Hostname: page-component-7bb8b95d7b-495rp Total loading time: 0 Render date: 2024-09-30T03:13:29.713Z Has data issue: false hasContentIssue false

Helicon and lower hybrid current drive comparisons in tokamak geometry

Published online by Cambridge University Press:  17 July 2023

Peter J. Catto*
Affiliation:
Plasma Science and Fusion Center, Massachusetts Institute of Technology, Cambridge, MA, USA
Muni Zhou
Affiliation:
Plasma Science and Fusion Center, Massachusetts Institute of Technology, Cambridge, MA, USA Department of Astrophysical Sciences, Princeton University, Princeton, NJ, USA School of Natural Science, Institute for Advanced Study, Princeton, NJ, USA
*
Email address for correspondence: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

The parallel current driven by applied helicon waves is evaluated in tokamak geometry along with the radio frequency (rf) power absorbed by the passing electrons. The results are compared with the corresponding expressions for lower hybrid current drive. The efficiency of both current drive schemes is found to be the same for any single wave frequency, single mode number limit. The evaluation of the parallel currents is performed using an adjoint technique. Tokamak geometry is retained by using an eigenfunction expansion appropriate for a transit averaged long mean free path treatment of electrons making correlated poloidal passes through the applied rf fields.

Type
Research Article
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press

1. Introduction

Investigating helicon (or whistler) waves as a means to drive a parallel current in a tokamak is a less studied area of current drive (Pinsker Reference Pinsker2015) than the more common mechanism of lower hybrid waves (Fisch Reference Fisch1978; Karney & Fisch Reference Karney and Fisch1979, Reference Karney and Fisch1985; Fisch & Boozer Reference Fisch and Boozer1980; Fisch & Karney Reference Fisch and Karney1981; Cordey, Edlington & Start Reference Cordey, Edlington and Start1982; Taguchi Reference Taguchi1983; Cohen Reference Cohen1987; Giruzzi Reference Giruzzi1987; Chiu et al. Reference Chiu, Chan, Harvey and Porkolab1989; Ehst & Karney Reference Ehst and Karney1991) which is extensively reviewed by Bonoli (Reference Bonoli2014). Nonetheless, helicon current drive (HCD) remains of considerable interest (Vdovin Reference Vdovin2013; Prater et al. Reference Prater, Moeller, Pinsker, Porkolab, Meneghini and Vdovin2014; Pinsker Reference Pinsker2015; Wang et al. Reference Wang, Wi, Kim, Kim, Jeong and Kwak2017; Lau et al. Reference Lau, Jaeger, Bertelli, Berry, Green, Murakami, Park, Pinsker and Prater2018, Reference Lau, Berry, Jaeger and Bertelli2019; Pinsker et al. Reference Pinsker, Prater, Moeller, Degrassie, Petty, Porkolab, Anderson, Garofalo, Lau, Nagy, Pace, Torreblanca, Watkins and Zeng2018) because of its many similarities to lower hybrid current drive (LHCD) and its ability to efficiently access higher density regions (Li et al. Reference Li, Ding, Dong and Liu2020a,Reference Li, Liu, Xiang and Lib; Lau et al. Reference Lau, Bookman, Dimits, Dudson, Martin, Pinsker, Thomas and van Compernolle2021; Li, Li & Liu Reference Li, Li and Liu2021; Yin et al. Reference Yin, Zheng, Gong, Yang, Yin, Song, Huang, Chen and Zhong2022). Both HCD and LHCD rely on a Landau resonance and the preferential heating of electrons, although HCD uses a perpendicular component of the applied electric field rather than the parallel component used for LHCD. An early analytic attempt at evaluating the parallel current driven by helicon waves and the associated efficiency appears in de Assis & Busnardo-Neto (Reference de Assis and Busnardo-Neto1988). They give rough estimates for a model collision operator in a constant magnetic field with some inadequately defined notation. Little else in the way of analytic estimates is available in the literature, even though a great deal of effort has been expended on simulations, planning experiments and designing antennas for DIII-D (Prater et al. Reference Prater, Moeller, Pinsker, Porkolab, Meneghini and Vdovin2014; Lau et al. Reference Lau, Jaeger, Bertelli, Berry, Green, Murakami, Park, Pinsker and Prater2018, Reference Lau, Berry, Jaeger and Bertelli2019, Reference Lau, Bookman, Dimits, Dudson, Martin, Pinsker, Thomas and van Compernolle2021; Pinsker et al. Reference Pinsker, Prater, Moeller, Degrassie, Petty, Porkolab, Anderson, Garofalo, Lau, Nagy, Pace, Torreblanca, Watkins and Zeng2018), KSTAR (Wang et al. Reference Wang, Wi, Kim, Kim, Jeong and Kwak2017); EAST (Li et al. Reference Li, Liu, Xiang and Li2020b); and HL-2M (Li et al. Reference Li, Ding, Dong and Liu2020a); as well as ITER and DEMO (Vdovin Reference Vdovin2013; Lau et al. Reference Lau, Jaeger, Bertelli, Berry, Green, Murakami, Park, Pinsker and Prater2018) and CFETR (Li et al. Reference Li, Li and Liu2021). The purpose of the investigation here is to derive analytic expressions for both the parallel current driven by helicon waves and its associated HCD efficiency in a tokamak, and to compare them with the recently derived expressions for LHCD (Catto Reference Catto2021) that are generalized herein. It turns out that the adjoint procedure (Antonsen & Chu Reference Antonsen and Chu1982) used for these evaluations have many similarities so the comparisons can be made in a meaningful way. Moreover, these results are derived in tokamak geometry by transit averaging and then using the Cordey (Reference Cordey1976) eigenfunctions and associated results (Hsu, Catto & Sigmar Reference Hsu, Catto and Sigmar1990; Xiao, Catto & Molvig Reference Xiao, Catto and Molvig2007; Parker & Catto Reference Parker and Catto2012). In addition, since the electrons are only weakly collisional, successive poloidal passes through the applied radio frequency (rf) fields are correlated and the quasilinear (QL) description employed (Catto & Tolman Reference Catto and Tolman2021a) accounts for this feature.

The next section introduces the transit averaged QL operator employed for the adjoint evaluations of the parallel current, as well as notation. The adjoint technique is briefly summarized in § 3, where the unlike collision operator is also presented. Section 4 summarizes the solution for the adjoint equation in a tokamak and also gives the like particle collision operator employed. The parallel current driven by helicon waves is derived in § 5, which also presents an improved evaluation of the lower hybrid results. The rf power absorbed by the passing electrons and the current drive efficiencies associated with HCD and LHCD are presented in § 6. Section 7 gives results when HCD and LHCD are both operative. The Appendix presents some cold plasma material that suggests that HCD and LHCD are able to drive comparable currents.

2. Background

In a tokamak the transit averaged QL operator for electrons, when the applied wave field is at a frequency $\omega$ much smaller than the electron cyclotron frequency ${\varOmega _e} = eB/{m_e}c$ and the unperturbed electron distribution function is nearly the Maxwellian ${f_0}$, is

(2.1)\begin{equation}\bar{Q}\{ {f_0}\} = \mathop \sum \limits_{\boldsymbol{k}} \frac{1}{{{\tau _f}}}\frac{\partial }{{\partial E}}\left( {{\tau_f}{v^2}D\frac{{\partial {f_0}}}{{\partial E}}} \right),\end{equation}

with ${\tau _f} = \oint_f {\textrm{d}\tau }$ the time for a full ( f) poloidal passing $(\sigma = {v_{||}}/|{v_{||}}| = \pm 1)$ or trapped $(\sigma = 0)$ poloidal circuit, the sum over the applied rf wave vectors $\boldsymbol{k}$, and D the velocity space diffusivity (Catto & Tolman Reference Catto and Tolman2021a). Here e is the charge on a proton, c is the speed of light, ${m_e}$ is the electron mass and $E = v^{2}/2 - e\varPhi/m_{e}$ is the total energy, with $\varPhi$ the electrostatic potential, $v = |\boldsymbol{v}|$ the electron speed and $v_{||} = \boldsymbol{n} \boldsymbol{\cdot} \boldsymbol{v}$ the parallel electron velocity along the tokamak magnetic field $\boldsymbol{B} = B\boldsymbol{n} = I\boldsymbol{\nabla }\zeta + \boldsymbol{\nabla }\zeta \times \boldsymbol{\nabla }\psi \;$. The unit vector along the magnetic field is $\boldsymbol{n}$, $\zeta \;$ is the toroidal angle variable, $\psi$ is the poloidal flux function and the flux function $I(\psi )$ is $I = R{B_t}$, where ${B_t}$ is the toroidal magnetic field, R is the major radius and ${B_p}$ is the poloidal magnetic field in $|\boldsymbol{\nabla }\psi |= R{B_p}$. The poloidal angle $\vartheta$ satisfies $\boldsymbol{\nabla }\zeta \boldsymbol{\cdot }\boldsymbol{\nabla }\psi = 0 = \boldsymbol{\nabla }\zeta \boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta$ and is chosen such that $\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta = |I|/q{R^2} = {q^{ - 1}}|\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\zeta |$, making the safety factor, $q = q(\psi )$, a flux function. Taking the toroidal current to be in the $\boldsymbol{\nabla }\zeta$ direction makes ${B_p} > 0$, $\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta > 0$ and $\vartheta$ increases in the ${\boldsymbol{B}_p} = \boldsymbol{\nabla }\zeta \times \boldsymbol{\nabla }\psi$ direction. Then the incremental time along a trajectory is $\textrm{d}{\tau} = \textrm{d}\vartheta/v_{||}\boldsymbol{n} \boldsymbol{\cdot} \boldsymbol{\nabla}\vartheta > 0$, with $\textrm{d}\vartheta$ and $v_{||}$ reversing signs together when a trapped electron reflects, such that $\textrm{d}\vartheta < 0$ for a passing electron with ${v_{||}} < 0$.

