperturbed electron distribution function f is

(2.2)\begin{equation}{v_{||}}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }f = C\{ f\} + Q\{ f\} .\end{equation}

In the preceding, Q denotes the QL operator, C is the collision operator, the three orthonormal spatial unit vectors are related by ${\boldsymbol{e}_1} \times {\boldsymbol{e}_2}{\kern 1pt} = \boldsymbol{n} = \boldsymbol{B}/B$, and the unperturbed magnetic field is

(2.3)\begin{equation}\boldsymbol{B} = B\boldsymbol{n} = \boldsymbol{\nabla }\alpha \times \boldsymbol{\nabla }\psi = I(\psi )\boldsymbol{\nabla }\zeta + \boldsymbol{\nabla }\zeta \times \boldsymbol{\nabla }\psi ,\end{equation}

where ${m_e}$ and e are the mass and magnitude of the charge on an electron, and $\Phi $ is the electrostatic potential. The spatial variables employed are the poloidal flux function $\psi$, and the poloidal, $\vartheta$, and toroidal, $\zeta$, angles, where q is the safety factor and $I = R{B_t}$, with ${B_t}$ the toroidal magnetic field and R the major radius. The detailed forms of Q and C will be given as needed.

Linearizing about an electron Maxwellian of density ${n_e}$ and temperature ${T_e}$,

(2.4)\begin{equation}{f_0} = {f_0}(\psi ,E) = {n_e}(\psi ){\left[ {\frac{{{m_e}}}{{2{\rm \pi}{T_e}(\psi )}}} \right]^{3/2}}{\textrm{e}^{ - {m_e}{v^2}/2{T_e}(\psi )}} = {n_e}(\psi ){\left[ {\frac{{{m_e}}}{{2{\rm \pi}{T_e}(\psi )}}} \right]^{3/2}}{\textrm{e}^{ - [{m_e}E + e\Phi (\psi )]/{T_e}(\psi )}},\end{equation}

by writing

(2.5)\begin{equation}f = {f_0} + {f_1} + \cdots ,\end{equation}

and using

(2.6)\begin{equation}C\{ f\} = {C_{ee}}\{ f\} + {C_{ei}}\{ f\} ,\end{equation}

with $C\{ {f_0}\} = 0$, and ${C_{ee}}$ and ${C_{ei}}$ the electron--electron and electron--ion collision operators, leads to the unperturbed linearized equation to be solved by the adjoint method, namely

(2.7)\begin{equation}{v_{||}}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }{f_1} = C\{ {f_1}\} + Q\{ {f_0}\} .\end{equation}

The electron--ion collision operator (Hinton & Hazeltine Reference Hinton and Hazeltine1976; Helander & Sigmar Reference Helander and Sigmar2002) is

(2.8)\begin{equation}{C_{ei}}\{ {f_1}\} = \frac{{{\nu _u}}}{{{x^3}}}L\left\{ {{f_1} - \frac{{{m_e}}}{{{T_e}}}{V_{||}}{v_{||}}{f_0}} \right\} = \frac{{{\nu _u}}}{{{x^3}}}\left[ {L\{ {f_1}\} + \frac{{{m_e}}}{{{T_e}}}{V_{||}}{v_{||}}{f_0}} \right],\end{equation}

with $x = v/{v_e}$ and ${v_e} = {(2{T_e}/{m_e})^{1/2}}$ the electron thermal speed. The Lorentz operator L is self-adjoint and defined here as

(2.9)\begin{equation}L\{ h\} = \frac{1}{2}{\nabla _v} \cdot [({v^2}\boldsymbol{I} - \boldsymbol{vv})\boldsymbol{\cdot }{\boldsymbol{\nabla }_v}h] = \frac{{2{B_0}}}{B}\xi \frac{\partial }{{\partial \lambda }}\left( {\lambda \xi \frac{{\partial h}}{{\partial \lambda }}} \right),\end{equation}

with $L\{ {v_{\|}}{f_0}\} = - {v_{\|}}{f_0}$, $\lambda = 2\mu {B_0}{\kern 1pt} /{\kern 1pt} {v^2} = {B_0}v_ \bot ^2/B{v^2}$, $\xi = {v_{\|}}/v$ and ${B_0}$ a normalization constant to be made explicit shortly. The parallel ion mean velocity is ${V_{\|}}$ and the unlike collision frequency is

(2.10)\begin{equation}{\nu _u} \equiv \sqrt 2 {\rm \pi}{Z^2}{e^4}{n_i}\ell n\varLambda /m_e^{1/2}T_e^{3/2} \to 3\sqrt {\rm \pi}{\nu _{ei}}/4,\end{equation}

where ${\nu _{ei}} = 4\sqrt {2{\rm \pi}} {Z^2}{e^4}{n_i}\ell n\varLambda /3{m^{1/2}}{T^{3/2}} = Z{\nu _{ee}}$ for a quasineutral plasma with $Z{n_i} = {n_e}$, ${\nu _{ee}}$ the electron--electron collision frequency, and Z the ion charge number.

The electron--electron collision operator is self-adjoint. As only a lowest-order solution is desired, the standard high speed expansion of the electron--electron collision operator in its self-adjoint form is employed, namely

(2.11)\begin{align}{C_{ee}}\{ h\} &= {\nu _\ell }\left\{ {\frac{1}{{{x^3}}}L\{ h\} + {\nabla_v} \cdot \left[ {\frac{{{T_e}{f_0}}}{{{m_e}{x^3}}}{\nabla_v}\left( {\frac{h}{{{f_0}}}} \right)} \right]} \right\}\nonumber\\ &= \frac{{2{B_0}{\nu _\ell }\xi }}{{B{x^3}}}\frac{\partial }{{\partial \lambda }}\left( {\lambda \xi \frac{{\partial h}}{{\partial \lambda }}} \right) + \frac{{{\nu _\ell }{T_e}}}{{{m_e}{v^2}}}\frac{\partial }{{\partial v}}\left[ {\frac{{{v^2}{f_0}}}{{{x^3}}}\frac{\partial }{{\partial v}}\left( {\frac{h}{{{f_0}}}} \right)} \right],\end{align}

where

(2.12)\begin{equation}{\nu _\ell } \equiv \sqrt 2 {\rm \pi}{Z^4}{e^4}{n_e}\ell n\varLambda /m_e^{1/2}T_e^{3/2} \to 3\sqrt {\rm \pi}{\nu _{ee}}/4,\end{equation}

where ${\nu _{ee}} = 4\sqrt {2{\rm \pi}} {e^4}{n_e}\ell n\varLambda /3{m^{1/2}}{T^{3/2}}$ and h is a perturbed distribution function. The preceding like particle collision operator is just the usual non-momentum conserving high speed expansion of the Rosenbluth potentials for collisions with a Maxwellian (Karney & Fisch Reference Karney and Fisch1979).

The adjoint method provides an explicit means of evaluating the parallel current provided the adjoint equation is easier to solve than the original kinetic equation. It has been used for LHCD by Antonsen & Chu (Reference Antonsen and Chu1982), Karney & Fisch (Reference Karney and Fisch1985) and Cohen (Reference Cohen1987). It has also been used to calculate the bootstrap current in stellarators (Helander, Geiger & Maaβberg Reference Helander, Geiger and Maaßberg2011), and in the plateau regime to evaluate the bootstrap current in a tokamak (Pusztai & Catto Reference Pusztai and Catto2010).

To account for the non-self-adjointness of the electron--ion collision operator it is necessary to use the modified adjoint equation

(2.13)\begin{equation}{v_{||}}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }h + C\{ h\} ={-} \left( {\frac{B}{I} - \frac{{m{\nu_u}}}{{T{x^3}}}{V_{||}}} \right){v_{||}}{f_0}.\end{equation}

Then, defining the flux surface average of any quantity A as

(2.14)\begin{equation}\langle A\rangle \equiv {{[{\oint {\textrm{d}\vartheta A/\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta } } ]} / {[{\oint {\textrm{d}\vartheta /\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta } } ]}},\end{equation}

using

(2.15)\begin{equation}\left\langle {\int {{\textrm{d}^3}v} f_0^{ - 1}{v_{||}}(h\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }{f_1} + {f_1}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }h)} \right\rangle = \left\langle {\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\left( {{B^{ - 1}}\int {{\textrm{d}^3}} v{v_{||}}h{f_1}/{f_0}} \right)} \right\rangle = 0,\end{equation}

and the self-adjointness of the electron--electron collision operator,

(2.16)\begin{equation}\left\langle {\int {{\textrm{d}^3}v} f_0^{ - 1}({f_1}{C_{ee}}\{ h\} - h{C_{ee}}\{ {f_1}\} )} \right\rangle = 0,\end{equation}

and the Lorentz operator

(2.17)\begin{equation}\left\langle {\int {{\textrm{d}^3}v} f_0^{ - 1}x_e^{ - 3}({f_1}L\{ h\} - hL\{ {f_1}\} )} \right\rangle = 0,\end{equation}

yields the desired adjoint relation

(2.18)\begin{equation}\left\langle {B\int {{\textrm{d}^3}} v{v_{||}}{f_1}} \right\rangle = I\left\langle {\int {{\textrm{d}^3}} vh\left( {f_0^{ - 1}Q\{ {f_0}\} - \frac{{m{\kern 1pt} {\nu_u}}}{{T{\kern 1pt} x_e^3}}{V_{||}}{v_{||}}} \right)} \right\rangle .\end{equation}

Therefore, only the adjoint equation need be solved to evaluate the parallel electron current driven by the lower hybrid waves. It is convenient to define $B_0^2 \equiv \langle {B^2}\rangle$.

Based on the direction of the poloidal magnetic field, the Ohmic current is in the positive toroidal direction. Consequently, the lower hybrid (LH) parallel electron flow is to be driven in the opposite or negative direction to make $\left\langle {B\int {{\textrm{d}^3}} v{v_{\|}}{f_1}} \right\rangle < 0$.

The parallel ion flow term in (2.18) is normally ignored as

(2.19)\begin{equation}\frac{{m{\nu _u}\left\langle {{V_{||}}\int {{\textrm{d}^3}} v{x^{ - 3}}h{v_{||}}} \right\rangle }}{{T\left\langle {\int {{\textrm{d}^3}} vhf_0^{ - 1}Q\{ {f_0}\} } \right\rangle }}\sim \frac{{{V_{||}}{\nu _{ee}}{f_0}}}{{{v_e}Q}}\sim {\left( {\frac{{{m_e}}}{M}} \right)^{1/2}}\frac{{{\rho _{pi}}}}{a}\frac{{{\nu _{ee}}{f_0}}}{{Q\{ {f_0}\} }} \ll 1,\end{equation}

where M is the ion mass, and ${V_{\|}}\sim{v_i}{\rho _{pi}}/a$, with ${v_i}$ the ion thermal speed, ${\rho _{pi}}$the poloidal ion gyroradius and a the minor radius. Consequently, ${V_{\|}}$ can be ignored and only

(2.20)\begin{align}\left\langle {B\int {{\textrm{d}^3}} v{v_{||}}{f_1}} \right\rangle = I\int {{\textrm{d}^3}} v\frac{{{v_{||}}h}}{{B{f_0}}}\left\langle {\frac{B}{{{v_{||}}}}Q\{ {f_0}\} } \right\rangle = I{{\left( {\int {{\textrm{d}^3}} v\frac{{{v_{||}}h}}{{B{f_0}}}{\kern 1pt} {\tau_f}\overline {Q\{ {f_0}\} } } \right)} / {({\oint {\textrm{d}\vartheta /\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta } } )}},\end{align}

need be evaluated, where the solution for h is shown in the next section to satisfy $\partial h/{\kern 1pt} \partial \vartheta = 0$ to lowest order. The transit average over a full (f) poloidal circuit is defined using $\textrm{d}\tau = \textrm{d}\vartheta /{v_{\|}}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta$ to be

(2.21)\begin{equation}\bar{A} \equiv {{\oint_f {\textrm{d}\tau A} } / {({\oint_f {\textrm{d}\tau } } )}}.\end{equation}

3. Eigenfunction solution procedure in tokamak geometry

Rewriting the adjoint equation yields a Spitzer & Härm (Reference Spitzer and Härm1953) equation in toroidal geometry,

(3.1)\begin{equation}{v_{||}}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }h + {C_{ee}}\{ h\} + {\nu _u}{x^{ - 3}}L\{ h\} ={-} {I^{ - 1}}B{v_{||}}{f_0}.\end{equation}

Writing $h = \bar{h} + \tilde{h} + \cdots$ with $\tilde{h} \ll \bar{h}$ then to lowest order

(3.2)\begin{equation}{v_{||}}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }\bar{h} = 0,\end{equation}

giving

(3.3)\begin{equation}\bar{h} = \bar{h}(\psi ,\alpha ,E,\mu ,\sigma ),\end{equation}

where $\sigma = {v_{\|}}/| {v_{\|}} |$ for the passing and $\sigma = 0$ for the trapped. Then transit averaging the next order equation,

(3.4)\begin{equation}{v_{||}}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }\tilde{h} + {C_{ee}}\{ \bar{h}\} + {\nu _u}{x^{ - 3}}L\{ \bar{h}\} ={-} {I^{ - 1}}B{v_{||}}{f_0},\end{equation}

leads to

(3.5)\begin{equation}\overline {{C_{ee}}\{ \bar{h}\} } + {\nu _u}{x^{ - 3}}\overline {L\{ \bar{h}\} } ={-} {I^{ - 1}}\overline {B{v_{||}}} {f_0}.\end{equation}

Integration over a full bounce for the trapped (t) electrons gives $\overline {B{v_{\|}}} = 0$, implying that the trapped response ${\bar{h}_t}$ vanishes,

(3.6)\begin{equation}{\bar{h}_t} = 0.\end{equation}

For the passing electrons

(3.7)\begin{equation}\overline {B{v_{||}}} \oint_f {\textrm{d}\tau } = \oint {\textrm{d}\vartheta {B^2}/\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta } = \langle {B^2}\rangle \oint {\textrm{d}\vartheta /\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta } .\end{equation}

Using the flux surface average to rewrite the passing (p) adjoint equation leads to

(3.8)\begin{equation}I\left\langle {\frac{B}{{{v_{||}}}}{C_{ee}}\{ {{\bar{h}}_p}\} } \right\rangle + \frac{{I{\nu _u}}}{{x_e^3}}\left\langle {\frac{B}{{{v_{||}}}}L\{ {{\bar{h}}_p}\} } \right\rangle ={-} \langle {B^2}\rangle {f_0},\end{equation}

where

(3.9)\begin{equation}\left\langle {\frac{B}{{{v_{||}}}}L\{ {{\bar{h}}_p}\} } \right\rangle = \frac{{2{B_0}}}{v}\frac{\partial }{{\partial \lambda }}\left( {\lambda \langle \xi \rangle \frac{{\partial {{\bar{h}}_p}}}{{\partial \lambda }}} \right),\end{equation}

and

(3.10)\begin{equation}\left\langle {\frac{B}{{{v_{||}}}}{C_{ee}}\{ {{\bar{h}}_p}\} } \right\rangle = \frac{{2{B_0}{\nu _\ell }}}{{{x^3}v}}\frac{\partial }{{\partial \lambda }}\left( {\lambda \langle \xi \rangle \frac{{\partial {{\bar{h}}_p}}}{{\partial \lambda }}} \right) + \frac{{{\nu _\ell }{T_e}}}{{{m_e}{v^3}}}\left\langle {\frac{B}{\xi }} \right\rangle \frac{\partial }{{\partial v}}\left[ {\frac{{{v^2}{f_0}}}{{{x^3}}}\frac{\partial }{{\partial v}}\left( {\frac{{{{\bar{h}}_p}}}{{{f_0}}}} \right)} \right].\end{equation}

The last term of (3.10) proportional to ${v^{ - 3}}\partial {\bar{h}_p}/\partial v$ is drag and the remainder is energy scattering.