The QL diffusivity D in tokamak geometry for successive correlated interactions through the applied rf (Catto & Tolman 2021) is

(2.2)\begin{equation}D = \frac{{\mathrm{\pi }{e^2}}}{{2m_e^2{v^2}{\tau _f}}}\sum\limits_\ell {\delta ({\oint_f {\textrm{d}\tau \chi - 2\mathrm{\pi }\ell } } )} {\left|{\oint_f {\textrm{d}\tau {\boldsymbol{e}_k}\boldsymbol{\cdot }\left[ {\boldsymbol{n}{v_{||}}{J_0}(\eta ) - i\boldsymbol{n} \times \boldsymbol{k}\frac{{{v_ \bot }}}{{{k_ \bot }}}{J_1}(\eta )} \right]\,{\textrm{e}^{ - \textrm{i}\int_{{\tau_0}}^\tau {\textrm{d}\tau^{\prime}\chi (\tau^{\prime})} }}} } \right|^2},\end{equation}

with the transit averaged resonance condition defined by

(2.3)\begin{equation}\oint_f {\textrm{d}\tau \chi } = \omega {\tau _f} - 2\mathrm{\pi }\sigma (qn - m),\end{equation}

and the exponential phase factor given by

(2.4)\begin{equation}\int_{{\tau_0}}^\tau {\textrm{d}\tau^{\prime}\chi (\tau^{\prime})}=\omega(\tau-\tau_{0})-\sigma(qn-m)\vartheta(\tau),\end{equation}

where ${\tau _0}$ is taken to be the trajectory time at the equatorial plane crossing where B is a minimum. In the preceding and what follows ${\boldsymbol{e}_k}$ is the Fourier amplitude of the applied electric field of wave vector $\boldsymbol{k} = {\boldsymbol{k}_ \bot } + {k_{||}}\boldsymbol{n}$ having ${\boldsymbol{k}_ \bot } = {k_ \bot }(\boldsymbol{\psi }\cos \varsigma + \boldsymbol{p}\sin \varsigma )$, ${k_{||}} = (qn - m)/qR$ and $\eta = {k_ \bot }{v_ \bot }/{\varOmega _e}$, with $\boldsymbol{\psi } = \boldsymbol{\nabla }\psi /|\boldsymbol{\nabla }\psi |= \boldsymbol{\nabla }\psi /R{B_p}$, and $\boldsymbol{p} = \boldsymbol{\nabla }\zeta \times \boldsymbol{\nabla }\psi /|\boldsymbol{\nabla }\zeta \times \boldsymbol{\nabla }\psi |= {\boldsymbol{B}_p}/{B_p}$. The orthonormal unit vectors satisfy $\boldsymbol{\psi } \times \boldsymbol{p} = \boldsymbol{n}$, n and m are the toroidal and poloidal mode numbers, respectively, and large aspect ratio is assumed to write $\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta \approx 1/qR$. The integer $\ell$ denotes the resonant path in velocity space as electrons can experience other (usually less important) resonances besides $\ell = 0$ in toroidal geometry as their velocity changes along a field line as they cannot remain in resonance indefinitely in an inhomogeneous magnetic field. Indeed, they can have other resonant interactions by crossing the various resonant paths or curves in velocity space defined by $\oint_f {\textrm{d}\tau \chi } = 2\mathrm{\pi }\ell$.

For LHCD the applied rf fields are tailored to make the first or ${v_{||}}{J_0}$ term dominate. For HCD the rf fields must be applied in a manner that makes the second or ${v_ \bot }{J_1}$ term dominate (Prater et al. Reference Prater, Moeller, Pinsker, Porkolab, Meneghini and Vdovin2014; Pinsker Reference Pinsker2015; Pinsker et al. Reference Pinsker, Prater, Moeller, Degrassie, Petty, Porkolab, Anderson, Garofalo, Lau, Nagy, Pace, Torreblanca, Watkins and Zeng2018; Yin et al. Reference Yin, Zheng, Gong, Yang, Yin, Song, Huang, Chen and Zhong2022), so ${J_1}$ will not be expanded by assuming its argument is small. Of course, both HCD and LHCD rely on there being a Landau resonance. Here the goal is to evaluate HCD by an adjoint technique (Antonsen & Chu Reference Antonsen and Chu1982) very similar to the one used recently to evaluate LHCD in full toroidal geometry to find the aspect ratio modifications to the driven current and efficiency (Catto Reference Catto2021). Unlike for LHCD, HCD does not seem to have an accepted analytic expression for the driven current or efficiency even without toroidal effects retained.

3. Adjoint technique summary

For an adjoint evaluation of HCD the preceding QL operator is all that is required along with the perturbed electron kinetic equation

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

and its adjoint equation for the adjoint function h associated with ${f_1}$

(3.2)\begin{equation}{v_{||}}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }h + C\{ h\} ={-} \left( {\frac{B}{I} - \frac{{{m_e}{\nu_u}}}{{{T_e}{x^3}}}{V_{||}}} \right){v_{||}}{f_0},\end{equation}

where the small term containing the parallel mean ion velocity ${V_{||}}$ is needed to account for the non-self-adjoint term in the electron-ion collision operator and ${f_0}$ is the Maxwellian

(3.3)\begin{equation}{f_0} = {f_0}(\psi ,E) = {n_e}(\psi ){\left( {\frac{{{m_e}}}{{2\mathrm{\pi }{T_e}}}} \right)^{3/2}}\,{\textrm{e}^{ - [{m_e}E + e\varPhi (\psi )]/{T_e}(\psi )}}.\end{equation}

The potential may be assumed to be a flux function to lowest order. In the electron kinetic equation ${f_1}$ is the perturbed electron distribution function with $f = {f_0} + {f_1}$, $Q\{ {f_0}\}$ is the QL operator prior to transit averaging, and $C\{ {f_1}\} = {C_{\textrm{ee}}}\{ {f_1}\} + {C_{\textrm{ei}}}\{ {f_1}\}$ is the sum of the self-adjoint linearized electron-electron collision operator plus the electron-ion collision operator

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

where $L\{ h\}$ is the self-adjoint Lorentz or pitch angle scattering operator

(3.5)\begin{equation}L\{ h\} = \frac{{2{B_0}}}{B}\xi \frac{\partial }{{\partial \lambda }}\left( {\lambda \xi \frac{{\partial h}}{{\partial \lambda }}} \right),\end{equation}

with $x = v/{v_e}$ and ${v_e} = {(2{T_e}/{m_e})^{1/2}}$ the electron thermal speed. The pitch angle is defined as $\lambda = 2\mu {B_0}/B{v^2} = {B_0}v_ \bot ^2/B{v^2}$ and $\xi = {v_{||}}/v$, with ${B_0}$ a normalizing flux function to be defined shortly. The unlike collision frequency ${\nu _u}$ is defined as

(3.6)\begin{equation}{\nu _u} = 3\sqrt {\pi } {\nu _{\textrm{ei}}}/4,\end{equation}

where ${\nu _{\textrm{ei}}} = 4\sqrt {2{\pi }} {Z^2}{e^4}{n_i}\ell n{\varLambda _C}/3m_e^{1/2}T_e^{3/2} = Z{\nu _{\textrm{ee}}}$ for a quasineutral plasma with the ion, ${n_i}$, and electron, ${n_e}$, densities satisfying $Z{n_i} = {n_e}$, Z the ion charge number, ${\nu _{\textrm{ee}}}$ the electron-electron collision frequency and $\ell n{\varLambda _C}$ the Coulomb logarithm.

Defining the flux surface average of any quantity A by

(3.7)\begin{equation}\langle A\rangle = {{\left({\oint {\textrm{d}\vartheta A/\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta } } \right)}\Bigg/ {\left({\oint {\textrm{d}\vartheta /\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta } } \right)}},\end{equation}

using $\langle \boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }({B^{ - 1}}\int {{\textrm{d}^3}v{v_{||}}h{f_1}/{f_0}} \rangle = 0$, and the self-adjointness of both ${C_{\textrm{ee}}}$ (that requires ${f_0}$ to be Maxwellian) and L to combine the equations leads to the convenient adjoint relation

(3.8)\begin{equation}\left\langle {B\int {{\textrm{d}^3}v{v_{||}}{f_1}} } \right\rangle = {{I\left( {\int {{\textrm{d}^3}v\frac{{{v_{||}}\bar{h}}}{{B{f_0}}}{\tau_f}\bar{Q}\{ {f_0}\} } } \right)} \Bigg/ {\left({\oint {\textrm{d}\vartheta /\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta } } \right)}},\end{equation}

where the ${V_{||}}\sim {v_i}{\rho _{\textrm{pi}}}/a$ term, with ${v_i}$ the ion thermal speed, ${\rho _{\textrm{pi}}}$ the poloidal ion gyroradius and a the minor radius, is negligible for the rf amplitudes of interest. Catto (Reference Catto2021) can be consulted for more details on this approximation and the treatment to follow. To write the final form of $\langle B\int {{\textrm{d}^3}v{v_{||}}{f_1}} \rangle$ the lowest-order result $\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }\bar{h} = 0$ is employed along with the transit average definition $\bar{A} = \oint_{} {\textrm{d}\tau A/{\tau _f}}$. It is convenient to let $B_0^2 = \langle {B^2}\rangle$.