In the large aspect ratio limit

(3.11)\begin{equation}\langle \xi \rangle = \frac{{2\sqrt {2\varepsilon } E(k)}}{{{\rm \pi} \sqrt {(1 - \varepsilon ){k^2} + 2\varepsilon } }},\end{equation}

and

(3.12)\begin{equation}\frac{1}{{{B_0}}}\left\langle {\frac{B}{\xi }} \right\rangle ={-} 2\frac{{\partial \langle \xi \rangle }}{{\partial \lambda }},\end{equation}

with E an elliptic integral of the second kind, ${k^2} = 2\varepsilon \lambda /[1 - (1 - \varepsilon )\lambda ]$, and $\varepsilon = r/R$ with r the minor radius. The equation to be solved is first rewritten as

(3.13)\begin{align}2(Z + 1)\frac{\partial }{{\partial \lambda }}\left[ {\lambda \langle \xi \rangle \frac{\partial }{{\partial \lambda }}\left( {\frac{{{{\bar{h}}_p}}}{{{f_0}}}} \right)} \right] + \frac{{{T_e}{x^3}}}{{{m_e}{B_0}{v^2}{f_0}}}\left\langle {\frac{B}{\xi }} \right\rangle \frac{\partial }{{\partial v}}\left[ {\frac{{{v^2}{f_0}}}{{{x^3}}}\frac{\partial }{{\partial v}}\left( {\frac{{{{\bar{h}}_p}}}{{{f_0}}}} \right)} \right] ={-} \frac{{\langle {B^2}\rangle v{x^3}}}{{I{B_0}{\nu _\ell }}}.\end{align}

Only the lowest -order $v$ dependence is required. Inserting

(3.14)\begin{equation}\frac{{{{\bar{h}}_p}}}{{{f_0}}} = \frac{{\langle {B^2}\rangle v{x^3}}}{{I{B_0}{\nu _\ell }}}\varLambda (\lambda ),\end{equation}

leads to

(3.15)\begin{equation}2(Z + 1)\frac{\partial }{{\partial \lambda }}\left( {\lambda \langle \xi \rangle \frac{{\partial \varLambda }}{{\partial \lambda }}} \right) + \frac{{{T_e}\varLambda }}{{{m_e}{B_0}{v^3}{f_0}}}\left\langle {\frac{B}{\xi }} \right\rangle \frac{\partial }{{\partial v}}\left[ {\frac{{{v^2}{f_0}}}{{{x^3}}}\frac{\partial }{{\partial v}}(v{x^3})} \right] ={-} 1.\end{equation}

Using

(3.16)\begin{align}\frac{{{T_e}}}{{{m_e}{v^3}{f_0}}}\frac{\partial }{{\partial v}}\left[ {\frac{{{v^2}{f_0}}}{{{x^3}}}\frac{\partial }{{\partial v}}(v{x^3})} \right] = \frac{1}{{2{x^3}{e^{ - {x^2}}}}}\frac{\partial }{{\partial x}}\left( {\frac{{{e^{ - {x^2}}}}}{x}\frac{{\partial {x^4}}}{{\partial x}}} \right) = \frac{2}{{{x^3}{e^{ - {x^2}}}}}\frac{\partial }{{\partial x}}({x^2}{e^{ - {x^2}}}) ={-} 4 + \frac{4}{{{x^2}}},\end{align}

then for $x \gg 1$ drag dominates over energy scattering and the equation to be solved reduces to

(3.17)\begin{equation}\frac{\partial }{{\partial \lambda }}\left( {\lambda \langle \xi \rangle \frac{{\partial \varLambda }}{{\partial \lambda }}} \right) + \frac{4}{{Z + 1}}\frac{{\partial \langle \xi \rangle }}{{\partial \lambda }}\varLambda ={-} \frac{1}{{2(Z + 1)}}.\end{equation}

The Cordey (Reference Cordey1976) eigenfunctions ${\varLambda _j}$ and associated eigenvalues ${\kappa _j}$ of the Sturm--Liouville differential equation

(3.18)\begin{equation}\frac{\partial }{{\partial \lambda }}\left( {\lambda \langle \xi \rangle \frac{{\partial {\Lambda_j}}}{{\partial \lambda }}} \right) = {\kappa _j}\frac{{\partial \langle \xi \rangle }}{{\partial \lambda }}{\varLambda _j} ={-} \frac{{{\kappa _j}}}{2}\left\langle {\frac{B}{{{B_0}\xi }}} \right\rangle {\varLambda _j},\end{equation}

are used to obtain a solution. Expanding in the eigenfunctions ${\varLambda _j}$, which satisfy ${\varLambda _j}(\lambda = 0) = 1$, ${\varLambda _j}(\lambda = {B_0}/{B_{\textrm{max}}}) = 0$ and, for $j \ne k$, the orthogonality condition

(3.19)\begin{equation}\int_0^{{B_0}/{B_{max}}} {\textrm{d}\lambda } {\varLambda _k}{\varLambda _j}\frac{{\partial \langle \xi \rangle }}{{\partial \lambda }} = 0,\end{equation}

by inserting

(3.20)\begin{equation}\varLambda = \sum\limits_{j = 1}^\infty {{A_j}{\varLambda _j}} ,\end{equation}

into the differential equation leads to

(3.21)\begin{equation}\sum\limits_{j = 1}^\infty {{A_j}\left( {{\kappa_j} + \frac{4}{{Z + 1}}\frac{{\partial \langle \xi \rangle }}{{\partial \lambda }}} \right){\varLambda _j}} ={-} \frac{1}{{2(Z + 1)}}.\end{equation}

Multiplying by ${\varLambda _k}$ and integrating over $\lambda$, yields the coefficients ${A_j}$ to be given by

(3.22)\begin{equation}2{\alpha _j}[(Z + 1){\kappa _j} + 4]{A_j} = {\beta _j},\end{equation}

where

(3.23)\begin{equation}{\alpha _j} ={-} \int_0^{{B_0}/{B_{max}}} {\textrm{d}\lambda } \varLambda _j^2\frac{{\partial \langle \xi \rangle }}{{\partial \lambda }}\mathop \to \limits_{\varepsilon \to 0} \frac{1}{{4j - 1}},\end{equation}

and

(3.24)\begin{equation}{\beta _j} = \int_0^{{B_0}/{B_{max}}} {\textrm{d}\lambda } {\varLambda _j}\mathop \to \limits_{\varepsilon \to 0} \frac{2}{3}.\end{equation}

Fortunately, aspect ratio expansions of the preceding eigenvalues and coefficients are available (Hsu, Catto & Sigmar Reference Hsu, Catto and Sigmar1990; Xiao, Catto & Molvig Reference Xiao, Catto and Molvig2007; Parker & Catto Reference Parker and Catto2012). In particular, the lowest eigenvalue is ${\kappa _1} \simeq 1 + 1.46\sqrt \varepsilon + 1.48\varepsilon + 0.13{\varepsilon ^{3/2}}$ so always of order unity, with the others increasing in size as for all Sturm--Liouville problems. Moreover, as $\varepsilon$ increases so do all of the ${\kappa _j}$ and the ${\beta _j}/{\alpha _j}$. These prior investigations find that only the leading few eigenfunctions are required, as might be expected from how quickly the eigenvalues increase. Being more explicit by using results from Hsu et al. (Reference Hsu, Catto and Sigmar1990),

(3.25)\begin{align}\kappa _j - (2{j^2} - j) = 1.46\sqrt \varepsilon \frac{{4j - 1}}{3}\left[\frac{{(2j - 1)!!}}{{(2j - 2)!!}}\right]^2 = 1.46\sqrt \varepsilon \left\{\begin{array}{@{}ll} 1&{j = 1}\\ {21/4}&{j = 2}\\ {825/64}&{j = 3} \end{array} \right.,\end{align}

where $(2j - 1)!! = 1 \cdot 3 \cdot 5 \ldots (2j - 1)$ and $(2j - 2)!! = 2 \cdot 4 \cdot 6 \ldots (2j - 2)$ (and both equal 1 for j = 1). Consequently, only the first couple of ${\beta _j}$ and ${\alpha _j}$ are required. For ${\beta _j}$