Recall the Ohmic current is in the positive toroidal direction. Therefore, the helicon waves must drive the current in the same direction, requiring $\langle B\int {{\textrm{d}^3}v{v_{||}}{f_1}} \rangle < 0$.

4. Solution of the adjoint equation

The advantage of the adjoint method is that only the simpler adjoint equation

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

need be solved instead of the more complicated electron kinetic equation. To solve the preceding equation, it is adequate to approximate the electron-electron collision operator by its standard high speed $({v^2} \gg v_e^2)$ expansion. Then its self-adjoint form becomes

(4.2)\begin{equation}{C_{\textrm{ee}}}\{ h\} = \frac{{2{\nu _\ell }{B_0}\xi }}{B}\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{equation}

where the like collision frequency is defined as

(4.3)\begin{equation}{\nu _\ell } = 3\sqrt {\pi } {\nu _{\textrm{ee}}}/4,\end{equation}

with ${\nu _{\textrm{ee}}} = 4\sqrt {2\mathrm{\pi }} {e^4}{n_e}\ell n{\varLambda _C}/3m_e^{1/2}T_e^{3/2}$. This like particle operator is the usual non-momentum conserving, high speed expansion of the Rosenbluth potentials for collisions with a Maxwellian used by Karney & Fisch (Reference Karney and Fisch1979, Reference Karney and Fisch1985). Recent estimates (Catto Reference Catto2021; Catto & Tolman Reference Catto and Tolman2021a,Reference Catto and Tolmanb) indicate ${f_0}$ must be nearly Maxwellian for QL theory to remain valid.

The adjoint equation is solved using the Cordey eigenfunctions (Cordey Reference Cordey1976; Hsu et al. Reference Hsu, Catto and Sigmar1990; Xiao et al. Reference Xiao, Catto and Molvig2007; Parker & Catto Reference Parker and Catto2012) by writing $h = \bar{h} + \tilde{h}$ with $\partial \bar{h}/\partial \vartheta = 0$ to obtain, upon annihilating ${v_{||}}\boldsymbol{n}\boldsymbol{\cdot }\boldsymbol{\nabla }\tilde{h}$ the term, the transit average equation

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

Integration over a full trapped (t) bounce gives $\overline {B{v_{||}}} = 0$ implying that ${\bar{h}_t} = 0$. For the passing (p) electrons flux surface averages are used to rewrite the adjoint equation as

(4.5)\begin{equation}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}\langle B/{v_{||}}\rangle }}{{{m_e}{B_0}{v^2}{f_0}}}\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{equation}

where $\overline {B{v_{||}}} \oint_p {\textrm{d}\tau } = \langle {B^2}\rangle \oint {\textrm{d}\vartheta /\boldsymbol{B}\boldsymbol{\cdot }\boldsymbol{\nabla }\vartheta }$.

For the purposes here, the recent solution of (4.5) by Catto (Reference Catto2021) is adequate and convenient. Ignoring order $\varepsilon = r/R \ll 1$ terms it is given by

(4.6)\begin{align} \begin{aligned} \dfrac{{{{\bar{h}}_p}}}{{{f_0}}} & \approx \dfrac{{v{x^3}}}{{R{\nu _\ell }}}\left\{ {\dfrac{{(1 + 0.62\sqrt \varepsilon ){\varLambda_1}(\lambda ) - 1.02[(Z + 5)/(7Z + 11)]\sqrt \varepsilon {\varLambda_2}(\lambda )}}{{[(Z + 1)(1 + 1.46\sqrt {2\varepsilon } ) + 4]}}} \right\}\\ & \equiv \dfrac{{v{x^3}}}{{R{\nu _\ell }}}\dfrac{{{\varLambda _{1 + 2}}(\sqrt \varepsilon ,Z,\lambda )}}{{\left[ {(Z + 1)(1 + 1.46\sqrt {2\varepsilon } ) + 4} \right]}}, \end{aligned}\end{align}

where the Cordey (Reference Cordey1976) eigenfunctions ${\varLambda _j}$ with eigenvalues ${\kappa _j}$ satisfy the Sturm--Liouville differential equation obtained by transit averaging,

(4.7)\begin{equation}\frac{\partial }{{\partial \lambda }}\left( {\lambda \langle \xi \rangle \frac{{\partial {\varLambda_j}}}{{\partial \lambda }}} \right) = {\kappa _j}\frac{{\partial \langle \xi \rangle }}{{\partial \lambda }}{\varLambda _{j\; }}.\end{equation}

In the preceding, $\partial \langle \xi \rangle /\partial \lambda ={-} \langle B/2{B_0}\xi \rangle$, ${\varLambda _j}(\lambda = 0) = 1$,

(4.8)\begin{gather}{\varLambda _{1 + 2}}(\sqrt \varepsilon ,Z,\lambda = 0) = 1 + [0.62 - 1.02(Z + 5)/(7Z + 11)]\sqrt \varepsilon ,\end{gather}
(4.9)\begin{gather}\langle \xi \rangle = 2\sqrt {2\varepsilon } E(k)/\mathrm{\pi }\sqrt {(1 - \varepsilon ){k^2} + 2\varepsilon } ,\end{gather}

where E(k) is the elliptic integral of the second kind, ${k^2} = 2\varepsilon \lambda /[1 - (1 - \varepsilon )\lambda ]$ and ${\varLambda _j} = 0$ at k = 1. The response ${\bar{h}_p}$ increases with speed because of the v dependence of $C\{ h\}$. For the passing

(4.10)\begin{equation}{\tau _p} = \oint_p {\textrm{d}\tau } \approx 4qR\sqrt {(1 - \varepsilon ){k^2} + 2\varepsilon } K(k)/v\sqrt {2\varepsilon } ,\end{equation}

with K(k) the elliptic integral of the first kind. The ${\varLambda _2}$ term has a somewhat small numerical coefficient. It was ignored in Catto (Reference Catto2021), but is retained here as it modestly alters the $\sqrt \varepsilon$ correction of interest (order $\varepsilon$ corrections are ignored). More eigenfunction details are available in Catto (Reference Catto2021), Hsu et al. (Reference Hsu, Catto and Sigmar1990), Xiao et al. (Reference Xiao, Catto and Molvig2007) and Parker & Catto (Reference Parker and Catto2012).

5. Helicon driven current in a tokamak compared with LHCD

Only the passing electrons contribute to helicon drive current, giving

(5.1)\begin{align}\left\langle {B\int {{\textrm{d}^3}v{v_{||}}{f_1}} } \right\rangle = \frac{{\langle {B^2}\rangle }}{{2\mathrm{\pi }q}}\int {{\textrm{d}^3}v\frac{{{v_{||}}{{\bar{h}}_p}}}{{B{f_0}}}{\tau _p}\bar{Q}\{ {f_0}\} } = \frac{{\langle {B^2}\rangle }}{{2\mathrm{\pi }q}}\sum\limits_{\boldsymbol{k}} {\int {{\textrm{d}^3}v\frac{{{v_{||}}{{\bar{h}}_p}}}{{Bv{f_0}}}{{\left. {\frac{\partial }{{\partial v}}} \right|}_\mu }\left( {{\tau_p}vD{{\left. {\frac{{\partial {f_0}}}{{\partial v}}} \right|}_\mu }} \right)} } ,\end{align}

at large aspect ratio, with the QL diffusivity for helicon waves simplifying to

(5.2)\begin{equation}D = \frac{{\mathrm{\pi }{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}{\tau _p}v_ \bot ^2J_1^2(\eta )}}{{8m_e^2k_ \bot ^2{v^2}}}\sum\limits_\ell {\delta [\omega {\tau _p} - 2\mathrm{\pi }\sigma (qn - m) - 2\mathrm{\pi }\ell ]\varTheta (v,\lambda ,n,m)} .\end{equation}

Here,

(5.3)\begin{align}\varTheta = \frac{1}{{\tau _p^2}}{|{\oint_p {\textrm{d}\tau \,{\textrm{e}^{ - \textrm{i}[\omega (\tau - {\tau_0}) - \sigma (qn - m)\vartheta (\tau )]}}} } |^2} = \frac{{{q^2}{R^2}}}{{\tau _p^2{v^2}}}{\left|{\int_{ - \mathrm{\pi }}^\mathrm{\pi } {\frac{{\textrm{d}\vartheta }}{\xi }\,{\textrm{e}^{ - \textrm{i}\left[ {\omega qR\int_0^\vartheta {\textrm{d}\vartheta /{v_{||}} - \sigma (qn - m)\vartheta } } \right]}}} } \right|^2} \le 1,\end{align}

and ${\textrm{d}^3}v \to 2\mathrm{\pi }B\,\textrm{d}E\,\textrm{d}\mu /{v_{||}} \to 2\mathrm{\pi }B{v^3}\,\textrm{d}v\,\textrm{d}\lambda /{B_0}{v_{||}}$. Integrating by parts in $E,\mu$ variables, yields

(5.4)\begin{equation}\left\langle {B\int {{\textrm{d}^3}v{v_{||}}{f_1}} } \right\rangle = \frac{{{m_e}\langle {B^2}\rangle }}{{2\mathrm{\pi }q{T_e}}}\sum\limits_{\boldsymbol{k}} {\int {{\textrm{d}^3}v\frac{{{v_{||}}}}{B}{\tau _p}vD{f_0}{{\left. {\frac{\partial }{{\partial v}}} \right|}_\mu }\left( {\frac{{{{\bar{h}}_p}}}{{{f_0}}}} \right)} } .\end{equation}

Catto (Reference Catto2021) has an extra ${v^2}$ multiplying ${\bar{h}_p}$ in his (4.7), implying the factors of $v$ in his (4.27) for the lower hybrid driven parallel current need to be corrected, as will be found shortly.

To sustain or enhance the poloidal magnetic field the parallel current driven by the helicon waves must be positive, requiring $\langle B\smallint {\textrm{d}^3}v{v_{||}}{f_1}\rangle < 0$. Therefore, passing electrons with ${v_{||}} < 0$ must drive the current implying $\sigma ={-} 1$ and $\ell + m - qn > 0$ in the argument of the delta function (requiring ${k_{||}} < 0$ for $\ell = 0$). As a result, writing it in a form allowing the speed integral to be performed gives

(5.5)\begin{equation}\delta [\omega {\tau _p} - 2\mathrm{\pi }(\ell + m - qn)] = v_{\omega /{k_{||}}}^2\delta (v - {v_{\omega /{k_{||}}}})/\omega v{\tau _p},\end{equation}

with

(5.6)\begin{equation}{v_{\omega /{k_{||}}}} = \omega v{\tau _p}/2\mathrm{\pi }(\ell + m - qn) > 0.\end{equation}

Then the exponential factor in the Maxwellian becomes

(5.7)\begin{equation}x_{\omega /{k_{||}}}^2 = {m_e}v_{\omega /{k_{||}}}^2/2{T_e},\end{equation}

where recall a high speed expansion of the like collision operator is used so $x_{\omega /{k_{||}}}^2 \gg 1$ in ${f_0}$. In addition, the barely passing do not contribute significantly to the current since $v{\tau _p}/4qR\mathop \to \limits_{k \to 1} {(2\varepsilon )^{ - 1/2}}\ell n(4/\sqrt {1 - {k^2}} ) \gg 1$. Therefore, because of the exponential from ${f_0}$, the freely passing $({k^2}\mathrm{\ \mathbin{\lower.3ex\hbox{$\buildrel< \over {\smash{\scriptstyle\sim}\vphantom{_x}}$}}\ }\varepsilon \ll 1)$ electrons make the dominant contribution. As a result, only $\ell = 0$ need be retained in

(5.8)\begin{equation}D \approx \frac{{\mathrm{\pi }{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}\lambda J_1^2(\eta )}}{{8m_e^2k_{ \bot \; }^2|{k_{||}}|}}\delta (v - {v_{\omega /{k_{||}}}}) \equiv {D_{\vec{k}}}\; \lambda J_1^2(\eta )\delta (v - {v_{\omega /{k_{||}}}}).\end{equation}

Defining the exponential factor of ${f_0}$ for $\ell = 0$ as ${X^2}(\lambda ) \equiv x_{\omega /{k_{||}}}^2(\ell = 0)$, gives

(5.9)\begin{equation}{X^2}(\lambda ) \approx ({\omega ^2}/k_{||}^2v_e^2)(1 + {k^2}/2\varepsilon ) \approx ({\omega ^2}/k_{||}^2v_e^2)[\lambda /(1 - \lambda )],\end{equation}

as $v{\tau _p}\mathop \to \limits_{{k^2} \ll 1} 2\mathrm{\pi }qR\sqrt {1 + {k^2}/2\varepsilon }$, ${v_{\omega /{k_{||}}}}\mathop \to \limits_{\ell = 0} \omega \sqrt {1 + {k^2}/2\varepsilon } /{k_{||}}$ and

(5.10)\begin{equation}\varTheta \mathop \to \limits_{\varepsilon \ll 1} {\left|{\int_{ - \mathrm{\pi }}^\mathrm{\pi } {\frac{{\textrm{d}\vartheta }}{{2\mathrm{\pi }}}\,{\textrm{e}^{ - \textrm{i}\ell \vartheta }}} } \right|^2} = {\delta _{0\ell }} = \left\{ {\begin{array}{@{}ll} 1& {\ell = 0}\\ 0& {\ell \ne 0} \end{array}} \right..\end{equation}

Then, integrating over ${v_{||}} < 0$ only, as D vanishes for ${v_{||}} > 0$, leads to

(5.11)\begin{align}\left\langle {\frac{B}{{{B_0}}}\int {{\textrm{d}^3}v{v_{||}}{f_1}} } \right\rangle ={-} 4\mathrm{\pi }R\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\frac{{{\omega ^2}}}{{k_{||}^2v_e^2}}{D_{\boldsymbol{k}}}} \int_0^{\lambda < 1 - \varepsilon } {\textrm{d}\lambda \lambda J_1^2(\eta )} \int_0^\infty {\textrm{d}v\delta (v - {v_{\omega /{k_{||}}}})} v{\left. {{f_0}\frac{\partial }{{\partial v}}} \right|_\mu }\left( {\frac{{{{\bar{h}}_p}}}{{{f_0}}}} \right).\end{align}

In $v,\lambda$ variables

(5.12)\begin{align}v{\left. {{f_0}\frac{\partial }{{\partial v}}} \right|_\mu }\left( {\frac{{{{\bar{h}}_p}}}{{{f_0}}}} \right) = \; v{f_0}\left[ {{{\left. {\frac{\partial }{{\partial v}}} \right|}_\lambda }\left( {\frac{{{{\bar{h}}_p}}}{{{f_0}}}} \right) + {{\left. {\frac{{\partial \lambda }}{{\partial v}}} \right|}_\mu }{{\left. {\frac{\partial }{{\partial \lambda }}} \right|}_v}\left( {\frac{{{{\bar{h}}_p}}}{{{f_0}}}} \right)} \right] = \left( {4 - \frac{{2\lambda }}{{{\varLambda_{1 + 2}}}}\frac{{\partial {\varLambda_{1 + 2}}}}{{\partial \lambda }}} \right){\bar{h}_p},\end{align}

since the lowest-order solution is ${\bar{h}_p}/{f_0} \propto {v^4}{\varLambda _{1 + 2}}(\lambda )$. The exponential in the Maxwellian makes the evaluation of the $\lambda$ integral insensitive to its upper limit $(\lambda < 1 - \varepsilon )$ as will be shown shortly. The pitch angle derivative of ${\varLambda _{1 + 2}}$ is finite at $\lambda = 0$ making $\varLambda _{1 + 2}^{ - 1}\partial {\varLambda _{1 + 2}}/\partial \lambda \sim 1$.

Only the $\lambda = 0$ limit of this term will matter when the pitch angle integral is performed in (5.18) by integrating by parts, leaving just

(5.13)\begin{equation}\left\langle {\frac{B}{{{B_0}}}\int {{d^3}v{v_{||}}{f_1}} } \right\rangle ={-} 16\mathrm{\pi }R\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\frac{{{\omega ^2}}}{{k_{||}^2v_e^2}}{D_{\boldsymbol{k}}}} \int_0^{\lambda < 1 - \varepsilon } {\textrm{d}\lambda \lambda J_1^2(\eta ){{\bar{h}}_p}} ,\end{equation}

to be evaluated, where at resonance $\eta = \sqrt \lambda {k_ \bot }\omega /{k_{||}}{\varOmega _e}$ and

(5.14)\begin{equation}{\bar{h}_p} = \; \frac{{4{n_e}{\omega ^4}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}{\varLambda _{1 + 2}}(\sqrt \varepsilon ,Z,\lambda )\,{\textrm{e}^{ - {X^2}(\lambda )}}}}{{3{\mathrm{\pi }^2}R{\nu _{\textrm{ee}}}k_{||}^4v_e^6[(Z + 1)(1 + 1.46\sqrt {2\varepsilon } ) + 4]}}.\end{equation}

Inserting ${\bar{h}_p}$ leads to the expression for the parallel current driven by helicon waves to be

(5.15)\begin{align}\begin{aligned} J_{||}^H & ={-} e\left\langle {\dfrac{B}{{{B_0}}}\int {{\textrm{d}^3}v{v_{||}}{f_1}} } \right\rangle = \dfrac{{8e{n_e}}}{{(Z + 1)(1 + 1.46\sqrt {2\varepsilon } ) + 4}}\\ & \quad \times\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\dfrac{{{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}}}{{3m_e^2k_{ \bot \; }^2}}} \dfrac{{{\omega ^6}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{{\nu _{\textrm{ee}}}|k_{||}^7|v_e^8}}\int_0^{\lambda < 1 - \varepsilon } {\textrm{d}\lambda \lambda J_1^2(\eta ){\varLambda _{1 + 2}}(\lambda )\,{\textrm{e}^{ - {X^2}(\lambda )}}} , \end{aligned}\end{align}

where $\eta \sim \sqrt \lambda {k_ \bot }{v_e}/{\varOmega _e} = \sqrt \lambda {k_ \bot }{\rho _e}$. As a result, ${X^2}(\lambda )$ provides a cut-off before $\lambda \to 1 - \varepsilon$. However, the procedure is more complex in the HCD case than in the LHCD case because of the Bessel function behaviour due the helicon wave interacting with gyromotion of the electron, while the lower hybrid wave only interacts with its parallel streaming.