(3.26)\begin{align}\begin{aligned} {\beta _j} & = \int_0^{{B_0}/{B_{\max }}} {\textrm{d}\lambda } {\varLambda _j}{\kern 1pt} = \dfrac{2}{3}(1 - \varepsilon )\left[ {{a_{j1}} - {{( - 1)}^{j + 1}}\dfrac{{(2j - 1)!!}}{{(2j - 2)!!}}\varepsilon } \right]\\ & \quad + O({\varepsilon ^{3/2}}) \to \dfrac{2}{3}(1 - \varepsilon )\left\{ {\begin{array}{@{}ll} {{a_{11}} - \varepsilon }&{j = 1}\\ {{a_{21}} + 3\varepsilon /2}&{j = 2} \end{array}} \right., \end{aligned}\end{align}

where based on appendix B and the fits in table 1 of Hsu et al. (Reference Hsu, Catto and Sigmar1990)

(3.27)\begin{equation}{a_{jj}} = 1 - \sum\limits_{j \ne m} {{a_{jm}}} \mathop \to \limits_{j = 1} {a_{11}} = 1 - 0.62\sqrt \varepsilon + 1.33\varepsilon ,\end{equation}

and

(3.28)\begin{align}\begin{aligned} {a_{jm}} & \simeq{-} \dfrac{{{{( - 1)}^{j + m}}(4m - 1)1.46\sqrt \varepsilon }}{{3[2{m^2} - m - (2{j^2} - j)]}}\left[ {\dfrac{{(2j - 1)!!(2m - 1)!!}}{{(2j - 2)!!(2m - 2)!!}}} \right]\\ & \quad + O(\varepsilon )\mathop \to \limits_{\scriptstyle j = 2 \atop \scriptstyle m = 1 } {a_{21}} ={-} 0.44\sqrt \varepsilon - 0.13\varepsilon . \end{aligned}\end{align}

In addition,

(3.29)\begin{align}{\alpha _j} &={-} \int_0^{{B_0}/{B_{\max }}} {\textrm{d}\lambda } \varLambda _j^2\frac{{\partial \langle \xi \rangle }}{{\partial \lambda }} = \sum\limits_{k = 1}^\infty {\frac{{a_{jk}^2}}{{4k - 1}} - \frac{4}{3}\varepsilon {{\left[ {\frac{{(2j - 1)!!}}{{(2j - 2)!!}}} \right]}^2}}\nonumber\\ &\quad + O({\varepsilon ^{3/2}}) = \left\{\begin{array}{@{}ll} {a_{11}^2/{\kern 1pt} 3 + O(\varepsilon )}&{j = 1}\\ {1/7 + O(\varepsilon )}&{j = 2} \end{array} \right.,\end{align}

giving

(3.30)\begin{equation}{\beta _1}/{\alpha _1} \simeq 2/{a_{11}} + O(\varepsilon ) = 2/(1 - 0.62\sqrt \varepsilon ) + O(\varepsilon ),\end{equation}

and

(3.31)\begin{equation}{\beta _2}/{\alpha _2} \simeq 14{a_{21}}{\kern 1pt} /3 + O(\varepsilon ) ={-} 2.04\sqrt \varepsilon + O(\varepsilon ).\end{equation}

Using the preceding results gives the lowest -order adjoint solution to be

(3.32)\begin{align} \frac{{{{\bar{h}}_p}}}{{{f_0}}} &= \frac{{\langle {B^2}\rangle v{x^3}}}{{2I{B_0}{\nu _\ell }}}\sum\limits_{j = 1}^\infty {\frac{{{\beta _j}{\varLambda _j}(\lambda )}}{{[(Z + 1){\kappa _j} + 4]{\alpha _j}}}}\nonumber\\ &\simeq \frac{{v{x^3}}}{{R{\nu _\ell }}}\left\{ {\frac{{(1 + 0.62\sqrt \varepsilon ){\Lambda_1}(\lambda )}}{{[(Z + 1)(1 + 1.46\sqrt {2\varepsilon } ) + 4]}} - \frac{{1.02\sqrt \varepsilon {\Lambda_2}(\lambda )}}{{[(Z + 1)(7 + 7.66\sqrt {2\varepsilon } ) + 4]}}} \right\},\end{align}

which will be used to evaluate the lower hybrid driven current. Notice that even the j = 2 term is small and can be ignored. The preceding solution also ignores ${x^{ - 3}} \ll 1$ corrections as small.

As ${\bar{h}_t}{\kern 1pt} = 0$ only the passing electrons contribute to 2.20, it simplifies to

(3.33)\begin{equation}\left\langle {B\int {{\textrm{d}^3}} v{v_{||}}{f_1}} \right\rangle \simeq \frac{{{B_0}I}}{{2{\rm \pi} qR}}\int {{\textrm{d}^3}} v\frac{{{v_{||}}{{\bar{h}}_p}}}{{B{f_0}}}{\tau _p}\overline {Q\{ {f_0}\} } ,\end{equation}

where large aspect ratio is assumed and the passing transit time for a full poloidal circuit is

(3.34)\begin{equation}{\tau _p} = \oint_p {\textrm{d}\tau } = \oint_p {\textrm{d}\vartheta /{v_{||}}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta } \simeq 4qR\sqrt {(1 - \varepsilon ){k^2} + 2\varepsilon } K(k)/v\sqrt {2\varepsilon } ,\end{equation}

with $\varepsilon = r/R$, ${k^2} = 2\varepsilon \lambda /[1 - (1 - \varepsilon )\lambda ]$, and K an elliptic integral of the first kind. The parallel current will next be formed by evaluating (3.33).

4. Current in a tokamak for a correlated QL diffusivity

The recently derived form for the correlated diffusivity (Catto & Tolman Reference Catto and Tolman2021) for the electrons in a tokamak is

(4.1)\begin{equation}\bar{D} = \frac{{2{{\rm \pi}^3}{q^2}{R^2}{e^2}|{\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}{|^2}}}{{m_e^2{v^2}{\tau _p}}}\sum\limits_\ell {\delta \left({\oint_p {\textrm{d}\tau \varLambda - 2{\rm \pi}\ell } } \right)\Theta (v,k)} ,\end{equation}

where the argument of the delta function is transit averaged with

(4.2)\begin{equation}\oint_p {\textrm{d}\tau \varLambda } \simeq \omega {\tau _p} - 2{\rm \pi}\sigma (qn - m),\end{equation}

allowing aspect ratio modifications to be properly evaluated. The phase integral $\Theta $ is defined as

(4.3)\begin{equation}\Theta \equiv {\left|{\oint_p {\frac{{\textrm{d}\vartheta }}{{2{\rm \pi}}}} \,{\textrm{e}^{ - i\int_{{\tau_0}}^\tau {\textrm{d}\tau^{\prime}\varLambda (\tau^{\prime})} }}} \right|^2} \le 1,\end{equation}

with

(4.4)\begin{equation}\int_{{\tau _0}}^\tau {\textrm{d}\tau ^{\prime}\varLambda } (\tau ^{\prime}) = \omega (\tau - {\tau _0}) - \sigma (qn - m)\vartheta (\tau ) = \omega qR\int_0^\vartheta {\textrm{d}\vartheta /{\kern 1pt} {v_{||}}} - \sigma (qn - m)\vartheta ,\end{equation}

where $\vartheta ({\tau _0}) = 0$ and

(4.5)\begin{equation}\textrm{d}\vartheta /\textrm{d}\tau = {v_{||}}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta .\end{equation}

For notational simplicity, only the poloidal mode number (m) is shown as a subscript on the Fourier coefficients of the applied rf parallel electric field ${\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}$. The frequency $(\omega )$, toroidal mode number (n) and radial mode index (s) subscripts are suppressed.