To understand the procedure and generalize the LHCD results, it is helpful to first consider the LHCD case in more detail than was given in Catto (Reference Catto2021). To obtain the parallel current driven by the LH waves, $J_{||}^{\textrm{LH}}$, the preceding steps are repeated starting with the replacement $k_ \bot ^{ - 2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}v_ \bot ^2J_1^2(\eta ) \to |{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n}{|^2}v_{||}^2J_0^2(\eta )$ and then using $v_ \bot ^2 \approx \lambda {v^2} \approx \lambda {\omega ^2}/k_{||}^2 \to v_{||}^2 \approx {\omega ^2}/k_{||}^2$. As a result, the driven parallel LH current is

(5.16)\begin{align}J_{||}^{\textrm{LH}} = \frac{{8e{n_e}}}{{(Z + 1)(1 + 1.46\sqrt {2\varepsilon } ) + 4}}\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\frac{{{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n}{|^2}}}{{3m_e^2}}} \frac{{{\omega ^6}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{{\nu _{\textrm{ee}}}|k_{||}^7|v_e^8}}\int_0^{\lambda < 1 - \varepsilon } {\textrm{d}\lambda J_0^2(\eta ){\varLambda _{1 + 2}}(\lambda )\,{\textrm{e}^{ - {X^2}(\lambda )}},} \end{align}

where again $\eta = \sqrt \lambda {k_ \bot }\omega /{k_{||}}{\varOmega _e}$. Then, an integration by parts is performed by using ${\textrm{e}^{ - {X^2}(\lambda )}} ={-} [{(1 - \lambda )^2}/z]\partial \,{\textrm{e}^{ - {X^2}(\lambda )}}/\partial \lambda$ to obtain an asymptotic expansion in inverse powers of $z = {\omega ^2}/k_{||}^2v_e^2 \gg 1$ to find

(5.17)\begin{equation}J_{||}^{\textrm{LH}} \approx \frac{{8e{n_e}{\varLambda _{1 + 2}}(\sqrt \varepsilon ,Z,0)}}{{(Z + 1)(1 + 1.46\sqrt {2\varepsilon } ) + 4}}\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\frac{{{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n}{|^2}}}{{3m_e^2}}} \frac{{{\omega ^4}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{{\nu _{\textrm{ee}}}|k_{||}^5|v_e^6}},\end{equation}

as $\int_0^{\lambda < 1 - \varepsilon } {\textrm{d}\lambda J_0^2{\varLambda _{1 + 2}}\,{\textrm{e}^{ - {X^2}}}} \approx {\varLambda _{1 + 2}}(\lambda = 0)/z$ since ${\varLambda _j}(0) = 1 = J_0^2(0)$. This result is 2/3 of the value given by Catto (Reference Catto2021) when ${\varLambda _2}$ is neglected. The difference is due to the extra power of ${v^2}$ multiplying ${\bar{h}_p}$ inside the v derivative beginning of his (4.7). Notice that only terms that remain finite at $\lambda = 0$ will contribute, as assumed in (5.13).

In the helicon case three integrations by parts in $\lambda$ are required. Again, only terms finite at $\lambda = 0$ contribute. Ignoring exponentially small terms

(5.18)\begin{align}\begin{aligned} & \int_0^{\lambda < 1 - \varepsilon } {\textrm{d}\lambda \lambda J_1^2{\varLambda _{1 + 2}}\,{\textrm{e}^{ - {X^2}}}} \approx \dfrac{1}{{{z^2}}}\int_0^{\lambda < 1 - \varepsilon } {\textrm{d}\lambda \,{\textrm{e}^{ - {X^2}}}} \dfrac{\partial }{{\partial \lambda }}\left\{ {{{(1 - \lambda )}^2}\dfrac{\partial }{{\partial \lambda }}[{{(1 - \lambda )}^2}\lambda J_1^2{\varLambda _{1 + 2}}]} \right\}\\ & \quad \approx \dfrac{1}{{{z^3}}}\dfrac{\partial }{{\partial \lambda }}{\left. {\left\{ {{{(1 - \lambda )}^2}\dfrac{\partial }{{\partial \lambda }}[{{(1 - \lambda )}^2}\lambda J_1^2{\varLambda _{1 + 2}}]} \right\}} \right|_{\lambda = 0}} = \dfrac{1}{{{z^3}}}{\left. {\dfrac{{\partial J_1^2}}{{\partial \lambda }}} \right|_{\lambda = 0}} = \dfrac{{k_ \bot ^2\rho _e^2}}{{2{z^2}}}. \end{aligned}\end{align}

Therefore, for HCD

(5.19)\begin{equation}J_{||}^H = \frac{{4e{n_e}{\varLambda _{1 + 2}}(\sqrt \varepsilon ,Z,0)}}{{(Z + 1)(1 + 1.46\sqrt {2\varepsilon } ) + 4}}\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\frac{{{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}}}{{3m_e^2k_{ \bot \; }^2}}} \frac{{{\omega ^2}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{{\nu _{\textrm{ee}}}|k_{||}^3|v_e^4}}k_ \bot ^2\rho _e^2.\end{equation}

The stronger ${T_e}$ dependence of $J_{||}^H$ compared with $J_{||}^{\textrm{LH}}$ was previously noted by Li et al. (Reference Li, Ding, Dong and Liu2020a, Reference Li, Li and Liu2021). Helicon waves are also sensitive to the density. If ${n_e}$ increases substantially, (A14) indicates $k_{||}^2$ will increase, causing a less effective rf to interaction with the bulk electrons.

Letting $z = {\omega ^2}/k_{||}^2v_e^2 = {c^2}/v_e^2n_{||}^2$, then HCD and LHCD depend on ${z^{3/2}}{e^{ - z}}$ and ${z^{5/2}}{e^{ - z}}$, respectively, and are maximized at z = 3/2 and z = 5/2, consistent with assuming ${v^2} > v_e^2$ in ${C_{\textrm{ee}}}$. Consequently, a slightly higher parallel refractive index is desirable for HCD compared with LHCD, as noted by Prater et al. (Reference Prater, Moeller, Pinsker, Porkolab, Meneghini and Vdovin2014). In addition to penetrating into the central core plasma, the current driven by helicon waves can be comparable to or larger than that driven by lower hybrid waves for the same ${\omega ^2}/k_{||}^2v_e^2$ based on the estimate of (A15) and simulations (Vdovin Reference Vdovin2013; Prater et al. Reference Prater, Moeller, Pinsker, Porkolab, Meneghini and Vdovin2014; Lau et al. Reference Lau, Jaeger, Bertelli, Berry, Green, Murakami, Park, Pinsker and Prater2018, Reference Lau, Berry, Jaeger and Bertelli2019; Li et al. Reference Li, Ding, Dong and Liu2020a,Reference Li, Liu, Xiang and Lib, Reference Li, Li and Liu2021).

6. The rf power and current drive efficiency

To form the HCD efficiency requires evaluating the rf power absorbed by the passing electrons. Integrating over ${v_{||}} < 0$ and ignoring the negligible power into the barely passing leads to

(6.1)\begin{align}\begin{aligned} P_{\textrm{cd}}^H & = \dfrac{{{m_e}}}{2}\left\langle {\int {{\textrm{d}^3}v{v^2}Q} } \right\rangle \approx \dfrac{{{m_e}{B_0}}}{{4\mathrm{\pi }qR}}\int {{\textrm{d}^3}v\dfrac{{{v_{||}}}}{B}{v^2}{\tau _p}\bar{Q}} \\ & = \dfrac{{{m_e}{B_0}}}{{4\mathrm{\pi }qR}}\int {{\textrm{d}^3}v\dfrac{{{v_{||}}}}{B}v} {\left. {\dfrac{\partial }{{\partial v}}} \right|_\mu }\left( {{\tau_p}vD{{\left. {\dfrac{{\partial {f_0}}}{{\partial v}}} \right|}_\mu }} \right) = \dfrac{{m_e^2}}{{qR{T_e}}}\int_0^{\lambda < 1 - \varepsilon } {\textrm{d}\lambda } \int_0^\infty {\textrm{d}v{\tau _p}{v^5}D{f_0}} \\ & = \dfrac{{{\mathrm{\pi }^{1/2}}}}{2}{m_e}{n_e}\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\dfrac{{{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}{\omega ^4}}}{{m_e^2k_{ \bot \; }^2|k_{||}^5|v_e^5}}\,} {\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}\int_0^{\lambda < 1 - \varepsilon } {\textrm{d}\lambda \lambda J_1^2(\eta )\,{\textrm{e}^{ - {X^2}(\lambda )}}} . \end{aligned}\end{align}

Similarly, fixing the ${v^2}$ and the numerical factor in Catto (Reference Catto2021) for LHCD, gives

(6.2)\begin{equation}P_{\textrm{cd}}^{\textrm{LH}} = \frac{{{\mathrm{\pi }^{1/2}}}}{2}{m_e}{n_e}\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\frac{{{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n}{|^2}{\omega ^4}}}{{m_e^2|k_{||}^5|v_e^5}}\,} {\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}\int_0^{1 - \varepsilon} {\textrm{d}\lambda J_0^2(\eta )\,{\textrm{e}^{ - {X^2}(\lambda )}}} .\end{equation}

Again using ${\textrm{e}^{ - {X^2}(\lambda )}}$ to integrate by parts gives

(6.3)\begin{equation}P_{\textrm{cd}}^H = \frac{{{\mathrm{\pi }^{1/2}}}}{4}{m_e}{n_e}\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\frac{{{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}}}{{m_e^2k_{ \bot \; }^2|{k_{||}}|{v_e}}}} k_ \bot ^2\rho _e^2\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}},\end{equation}

and

(6.4)\begin{equation}P_{\textrm{cd}}^{\textrm{LH}} = \frac{{{\mathrm{\pi }^{1/2}}}}{2}{m_e}{n_e}\mathop \sum \limits_{\vec{k}}^{{k_{||}} < 0} \frac{{{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n}{|^2}{\omega ^2}}}{{m_e^2|k_{||}^3|v_e^3}}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}.\end{equation}

The numerical coefficient of $P_{\textrm{cd}}^{\textrm{LH}}$ is half that of (5.7) in Catto (Reference Catto2021).

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

(6.5)\begin{equation}\frac{{J_{||}^H}}{{P_{\textrm{cd}}^H}} = \frac{{16e\{ 1 + [0.62 - 1.02(Z + 5)/(7Z + 11)]\sqrt \varepsilon \} \sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\frac{{{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}{\omega ^2}}}{{{\nu _{\textrm{ee}}}m_e^2k_{ \bot \; }^2|k_{||}^3|v_e^4}}} \; k_ \bot ^2\rho _e^2\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{3{\mathrm{\pi }^{1/2}}{m_e}[(Z + 1)(1 + 1.46\sqrt {2\varepsilon } ) + 4]\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\frac{{{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}}}{{m_e^2k_{ \bot \; }^2|{k_{||}}|{v_e}}}} k_ \bot ^2\rho _e^2\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}},\end{equation}

while

(6.6) \begin{equation}\frac{{J_{||}^{\textrm{LH}}}}{{P_{\textrm{cd}}^{\textrm{LH}}}} = \frac{{16e\{ 1 + [0.62 - 1.02(Z + 5)/(7Z + 11)]\sqrt \varepsilon \} \sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0}{{\frac{{{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n}{|^2}{\omega ^4}}}{{{\nu _{\textrm{ee}}}m_e^2|k_{||}^5|v_e^6}}\,} } {\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{3{\mathrm{\pi }^{1/2}}{m_e}[(Z + 1)(1 + 1.46\sqrt {2\varepsilon } ) + 4]\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\frac{{{e^2}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n}{|^2}{\omega ^2}}}{{m_e^2|k_{||}^3|v_e^3}}} \,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}.\end{equation}

In the terms in the sums in these two general forms and in (5.17), (5.19), (6.1) and (6.2), the Fourier mode amplitudes $|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}$ and $|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n}{|^2}$ and the wavenumbers ${k_ \bot }$ and ${k_{||}}$ are to be obtained from a full wave code for precise evaluations.

Remarkably, for any (not necessarily the same) single $\omega$ and $\boldsymbol{k}$ the preceding forms are identical

(6.7)\begin{align}\frac{{J_{||}^H/e{n_e}{v_e}}}{{P_{\textrm{cd}}^H/{n_e}{m_e}v_e^2{\nu _{\textrm{ee}}}}} = \frac{{16\{ 1 + [0.62 - 1.02(Z + 5)/(7Z + 11)]\sqrt \varepsilon \} {\omega ^2}}}{{3{\mathrm{\pi }^{1/2}}[(Z + 1)(1 + 1.46\sqrt {2\varepsilon } ) + 4]k_{||}^2v_e^2}} = \frac{{J_{||}^{\textrm{LH}}/e{n_e}{v_e}}}{{P_{\textrm{cd}}^{\textrm{LH}}/{n_e}{m_e}v_e^2{\nu _{\textrm{ee}}}}}.\end{align}

Approximation (6.7) suggests that comparable current drive efficiencies are possible with helicon and lower hybrid waves, with HCD providing core access as well as profile control. In addition, based on the cold plasma estimate (A15) of the Appendix, it seems possible to drive comparable parallel currents with helicon and lower hybrid waves, as suggested by the simulations of Prater et al. (Reference Prater, Moeller, Pinsker, Porkolab, Meneghini and Vdovin2014) and Lau et al. (Reference Lau, Jaeger, Bertelli, Berry, Green, Murakami, Park, Pinsker and Prater2018). The $\varepsilon = 0$ form of (6.7) is essentially the same as the non-relativistic, $z = v_p^2 \gg 1$ limit (31) in Karney & Fisch (Reference Karney and Fisch1985).

7. Combined helicon and lower hybrid

The preceding idealized analytic evaluations of the currents driven and the form of the QL operator suggest it is possible to combine HCD with LHCD for the same applied resonant wave frequency (Yin et al. Reference Yin, Zheng, Gong, Yang, Yin, Song, Huang, Chen and Zhong2022). Helicon and lower hybrid waves then drive current by simultaneously acting on the perpendicular and parallel electron distribution function, respectively. Indeed, an antenna used for LHCD can drive some level of helicon waves and vice versa. Retaining only $\ell = 0$, using the approximation ${\textrm{e}^{ - \textrm{i}\int_{{\tau _0}}^\tau {\textrm{d}\tau ^{\prime}\chi (\tau ^{\prime})} }} \approx 1$ and inserting the delta function, the combined QL diffusivity for the passing electrons becomes

(7.1)\begin{equation}D = \frac{{\mathrm{\pi }{e^2}}}{{2m_e^2{v^2}{\tau _p}}}{\left|{\oint_p {\textrm{d}\tau } {\boldsymbol{e}_k}\boldsymbol{\cdot }\left[ {\boldsymbol{n}{v_{||}}{J_0}(\eta ) - \textrm{i}\boldsymbol{n} \times \boldsymbol{k}\frac{{{v_ \bot }}}{{{k_ \bot }}}{J_1}(\eta )} \right]} \right|^2}\frac{{v_{\omega /{k_{||}}}^2\delta (v - {v_{\omega /{k_{||}}}})}}{{\omega v{\tau _p}}},\end{equation}

with

(7.2)\begin{equation}{v_{\omega /{k_{||}}}} = \; \omega v{\tau _p}/2\mathrm{\pi (}m - qn\textrm{)} > 0.\end{equation}

The combined parallel current that can be driven is therefore

(7.3)\begin{align}\begin{aligned} J_{||}^{H + \textrm{LH}} & = \dfrac{{8{e^3}{n_e}\{ 1 + [0.62 - 1.02(Z + 5)/(7Z + 11)]\sqrt \varepsilon \} }}{{3m_e^2{\nu _{\textrm{ee}}}[(Z + 1)(1 + 1.46\sqrt {2\varepsilon } ) + 4]}}\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\dfrac{{{\omega ^6}\,{\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}}}{{|k_{||}^7|v_e^8}}} \left\{\vphantom{\dfrac{{k_{||}^2v_e^2\rho_e^2}}{{2{\omega^2}}}}{|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n}{|^2}} \right.\\ & \quad \left. { + \,i\dfrac{{|{k_{||}}|{v_e}{\rho_e}}}{{2\omega }}[({\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n})(\boldsymbol{e}_k^\ast \boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}) - (\boldsymbol{e}_k^\ast \boldsymbol{\cdot }\boldsymbol{n})({\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k})] + \dfrac{{k_{||}^2v_e^2\rho_e^2}}{{2{\omega^2}}}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}} \right\}, \end{aligned}\end{align}

where the integrals are performed by integration by parts as before, but with one new integral appearing in the cross terms, namely

(7.4)\begin{align}\int_0^{\lambda < 1 - \varepsilon } {\textrm{d}\lambda {\lambda ^{1/2}}{J_1}(\eta ){J_0}(\eta ){\varLambda _{1 + 2}}(\lambda )\,{\textrm{e}^{ - {X^2}(\lambda )}}} \approx \frac{1}{z}{\left. {\frac{{\partial ({\lambda^{1/2}}{J_1})}}{{\partial \lambda }}} \right|_{\lambda = 0}} = \frac{{{k_ \bot }{\rho _e}}}{{2{z^{3/2}}}}{\varLambda _{1 + 2}}(\sqrt \varepsilon ,Z,\lambda = 0).\end{align}

The rf power absorbed by passing electrons is

(7.5)\begin{align}\begin{aligned} P_{\textrm{cd}}^{H + \textrm{LH}} & = \dfrac{{{\mathrm{\pi }^{1/2}}{e^2}{n_e}}}{{2{m_e}}}\sum\limits_{\boldsymbol{k}}^{{k_{||}} < 0} {\dfrac{{{\omega ^2}}}{{|k_{||}^3|v_e^3}}\,} {\textrm{e}^{ - {\omega ^2}/k_{||}^2v_e^2}}\left\{\vphantom{\dfrac{{k_{||}^2v_e^2\rho_e^2}}{{2{\omega^2}}}}{|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n}{|^2}} \right.\\ & \quad \left. { + \,i\dfrac{{|{k_{||}}|{v_e}{\rho_e}}}{{2\omega }}[{({\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n})(\boldsymbol{e}_k^\ast \boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}) - (\boldsymbol{e}_k^\ast \boldsymbol{\cdot }\boldsymbol{n})({\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k})} ]+ \dfrac{{k_{||}^2v_e^2\rho_e^2}}{{2{\omega^2}}}|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}} \right\}. \end{aligned}\end{align}

As a result, the efficiency is the ratio $J_{||}^{H + \textrm{LH}}/P_{\textrm{cd}}^{H + \textrm{LH}}$, which for a single frequency and wavenumber recovers the same expression as given at the end of § 6. Moreover, the cross term may allow more current to be driven and more power to be absorbed, perhaps as observed in the EAST numerical simulations of Yin et al. (Reference Yin, Zheng, Gong, Yang, Yin, Song, Huang, Chen and Zhong2022). Synergy is possible, but not assured, since helicon and lower hybrid waves independently act on perpendicular and parallel motion, respectively, to make the electrons less collisional.