For a Maxwellian the transit averaged QL operator is

(4.6)\begin{equation}\overline {Q\{ {f_0}\} } = \sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s } {\frac{1}{{{\tau _p}}}\frac{\partial }{{\partial E}}\left( {{\tau_p}{v^2}\bar{D}\frac{{\partial {f_0}}}{{\partial E}}} \right)} = \sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,S } {\frac{1}{{{\tau _p}v}}{{\left. {\frac{\partial }{{\partial v}}} \right|}_\mu }} \left( {{\tau_p}v\bar{D}\frac{{\partial {f_0}}}{{\partial v}}} \right).\end{equation}

with $\ell$ the bounce harmonic index as sucessive passes are correlated. As only the passing contribute, the large aspect ratio form of the flow becomes

(4.7)\begin{align}\begin{aligned} \left\langle {B\int {{d^3}} v{v_{||}}{f_1}} \right\rangle & \simeq \dfrac{{{B_0}I}}{{2{\rm \pi} qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s } {{\textrm{d}^3}v} \dfrac{{{v_{||}}{{\bar{h}}_p}}}{{B{f_0}v}}{\left. {\dfrac{\partial }{{\partial v}}} \right|_\mu }\left( {v{\tau_p}\bar{D}{{\left. {\dfrac{{\partial {f_0}}}{{\partial v}}} \right|}_\mu }} \right)\\ & = \dfrac{{{m_e}{B_0}I}}{{2{\rm \pi}{T_e}qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s } {\int {{\textrm{d}^3}v} } \dfrac{{{v_{||}}{\tau _p}}}{{Bv}}\bar{D}{f_0}{\left. {\dfrac{\partial }{{\partial v}}} \right|_\mu }\left( {\dfrac{{{v^2}{{\bar{h}}_p}}}{{{f_0}}}} \right).\end{aligned}\end{align}

To proceed it is convenient to perform the v integral first by writing

(4.8)\begin{equation}\bar{D} = {v^{ - 2}}\tau _p^{ - 1}\sum\limits_\ell {{d_\ell }\Theta (v,k)\delta ({\oint_p {\textrm{d}\tau \varLambda - 2{\rm \pi}\ell } } )} ,\end{equation}

with

(4.9)\begin{equation}{d_\ell } \equiv 2{{\rm \pi}^3}{q^2}{R^2}{e^2}m_e^{ - 2}|{\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}{|^2}.\end{equation}

Then

(4.10)\begin{equation}\left\langle {B\int {{\textrm{d}^3}} v{v_{||}}{f_1}} \right\rangle = \frac{{{m_e}{B_0}I}}{{2{\rm \pi}{T_e}qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s,\ell } {{d_\ell }} \int {{\textrm{d}^3}} v\frac{{{v_{||}}{f_0}}}{{B{v^3}}}\Theta (v,k)\delta ({\oint_p {\textrm{d}\tau \varLambda - 2{\rm \pi}\ell } } ){\left. {\frac{\partial }{{\partial v}}} \right|_\mu }\left( {\frac{{{v^2}{{\bar{h}}_p}}}{{{f_0}}}} \right).\end{equation}

Taylor expanding the argument of the delta function leads to

(4.11)\begin{equation}\delta ({\oint_p {\textrm{d}\tau \varLambda - 2{\rm \pi}\ell } } )= \frac{{\delta (v - {v_k})}}{{\omega |\partial {\tau _p}/\partial v{\kern 1pt} {|_{{v_k}}}}} = \frac{{v_k^2\delta (v - {v_k})}}{{\omega v{\tau _p}}},\end{equation}

with the speed at resonance defined as

(4.12)\begin{equation}{v_k} \equiv \omega v{\tau _p}/2{\rm \pi}(\ell + |qn - m|).\end{equation}

As a result,

(4.13)\begin{equation}\left\langle {B\int {{\textrm{d}^3}} v{v_{||}}{f_1}} \right\rangle = \frac{{{m_e}{B_0}I}}{{2{\rm \pi} {T_e}qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s,\ell } {{d_\ell }} \int {{\textrm{d}^3}v} \frac{{{v_{||}}{f_0}}}{{Bv}}\frac{{\Theta (v,k)}}{{\omega v{\tau _p}}}\delta (v - {v_k}){\left. {\frac{\partial }{{\partial v}}} \right|_\mu }\left( {\frac{{{v^2}{{\bar{h}}_p}}}{{{f_0}}}} \right).\end{equation}

Using ${\nabla _v}v = \boldsymbol{v}/v$, ${\nabla _v}\lambda = (2{B_0}{\kern 1pt} /B{v^2}){\boldsymbol{v}_ \bot } - (2\lambda /{v^2})\boldsymbol{v}$, and ${\nabla _v}\varphi = v_ \bot ^{ - 2}\boldsymbol{n} \times \boldsymbol{v}$, gives ${\textrm{d}^3}v = \textrm{d}v\,\textrm{d}\lambda \textrm{d}\varphi /{\nabla _v}v \times {\nabla _v}\lambda \boldsymbol{\cdot }{\boldsymbol{\nabla }_v}\varphi = \textrm{d}v\,\textrm{d}\lambda \,\textrm{d}\varphi B{v^3}/{\kern 1pt} 2{B_0}{v_{\|}}$. The LH parallel electron flow must be driven in the negative direction to sustain the poloidal magnetic field, making $\left\langle {B\int {{\textrm{d}^3}} v{v_{\|}}{f_1}} \right\rangle < 0$. Hence, a minus sign must be inserted as the velocity space integration is only over ${v_{\|}} < 0$ electrons with $\bar{D}$ vanishing for ${v_{\|}} > 0$. Consequently, in the resonance condition, $\sigma = - 1$ and m > qn. Therefore, the magnitude of the parallel wave number $|{k_{\|}}| \equiv |qn - m|/qR$, is used as ${k_{\|}} < 0$.

Multiplying the velocity integral by 1/2 and inserting the diffusivity

(4.14)\begin{align}\begin{aligned} \left\langle {B\int {{\textrm{d}^3}} v{v_{||}}{f_1}} \right\rangle & = \dfrac{{ - {m_e}I}}{{4{T_e}qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s,\ell } {{d_\ell }} \int_0^{{B_0}/{B_{\textrm{max}}}} {\textrm{d}\lambda } \int_0^\infty {\textrm{d}v\Theta (v,k)\delta (v - {v_k})} \dfrac{{{v^2}{f_0}}}{{\omega v{\tau _p}}}{\left. {\dfrac{\partial }{{\partial v}}} \right|_\mu }\left( {\dfrac{{{v^2}{{\bar{h}}_p}}}{{{f_0}}}} \right)\\ & ={-} \dfrac{{{m_e}I}}{{4{T_e}qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s,\ell } {{d_\ell }} \int_0^{{B_0}/{B_{\textrm{max}}}} {\textrm{d}\lambda } \Theta ({v_k},k)\dfrac{{{v^2}{d_\ell }{f_0}}}{{\omega v{\tau _p}}}{\left. {\dfrac{\partial }{{\partial v}}} \right|_\mu }{\left. {\left( {\dfrac{{{v^2}{{\bar{h}}_p}}}{{{f_0}}}} \right)} \right|_{v = {v_k}}}. \end{aligned}\end{align}

The preceding equation is a convenient form of the parallel flow and can be used to evaluate the driven parallel electron current for a more general like particle collision operator than was considered in the preceding section.

Defining

(4.15)\begin{equation}x_k^2 = {m_e}v_k^2/2{T_e} = {m_e}{(\omega v{\tau _p})^2}/8{{\rm \pi}^2}{(\ell + |qn - m|)^2}{T_e},\end{equation}

with $x_k^2 \gg 1$ in ${f_0} \propto {\textrm{e}^{ - x_k^2}}$, the $v{\tau _p}{\kern 1pt} \propto K(k)\mathop \to \limits_{k \to 1} \infty$ indicates that the freely passing will dominate the driven current. Consequently, $x_k^2 \gg 1$ is used to integrate by parts by first noting

(4.16)\begin{equation}{\left. {\frac{{\partial {f_0}}}{{\partial \lambda }}} \right|_v} ={-} {f_0}{\left. {\frac{{\partial x_k^2}}{{\partial \lambda }}} \right|_v} ={-} {f_0}\frac{{{m_e}{\omega ^2}v{\tau _p}}}{{4{{\rm \pi}^2}{T_e}{{(\ell + |{\kern 1pt} qn - m|)}^2}}}{\left. {\frac{{\partial (v{\tau_p})}}{{\partial \lambda }}} \right|_v}.\end{equation}

Then ignoring $x_k^{ - 2} \ll 1$ corrections

(4.17) \begin{align}\left\langle {B\int {{\textrm{d}^3}} v{v_{||}}{f_1}} \right\rangle &={-} \frac{{{{\rm \pi}^2}I}}{{qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s,\ell } {\frac{{{d_\ell }{{(\ell + |{\kern 1pt} qn - m|)}^2}{f_0}}}{{\omega {{(\omega {\tau _p})}^2}\partial (v{\tau _p})/\partial \lambda }}} \Theta ({x_k},k = 0)\nonumber\\&\quad \times{\left. {\left[{{{\left. {\frac{\partial }{{\partial v}}} \right|}_\lambda }\left( {\frac{{{v^2}{{\bar{h}}_p}}}{{{f_0}}}} \right) - {{\left. {2\lambda v\frac{\partial }{{\partial \lambda }}} \right|}_v}\left( {\frac{{{{\bar{h}}_p}}}{{{f_0}}}} \right)} \right]} \right|_{\scriptstyle v = {v_k} \atop \scriptstyle\lambda = 0 = k }} + \cdots .\end{align}

Changing to v and $\lambda$ velocity variables using

(4.18)\begin{equation}{\left. {\frac{\partial }{{\partial v}}} \right|_\mu } = {\left. {\frac{\partial }{{\partial v}}} \right|_\lambda } + {\left. {\frac{{\partial \lambda }}{{\partial v}}} \right|_\mu }{\left. {\frac{\partial }{{\partial \lambda }}} \right|_{{v_{||}}}} = {\left. {\frac{\partial }{{\partial v}}} \right|_\lambda } - \frac{{2\lambda }}{v}{\left. {\frac{\partial }{{\partial \lambda }}} \right|_v},\end{equation}

gives

(4.19)\begin{equation}{\left. {\frac{\partial }{{\partial v}}} \right|_\mu }\left( {\frac{{{v^2}{{\bar{h}}_p}}}{{{f_0}}}} \right) = {\left. {\frac{\partial }{{\partial v}}} \right|_\lambda }\left( {\frac{{{v^2}{{\bar{h}}_p}}}{{{f_0}}}} \right) - {\left. {2\lambda v\frac{\partial }{{\partial \lambda }}} \right|_v}\left( {\frac{{{{\bar{h}}_p}}}{{{f_0}}}} \right).\end{equation}

Therefore,

(4.20) \begin{align} \left\langle {B\int {{\textrm{d}^3}} v{v_{||}}{f_1}} \right\rangle &={-} \frac{{{{\rm \pi}^2}I}}{{qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s,\ell } {\frac{{{d_\ell }{{(\ell + |{\kern 1pt} qn - m|)}^2}{f_0}}}{{\omega {{(\omega {\tau _p})}^2}\partial (v{\tau _p})/\partial \lambda }}} \Theta ({x_k},k = 0)\nonumber\\&\quad \times{\left. {\left[ {{{\left. {\frac{\partial }{{\partial v}}} \right|}_\lambda }\left( {\frac{{{v^2}{{\bar{h}}_p}}}{{{f_0}}}} \right) - {{\left. {2\lambda v\frac{\partial }{{\partial \lambda }}} \right|}_v}\left( {\frac{{{{\bar{h}}_p}}}{{{f_0}}}} \right)} \right]} \right|_{\scriptstyle v = {v_k} \atop \scriptstyle\lambda = 0 = k }} + \cdots .\end{align}

As the pitch angle derivative of ${\bar{h}_p}$ is well behaved at $\lambda = 0$ the last term vanishes leading to a lower hybrid driven current ${J_{\|}}$ of

(4.21) \begin{equation}{J_{||}} \equiv{-} eB_0^{ - 1}\left\langle {B\int {{\textrm{d}^3}} v{v_{||}}{f_1}} \right\rangle \simeq{-} \frac{{{{\rm \pi}^2}e}}{q}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s,\ell } {\frac{{{d_\ell }{{(\ell + |{\kern 1pt} qn - m|)}^2}{f_0}}}{{\omega {{(\omega {\tau _p})}^2}\partial (v{\tau _p})/\partial \lambda }}} \Theta ({x_k},k = 0){\left. {{{\left. {\frac{\partial }{{\partial v}}} \right|}_\lambda }\left( {\frac{{{v^2}{{\bar{h}}_p}}}{{{f_0}}}} \right)} \right|_{\scriptstyle v = {v_k} \atop \scriptstyle\lambda = 0 = k }}.\end{equation}

To simplify further

(4.22)\begin{equation}{(v{\tau _p})_{k = 0}}{\kern 1pt} = 2{\rm \pi} qR,\end{equation}

is used along with

(4.23) \begin{align}\frac{1}{{{\tau _p}}}{\left. {\frac{{\partial {\tau_p}}}{{{\kern 1pt} \partial \lambda }}} \right|_{\lambda = 0 = k}} &= \frac{{{{[(1 - \varepsilon ){k^2} + 2\varepsilon ]}^2}}}{{4\varepsilon {k^2}}}{\left. {\left[ {\frac{E}{{(1 - {k^2})K}} - K + \frac{{(1 - \varepsilon ){k^2}}}{{(1 - \varepsilon ){k^2} + 2\varepsilon }}} \right]} \right|_{k = 0}}\nonumber\\ &= \frac{\varepsilon }{{{k^2}}}\left[ {\frac{{{k^2}}}{2} + \frac{{(1 - \varepsilon ){k^2}}}{{2\varepsilon }}} \right] \simeq \frac{1}{2}.\end{align}

Moreover, only $\ell = 0$ contributes as

(4.24) \begin{align}\Theta (k = 0,{x_k} = {x_{k = 0}}) &\equiv {\left|{\oint_p {\frac{{\textrm{d}\vartheta }}{{2{\rm \pi}}}\,} {\textrm{e}^{ - \textrm{i}\int_{{\tau_0}}^\tau {\textrm{d}\tau^{\prime}} \varLambda (\tau^{\prime})}}} \right|^2}\nonumber\\ &\simeq {\left|{\oint_p {\frac{{\textrm{d}\vartheta }}{{2{\rm \pi}}}\,} {\textrm{e}^{ - \textrm{i}(\omega qR/v - |qn - m|)\vartheta }}} \right|^2} \simeq {\left|{\oint_p {\frac{{d\vartheta }}{{2{\rm \pi}}}\,} {\textrm{e}^{ - \textrm{i}\ell \vartheta }}} \right|^2} = {\delta _{0\ell }},\end{align}

giving

(4.25)\begin{equation}{v_{k = 0}} = \omega {(v{\tau _p})_{k = 0}}/{\kern 1pt} 2{\rm \pi} |qn - m|= \omega qR/|qn - m|.\end{equation}

Consequently, the LH current becomes

(4.26)\begin{equation}{J_{||}} \simeq \frac{{{{\rm \pi}^{1/2}}{n_e}R}}{{2v_e^3}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s } {\frac{{{e^3}|{\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}{|^2}}}{{m_e^2\omega }}} \,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}{\left. {\frac{\partial }{{\partial v}}} \right|_\lambda }{\left. {\left( {\frac{{{v^2}{{\bar{h}}_p}}}{{{f_0}}}} \right)} \right|_{\scriptstyle v = {v_k} \atop \scriptstyle\lambda = 0 = k }}.\end{equation}

Keeping only the j = 1 term in ${\bar{h}_p}$ yields the final result for the driven lower hybrid current in a tokamak to be

(4.27)\begin{equation}{J_{||}} \simeq \frac{{4e{n_e}(1 + 0.62\sqrt \varepsilon )}}{{[(Z + 1)(1 + 2.06\sqrt \varepsilon ) + 4]{\nu _{ee}}}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s } {\frac{{{e^2}|{\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}{|^2}{\omega ^4}}}{{m_e^2|{\kern 1pt} {k_{||}}{\kern 1pt} {|^5}v_e^6}}} \,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}.\end{equation}

Importantly, the key result in the preceding expression is that it retains the proper leading order analytic dependence on the electron trapping parameter $\sqrt \varepsilon $ for the first time. Notice in particular that the aspect ratio dependence of the Z + 1 pitch angle scattering factor differs from that of the electron-electron energy scattering factor 4. Moreover, the overall current drive efficiency is reduced somewhat by the inverse aspect ratio dependences as the increase in the numerator is unable to overcome the decrease from the denominator. Presumably similar behaviour persists for more general cross sections and will make off-axis current profile control slightly more difficult in spherical tokamaks. Perhaps not surprisingly, energy scattering is less affected by toroidal geometry that pitch angle scattering, which is the reason only the leading Cordey (Reference Cordey1976) eigenfunction need be retained. As the freely passing dominate the electron response, (4.27) reduces to the standard $\sqrt \varepsilon {\kern 1pt} \to 0$ result. The driven parallel current (4.27) will be used in the next section to evaluate the current drive efficiency.