8. Summary

The new results are the analytic expressions for the parallel current that can be driven in a tokamak by a helicon wave, (5.19), and the associated efficiency of HCD, (6.5) and (6.7). In addition, the numerical coefficients of the corresponding tokamak expressions for LHCD (Catto Reference Catto2021), (5.17), (6.4), (6.6) and (6.7), are generalized as well as corrected, by using a more systematic derivation. Interestingly, for any single applied frequency and wave vector the efficiency of HCD and LHCD are shown to be the same, (6.7). In addition, HCD and LHCD can be combined for the same applied frequency as shown in (7.3) and (7.5) of § 7, and, of course, recover the same single wave vector efficiency, (6.7). If a combination of HCD and LHCD is possible from the same antenna and rf source, more current than the sum might be driven and steady state tokamak operation might be slightly more feasible.

Acknowledgements

The United States Government retains a non-exclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this manuscript, or allow others to do so, for United States Government purposes.

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

Funding

The authors are grateful to P. Bonoli, A. Ram and M. Porkolab for valuable suggestions and stimulating exchanges. This work was supported by the U.S. Department of Energy under contract number DE-FG02-91ER-54109, and research at Princeton University was supported by a Princeton Presidential Postdoctoral Research Fellowship.

Declaration of interests

The authors report no conflict of interest.

Appendix A. Helicon and lower hybrid waves in a cold plasma

To highlight the differences between helicon and lower hybrid waves in the simplest fashion, cold plasma theory can be employed. Helicon current drive employs a perpendicular component of the applied electric field, while the parallel component is used for LHCD.

In a cold plasma, Maxwell's equations lead to the need for the Fourier transform of the applied electric field, ${\boldsymbol{e}_k}$, to satisfy

(A1)\begin{equation}[\boldsymbol{\varepsilon } - {n^2}(\boldsymbol{I} - {k^2}\boldsymbol{kk})]\boldsymbol{\cdot }{\boldsymbol{e}_k} = 0,\end{equation}

with $\boldsymbol{k}$ a wave vector, $\boldsymbol{k} = |\boldsymbol{k}|$, $\boldsymbol{I}$ the unit dyad and $n = kc/\omega$ the index of refraction, and the dielectric tensor written as

(A2)\begin{equation}\boldsymbol{\varepsilon } = {\varepsilon _ \bot }(\boldsymbol{I} - \boldsymbol{nn}) + {\varepsilon _{||}}\boldsymbol{nn} - i{\varepsilon _ \times }\boldsymbol{n} \times \boldsymbol{I},\end{equation}

where the magnetic field is $\boldsymbol{B} = B\boldsymbol{n}$. The components of the dielectric tensor are defined as

(A3)\begin{gather}{\varepsilon _ \bot } = 1 - \sum\limits_s {\frac{{\omega _{\textrm{ps}}^2}}{{{\omega ^2} - \varOmega _s^2}}} \approx 1 + \frac{{\omega _{\textrm{pe}}^2}}{{\varOmega _e^2}} - \frac{{\omega _{\textrm{pi}}^2}}{{{\omega ^2}}} = 1 + \frac{{\omega _{\textrm{pe}}^2}}{{\varOmega _e^2}}\left( {1 - \frac{{{\varOmega_i}{\varOmega_e}}}{{{\omega^2}}}} \right),\end{gather}
(A4)\begin{gather}{\varepsilon _{||}} = 1 - \sum\limits_s {\frac{{\omega _{\textrm{ps}}^2}}{{{\omega ^2}}}} \approx{-} \frac{{\omega _{\textrm{pe}}^2}}{{{\omega ^2}}},\end{gather}
(A5)\begin{gather}{\varepsilon _ \times } ={-} \sum\limits_s {\frac{{\omega _{\textrm{ps}}^2{\varOmega _s}}}{{\omega ({\omega ^2} - \varOmega _s^2)}}} \approx{-} \frac{{\omega _{\textrm{pe}}^2}}{{\omega {\varOmega _e}}} - \frac{{\omega _{\textrm{pi}}^2{\varOmega _i}}}{{{\omega ^3}}} \approx{-} \frac{{\omega _{\textrm{pe}}^2}}{{\omega {\varOmega _e}}},\end{gather}

with $\omega _{\textrm{pe}}^2{\varOmega _i} = \omega _{\textrm{pi}}^2{\varOmega _e}$, where the approximate forms are valid for ${\varOmega _e} \gg \omega \gg {\varOmega _i} = {Z_i}eB/{m_i}c$, which is the frequency range of interest here. The species (s) plasma frequency is defined as $\omega _{\textrm{ps}}^2 = 4\mathrm{\pi }Z_s^2{e^2}{n_s}/{m_s}$, and ${\beta _e} = v_e^2\omega _{\textrm{pe\; }}^2/\varOmega _e^2{c^2}$.The orderings give $\varepsilon _ \times ^2/{\varepsilon _{||}}\sim \omega _{pe}^2/\varOmega _e^2\sim 1$ and ${\varepsilon _{||}} \gg {\varepsilon _ \times } \gg {\varepsilon _ \bot }\sim 1$. Moreover, for a resonant Landau interaction $\omega /{k_{||}}{v_e} > 1$ suggesting the ordering ${n_{||}} = {k_{||}}c/\omega \mathrm{\ \mathbin{\lower.3ex\hbox{$\buildrel> \over {\smash{\scriptstyle\sim}\vphantom{_x}}$}}\ }1$.

Letting

(A6)\begin{equation}{\boldsymbol{e}_k} = {e_{||}}\boldsymbol{n} + {e_ \bot }({\boldsymbol{k}_ \bot } + \varUpsilon \boldsymbol{n} \times \boldsymbol{k})/{k_ \bot }\sqrt {1 + |\varUpsilon {|^2}} ,\end{equation}

the three components of the cold plasma equation are related by

(A7)\begin{equation}{e_{||}}/{e_ \bot } = {n_{||}}{n_ \bot }/[(n_ \bot ^2 - {\varepsilon _{||}})\sqrt {1 + |\varUpsilon {|^2}} ],\end{equation}

and

(A8)\begin{equation}\varUpsilon = i{\varepsilon _ \times }/({\varepsilon _ \bot } - {n^2}),\end{equation}

and the cold plasma dispersion relation is

(A9)\begin{equation}n_{||}^2{\varepsilon _{||}}/({\varepsilon _{||}} - n_ \bot ^2) = {\varepsilon _ \bot } - \varepsilon _ \times ^2/({\varepsilon _ \bot } - {n^2}),\end{equation}

which when rewritten in powers of $n_ \bot ^2$ yields

(A10)\begin{equation}{\varepsilon _ \bot }n_ \bot ^4 + [({\varepsilon _{||}} + {\varepsilon _ \bot })(n_{||}^2 - {\varepsilon _ \bot }) + \varepsilon _ \times ^2]n_ \bot ^2 + {\varepsilon _{||}}[{(n_{||}^2 - {\varepsilon _ \bot })^2} - \varepsilon _ \times ^2] = 0.\end{equation}

The two distinct branches satisfying accessibility $(n_ \bot ^2 > 0)$ are found by assuming ${[({\varepsilon _{||}} + {\varepsilon _ \bot })(n_{||}^2 - {\varepsilon _ \bot }) + \varepsilon _ \times ^2]^2} \gg 4{\varepsilon _ \bot }{\varepsilon _{||}}[{(n_{||}^2 - {\varepsilon _ \bot })^2} - \varepsilon _ \times ^2]$ to find

(A11)\begin{equation}n_ \bot ^2 \approx \left\{ {\begin{array}{@{}ll} { - [({\varepsilon_{||}} + {\varepsilon_ \bot })(n_{||}^2 - {\varepsilon_ \bot }) + \varepsilon_ \times^2]/{\varepsilon_ \bot }}& {\textrm{lower}\;\textrm{hybrid/slow}}\\ { - {\varepsilon_{||}}[{{(n_{||}^2 - {\varepsilon_ \bot })}^2} - \varepsilon_ \times^2]/[({\varepsilon_{||}} + {\varepsilon_ \bot })(n_{||}^2 - {\varepsilon_ \bot }) + \varepsilon_ \times^2]}& {\textrm{helicon/fast}} \end{array}} \right..\end{equation}

For the lower hybrid branch $\varepsilon _ \times ^2 + {\varepsilon _ \bot }n_ \bot ^2 + {\varepsilon _{||}}(n_{||}^2 - {\varepsilon _ \bot }) \approx 0$. Keeping $|\varUpsilon {|^2} \ll 1$, requires ${\textrm{n}^2} \gg {\mathrm{\varepsilon }_ \times } \gg {\mathrm{\varepsilon }_ \bot }$, leading to

(A12)\begin{equation}{\omega ^2} = \frac{{{\varOmega _i}{\varOmega _e}(1 + \omega _{\textrm{pe}}^2/k_ \bot ^2{c^2} + k_{||}^2{\varOmega _e}/k_ \bot ^2{\varOmega _i})}}{{(1 + \varOmega _e^2/\omega _{\textrm{pe}}^2)(1 + \omega _{\textrm{pe}}^2/k_ \bot ^2{c^2}) + \omega _{\textrm{pe}}^2/k_ \bot ^2{c^2}}} \approx \frac{{{\varOmega _i}{\varOmega _e}(1 + k_{||}^2{\varOmega _e}/k_ \bot ^2{\varOmega _i})}}{{1 + \varOmega _e^2/\omega _{\textrm{pe}}^2}},\end{equation}

where the last form assumes $k_ \bot ^2{c^2} \gg \omega _{\textrm{pe}}^2$ or $n_ \bot ^2 \gg {\varepsilon _{||}}$ to further increase the size ${e_{||}}$ to allow ${e_{||}}/{e_ \bot } \approx {k_{||}}/{k_ \bot }$. Lower hybrid accessibility requires $n_{||}^2 > 1 + \omega _{\textrm{pe}}^2/\varOmega _e^2$ (Golant Reference Golant1972), which does not allow it to penetrate to the core in reactor relevant tokamaks.