5. RF power and current drive efficiency

The rf power required to drive the lower hybrid current, ${P_{cd}}$, is found from

(5.1)\begin{equation}{P_{cd}} = \frac{{{m_e}}}{2}\left\langle {\int {{\textrm{d}^3}v{v^2}} \overline {Q\{ {f_0}\} } } \right\rangle = \frac{{{m_e}}}{2}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,S } {\left\langle {\int {{\textrm{d}^3}v{v^2}} \frac{1}{{{\tau_p}v}}{{\left. {\frac{\partial }{{\partial v}}} \right|}_\mu }\left( {{\tau_p}v\bar{D}\frac{{\partial {f_0}}}{{\partial v}}} \right)} \right\rangle } .\end{equation}

Only the passing electrons contribute as there is no resonance for the trapped since

(5.2)\begin{equation}\oint_t {\textrm{d}\tau } \varLambda \simeq \omega {\tau _t},\end{equation}

and

(5.3)\begin{equation}\Theta \propto {|{\oint_t {\textrm{d}\tau {v_{||}}} \,{\textrm{e}^{ - \textrm{i}\omega (\tau - {\tau_0})}}} |^2} \to 0.\end{equation}

Recalling ${\textrm{d}^3}v = \textrm{d}v\,\textrm{d}\lambda \,\textrm{d}\varphi B{v^3}/2{B_0}{v_{\|}}$, and using $\oint {\textrm{d}\vartheta /\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta } \simeq 2{\rm \pi} qR$

(5.4)\begin{equation}{P_{cd}} = \frac{{{B_0}{m_e}}}{{4{\rm \pi} qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,S } {\int {{\textrm{d}^3}v} } \frac{{{v_{||}}v}}{B}{\left. {\frac{\partial }{{\partial v}}} \right|_\mu }\left( {{\tau_p}v\bar{D}\frac{{\partial {f_0}}}{{\partial v}}} \right) = \frac{{{B_0}m_e^2}}{{{\rm \pi}{T_e}qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,S } {\int {{\textrm{d}^3}v} } \frac{{{v_{||}}}}{B}{\tau _p}{v^2}\bar{D}{f_0}.\end{equation}

Multiplying by 1/2 since only ${v_{\|}} < 0$ contribute, the preceding becomes

(5.5)\begin{equation}{P_{cd}} = \frac{{m_e^2}}{{2{T_e}qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,S } {\int_0^{{B_0}/{B_{\textrm{min}}}} {\textrm{d}\lambda } } \int_0^\infty {\textrm{d}v} {\tau _p}{v^5}\bar{D}{f_0}.\end{equation}

Inserting $\bar{D}$

(5.6) \begin{align}\begin{aligned} {P_{cd}} & = \dfrac{{m_e^2}}{{2{T_e}qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s,\ell } {\dfrac{{{d_\ell }}}{\omega }} \int_0^{{B_0}/{B_{\textrm{max}}}} {\dfrac{{\textrm{d}\lambda }}{{v{\tau _p}}}} \int_0^\infty {\textrm{d}v} {v^5}{f_0}\Theta (v,k)\delta (v - {v_k})\\ &= \dfrac{{m_e^2}}{{2{T_e}qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s,\ell } {\dfrac{{{d_\ell }}}{\omega }} \int_0^{{B_0}/{B_{\textrm{max}}}} {\dfrac{{\textrm{d}\lambda }}{{v{\tau _p}}}} {v^5}{ {{f_0}\Theta (v,k)} |_{v = {v_k}}}\\ & ={-} \dfrac{{2{{\rm \pi}^2}{m_e}}}{{qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s } {\dfrac{{{d_{\ell = 0}}|{\kern 1pt} qn - m{|^2}}}{{{\omega ^3}}}} \int_0^{{B_0}/{B_{\textrm{max}}}} {\dfrac{{\textrm{d}\lambda }}{{{{(v{\tau _p})}^2}}}} \dfrac{{{v^5}\Theta (v,k)}}{{\partial (v{\tau _p}){\kern 1pt} /{\kern 1pt} \partial \lambda }}{\left. {\dfrac{{\partial {f_0}}}{{\partial \lambda }}} \right|_{v = {v_k}}}. \end{aligned}\end{align}

Integrating by parts and noting that only $\ell = 0{\kern 1pt}$ contributes, the rf power density required to drive the lower hybrid current ${J_{\|}}$ is just

(5.7)\begin{equation}{P_{cd}} \simeq \frac{{2{{\rm \pi}^2}{m_e}}}{{qR}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s } {\frac{{{d_{\ell = 0}}|qn - m{|^2}{v^5}}}{{{\omega ^3}{{(v{\tau _p})}^2}\partial (v{\tau _p})/\partial \lambda }}} { {{f_0}} |_{v = {v_k}}} = {{\rm \pi}^{1/2}}{n_e}{m_e}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s } {\frac{{{e^2}|{\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}{|^2}{\omega ^2}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{m_e^2|{k_{||}}{|^3}v_e^3}}} .\end{equation}

The current drive efficiency is defined by forming the ratio ${J_{\|}}/{P_{cd}}$

(5.8) \begin{equation}\frac{{{J_{||}}}}{{{P_{cd}}}} = \frac{\displaystyle{4e(1 + 0.62\sqrt {2\varepsilon } )\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s } {\frac{{|{\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}{|^2}{\omega ^4}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{|{k_{||}}{|^5}v_e^6}}} }}{\displaystyle{\sqrt {\rm \pi} {m_e}[(Z + 1)(1 + 2.06\sqrt \varepsilon ) + 4]{\nu _{ee}}\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s } {\frac{{|{\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}{|^2}{\omega ^2}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{|{k_{||}}{|^3}v_e^3}}} }},\end{equation}

or in its normalized form

(5.9) \begin{equation}\frac{{{J_{||}}/e{n_e}{v_e}}}{{{P_{cd}}/{n_e}{m_e}v_e^2{\nu _{ee}}}} \equiv \eta = \frac{\displaystyle{4(1 + 0.62\sqrt {2\varepsilon } )\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s } {\frac{{|{\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}{|^2}{\omega ^4}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{|{k_{||}}{|^5}v_e^5}}} }}{\displaystyle{\sqrt {\rm \pi} [(Z + 1)(1 + 2.06\sqrt \varepsilon ) + 4]\sum\limits_{\scriptstyle\omega ,m \atop \scriptstyle n,s } {\frac{{|{\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}{|^2}{\omega ^2}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{|{k_{||}}{|^3}v_e^3}}} }}.\end{equation}

Various normalizations of ${J_{\|}}$ and ${P_{cd}}$ (often because of a $\sqrt 2$ difference in the definition of the electron thermal speed), and differing definitions of collision frequency appear in the LHCD literature, as well as incompletely defined notation.

For a single frequency and single toroidal mode number (5.9) leads to the dimensionless current drive efficiency with electron trapping retained of

(5.10) \begin{align} \eta &= \frac{\displaystyle{4(1 + 0.62\sqrt {2\varepsilon } )\sum\limits_{m,s} {\frac{{|{\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}{|^2}{\omega ^5}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{|{k_{||}}{|^5}v_e^5}}} }}{\displaystyle{\sqrt {\rm \pi} [(Z + 1)(1 + 2.06\sqrt \varepsilon ) + 4]\sum\limits_{m,s} {\frac{{|{\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}{|^2}{\omega ^3}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{|{k_{||}}{|^3}v_e^3}}} }}\nonumber\\ &\to \frac{{4(1 + 0.62\sqrt {2\varepsilon } ){\omega ^2}}}{{\sqrt {\rm \pi}[(Z + 1)(1 + 2.06\sqrt \varepsilon ) + 4]|{\kern 1pt} {k_{||}}{|^2}v_e^2}},\end{align}

where the sums over poloidal mode number and radial mode structure (either as a Fourier or eikonal representation) are retained except in the final form.

6. Discussion

The key results derived here are the analytic expressions for the LHCD (4.26) and (4.27) and the LHCD efficiency (5.9) and (5.10) in tokamak geometry that properly retain the trapped electron modifications to leading order. The procedure used here can be extended to include lower speed electrons by using the full Rosenbluth form of the electron--electron collision operator with a momentum conserving modification. Keeping additional Cordey (Reference Cordey1976) eigenfunctions will only make minor, unimportant changes to the results here. However, generalizing the Cordey eigenfunctions to finite aspect ratio in a more realistic geometry will lead to improved results as the separatrix is approached. Such a treatment might be valuable as lower hybrid is viewed as an effective means to drive and control the off axis current profile (Bonoli Reference Bonoli2014). Of course, near the separatrix and beyond the long mean free path limit will become inappropriate and the Cordey eigenfunctions are no longer useful.

The QL derivation of Catto & Tolman (Reference Catto and Tolman2021) remains valid as long as the collisional boundary layer about the resonant electron trajectories defined in (4.11) and (4.12) remains narrow enough in velocity space to satisfy $1 \gg {({\nu _{ee}}/{k_{\|}}{v_e})^{1/3}} \propto n_e^{1/3}{\kern 1pt} /T_e^{7/6}$. The QL treatment they derive and the results here are expected to fail once the applied rf amplitude substantially distorts the electron distribution function from Maxwellian (Catto Reference Catto2020; Catto & Tolman Reference Catto and Tolman2021). This level of distortion is estimated by Catto & Tolman (Reference Catto and Tolman2021) to occur once the applied rf becomes sufficiently strong that the nonlinear term in the linearized Fokker--Planck equation for the electron kinetic response no longer satisfies $e|{\boldsymbol{e}_m}\boldsymbol{\cdot }\boldsymbol{n}|/{m_e} \ll \nu _{ee}^{2/3}k_{\|}^{1/3}v_e^{4/3}{\kern 1pt} \propto n_e^{2/3}/T_e^{1/3}$ and integrating over unperturbed electron trajectories to obtain the QL operator becomes inappropriate. This estimate may provide a hint as to why LHCD becomes less efficient at higher densities (Bonoli Reference Bonoli2014).

Acknowledgements

Work supported by the U.S. Department of Energy grant DE-FG02-91ER-54109 at MIT. The author is grateful for suggestions from Dr P. Bonoli.

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

Declaration of interests

The author reports no conflict of interest.

References

Antonsen, T.M. & Chu, M.S. 1982 Radio frequency current generation by waves in toroidal geometry. Phys. Fluids 25, 12951296.CrossRefGoogle Scholar
Bonoli, P.T. 2014 Review of recent experimental and modeling progress in the lower hybrid range of frequencies at ITER relevant parameters. Phys. Plasmas 21, 061508.CrossRefGoogle Scholar
Catto, P.J. 2020 Collisional effects on resonant particles in quasilinear theory. J. Plasma Phys. 86, 815860302.CrossRefGoogle Scholar
Catto, P.J. & Tolman, E.A. 2021 Reimagining full wave quasilinear theory in a tokamak. J. Plasma Phys. 87, 905870215.CrossRefGoogle Scholar
Chiu, S.C., Chan, V.S., Harvey, R.W. & Porkolab, M. 1989 Theory of fast wave current drive for tokamak plasmas. Nucl. Fusion 29, 21752186.CrossRefGoogle Scholar
Cohen, R.H. 1987 Effect of trapped electrons on current drive. Phys. Fluids 30, 24422449.CrossRefGoogle Scholar
Cordey, J.G. 1976 Effects of particle trapping on the slowing-down of fast ions in a toroidal plasma. Nucl. Fusion 16, 499507.CrossRefGoogle Scholar
Cordey, J.G., Edlington, T. & Start, D.F.H. 1982 A theory of currents induced by radio-frequency waves in toroidal plasmas. Plasma Phys. 24, 7389.CrossRefGoogle Scholar
Ehst, D.A. & Karney, C.F.F. 1991 Approximate formula for radio-frequency current drive efficiency with magnetic trapping. Nucl. Fusion 31, 19331938.CrossRefGoogle Scholar
Fisch, N.J. 1978 Confining a tokamak plasma with rf-driven currents. Phys. Rev. Lett. 41, 873876.CrossRefGoogle Scholar
Fisch, N.J. & Boozer, A.H. 1980 Creating an axisymmetric plasma resistivity with waves. Phys. Rev. Lett. 45, 720722.CrossRefGoogle Scholar
Fisch, N.J. & Karney, C.F.F. 1981 Current generation with low-frequency waves. Phys. Fluids 24, 2739.CrossRefGoogle Scholar
Giruzzi, G. 1987 Impact of electron trapping on rf current drive in tokamaks. Nucl. Fusion 27, 19341939.CrossRefGoogle Scholar
Helander, P., Geiger, J. & Maaßberg, H. 2011 On the bootstrap current in stellarators and tokamaks. Phys. Plasmas 18, 092505.CrossRefGoogle Scholar
Helander, P. & Sigmar, D.J. 2002 Collisional Transport in Magnetized Plasmas, pp. 191195. Cambridge University Press.Google Scholar
Hinton, F.L. & Hazeltine, R.D. 1976 Theory of plasma transport in toroidal confinement systems. Rev. Mod. Phys. 48, 239308.CrossRefGoogle Scholar
Hsu, C.T., Catto, P.J. & Sigmar, D.J. 1990 Neoclassical transport of isotropic fast ions. Phys. Plasmas B 2, 280290.Google Scholar
Karney, C.F.F. & Fisch, N.J. 1979 Numerical studies of current generation by radio-frequency traveling waves. Phys. Fluids 22, 18171824.CrossRefGoogle Scholar
Karney, C.F.F. & Fisch, N.J. 1985 Efficiency of current drive by fast waves. Phys. Fluids 28, 116126.CrossRefGoogle Scholar
Kennel, C.F. & Engelmann, F. 1966 Velocity space diffusion from weak plasma turbulence in a magnetic field. Phys. Fluids 9, 23772388.CrossRefGoogle Scholar
Parker, J.B. & Catto, P.J. 2012 Variational calculation of neoclassical ion heat flux and poloidal flow in the banana regime for axisymmetric magnetic geometry. Plasma Phys. Control. Fusion 54, 085011.CrossRefGoogle Scholar
Pusztai, I. & Catto, P.J. 2010 Neoclassical plateau regime transport in a tokamak pedestal.. Plasma Phys. Control. Fusion 52, 075016.CrossRefGoogle Scholar
Spitzer, L. & Härm, R. 1953 Transport phenomena in a completely ionized gas. Phys. Rev. 89, 977981.CrossRefGoogle Scholar
Taguchi, M. 1983 The effect of trapped electron on the wave-induced current. J. Phys. Soc. Japan 52, 20352040.CrossRefGoogle Scholar
Xiao, Y., Catto, P.J. & Molvig, K. 2007 Collisional damping for ion temperature graient mode driven zonal flow. Phys. Plasmas 14, 032302.CrossRefGoogle Scholar
Yoshioka, K. & Antonsen, T.M. Jr. 1986 Neoclassical effects on rf current drive in tokamaks. Nucl. Fusion 26, 839847.CrossRefGoogle Scholar