The helicon or whistler branch can be approximated by

(A13)\begin{equation}n_ \bot ^2 = [\varepsilon _ \times ^2 - {(n_{||}^2 - {\varepsilon _ \bot })^2}]/[(n_{||}^2 - {\varepsilon _ \bot }) + \varepsilon _ \times ^2/{\varepsilon _{||}}].\end{equation}

Substituting in the components of the dielectric tensor and rewriting gives

(A14)\begin{equation}{\omega ^2} = \frac{{{\varOmega _i}{\varOmega _e}(1 + k_{||}^2{c^2}/\omega _{\textrm{pi}}^2)}}{{1 + k_ \bot ^2/{k^2} + \varOmega _e^2/\omega _{\textrm{pe}}^2 + \omega _{\textrm{pe}}^2/{k^2}{c^2}}} \to {\varOmega _i}{\varOmega _e}\left( {\frac{{k_{||}^2{c^2}}}{{\omega_{\textrm{pi}}^2}}} \right)\left( {\frac{{{k^2}{c^2}}}{{\omega_{\textrm{pe}}^2}}} \right),\end{equation}

where the first form allows $k_ \bot ^2{c^2}\sim \omega _{\textrm{pe}}^2$, while the last assumes $\omega _{\textrm{pe}}^2/{k^2}{c^2} \gg \; 2 + \varOmega _e^2/\omega _{\textrm{pe}}^2$ and $k_{||}^2{c^2}/\omega _{\textrm{pi}}^2 \gg 1$ to recover the form in Preinhaelter & Vaclavik (Reference Preinhaelter and Václavík1967) and used in de Assis & Busnardo-Neto (Reference de Assis and Busnardo-Neto1988). The last form is sometimes referred to as an oblique whistler wave. Moreover, helicon waves can propagate in the high harmonic fast wave (HHFW) range of frequencies (Ono Reference Ono1995; Lashmore-Davies et al. Reference Lashmore-Davies, Fuchs and Cairns1998) and are sometimes referred to as HHFWs (Yin et al. Reference Yin, Zheng, Gong, Yang, Yin, Song, Huang, Chen and Zhong2022).

For the same value of ${\omega ^2}/k_{||}^2v_e^2$ the ratio of the parallel currents driven by helicon and lower hybrid waves (or the power absorbed) is roughly

(A15)\begin{align}\begin{aligned} & \dfrac{{J_{||}^H}}{{J_{||}^{\textrm{LH}}}}\sim \dfrac{{\rho _e^2|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n} \times \boldsymbol{k}{|^2}}}{{|{\boldsymbol{e}_k}\boldsymbol{\cdot }\boldsymbol{n}{|^2}}}\sim \dfrac{{k_ \bot ^2\rho _e^2|\varUpsilon |\; e_ \bot ^2}}{{(1 + |\varUpsilon {|^2})e_{||}^2}}\sim k_ \bot ^2\rho _e^2\dfrac{{{\varepsilon _ \times }{{(n_ \bot ^2 - {\varepsilon _{||}})}^2}}}{{\; n_ \bot ^2n_{||}^2({n^2} - {\varepsilon _ \bot })}}\\ & \quad \sim k_ \bot ^2\rho _e^2\dfrac{{\omega \omega _{\textrm{pe}}^2{{(\omega _{\textrm{pe}}^2 + k_ \bot ^2{c^2})}^2}}}{{{\varOmega _e}k_{||}^2\; {c^2}k_ \bot ^2{c^2}{k^2}{c^2}}}, \end{aligned}\end{align}

which can be order unity.

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. 2021 Lower hybrid current drive in a tokamak for correlated passes through resonance. J. Plasma Phys. 87, 905870309.CrossRefGoogle Scholar
Catto, P.J. & Tolman, E.A. 2021 a Reimagining full wave quasilinear theory in a tokamak. J. Plasma Phys. 87, 905870215.CrossRefGoogle Scholar
Catto, P.J. & Tolman, E.A. 2021 b Collisional broadening of nonlinear resonant wave-particle interactions. J. Plasma Phys. 87, 905870606.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
de Assis, A.S. & Busnardo-Neto, J. 1988 Current drive with whistler waves. Phys. Fluids 31, 943945.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
Golant, V.E. 1972 Plasma penetration near the lower hybrid frequency. Sov. Phys. – Tech. Phys. 16, 19801988.Google 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
Lashmore-Davies, C.N., Fuchs, V. & Cairns, R.A. 1998 A full wave theory of high-harmonic fast wave absorption in high-beta plasmas. Phys. Plasmas 5, 22842290.CrossRefGoogle Scholar
Lau, C., Berry, L.A., Jaeger, E.F. & Bertelli, N. 2019 Cold plasma finite element wave model for helicon waves. Plasma Phys. Control. Fusion 61, 045008.CrossRefGoogle Scholar
Lau, C., Bookman, M., Dimits, A., Dudson, B., Martin, E., Pinsker, R.I., Thomas, M. & van Compernolle, B. 2021 Helicon full wave modeling with scrape off layer turbulence on the DIII-D tokamak. Nucl. Fusion 61, 126072.CrossRefGoogle Scholar
Lau, C., Jaeger, E.F., Bertelli, N., Berry, L.A., Green, D.L., Murakami, M., Park, J.M., Pinsker, R.I. & Prater, R. 2018 AORSA full wave calculation of helicon waves in DIII-D and ITER. Nucl. Fusion 58, 066004.CrossRefGoogle Scholar
Li, J., Ding, X.T., Dong, J.Q. & Liu, S.F. 2020 a Helicon heating and current drive in toroidal plasmas. Plasma Phys. Control. Fusion 62, 095013.CrossRefGoogle Scholar
Li, X., Li, G. & Liu, H. 2021 Optimization of helicon wave off axis current drive in CFETR tokamak. Fusion Engng Des. 172, 112897.CrossRefGoogle Scholar
Li, X., Liu, H., Xiang, N. & Li, M. 2020 b Theoretical analysis of helicon wave current drive in EAST with high ${\beta _e}$ operation. Phys. Lett. A 384, 126779.CrossRefGoogle Scholar
Ono, M. 1995 High harmonic fast waves in high beta plasmas. Phys. Plasmas 2, 40754082.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
Pinsker, R.I. 2015 Whistlers, helicons, and lower hybrid waves: the physics of radio frequency wave propagation and absorption for current drive via Landau damping. Phys. Plasmas 22, 090901.CrossRefGoogle Scholar
Pinsker, R.I., Prater, R., Moeller, C.P., Degrassie, J.S., Petty, C.C., Porkolab, M., Anderson, J.P., Garofalo, A.M., Lau, C., Nagy, A., Pace, D.C., Torreblanca, H., Watkins, J.G. & Zeng, L. 2018 Experiments on helicons in DIII-D—investigation of the physics of a reactor-relevant non-inductive current drive technology. Nucl. Fusion 58, 106007.CrossRefGoogle Scholar
Prater, R., Moeller, C.P., Pinsker, R.I., Porkolab, M., Meneghini, O. & Vdovin, V.L. 2014 Application of very high harmonic fast waves for off-axis current drive in the DIII-D and FNSF-AT tokamaks. Nucl. Fusion 54, 083024.CrossRefGoogle Scholar
Preinhaelter, J. & Václavík, J. 1967 The quasilinear theory of beam instabilites of the low frequency non-electrostatic plasma waves. Plasma Phys. 9, 653664.CrossRefGoogle Scholar
Taguchi, M. 1983 The effect of trapped electron on the wave-induced current. J. Phys. Soc. Japan 52, 20352040.CrossRefGoogle Scholar
Vdovin, V.L. 2013 Current generation by helicons and lower hybrid waves in modern tokamaks and reactors ITER and DEMO. Plasma Phys. Rep. 39, 95119.CrossRefGoogle Scholar
Wang, S.J., Wi, H.H., Kim, H.J., Kim, J., Jeong, J.H. & Kwak, J.G. 2017 Helicon wave coupling in KSTAR plasmas for off axis current drive in high electron pressure plasmas. Nucl. Fusion 57, 046010.CrossRefGoogle Scholar
Xiao, Y., Catto, P.J. & Molvig, K. 2007 Collisional damping for ion temperature gradient mode driven zonal flow. Phys. Plasmas 14, 032302.CrossRefGoogle Scholar
Yin, L., Zheng, P., Gong, X., Yang, C., Yin, X., Song, C., Huang, Q., Chen, Y. & Zhong, Y. 2022 New synergy effects of the lower hybrid wave and the high harmonic fast wave current drive. Nucl. Fusion 62, 066023.CrossRefGoogle Scholar