
Citation: | Zhiwen CHENG, Guangyu WEI, Lei YE, Zhiyong QIU. Nonlinear saturation of reversed shear Alfvén eigenmode via high-frequency quasi-mode generation[J]. Plasma Science and Technology, 2025, 27(1): 015101. DOI: 10.1088/2058-6272/ad8814 |
A nonlinear saturation mechanism for reversed shear Alfvén eigenmode (RSAE) is proposed and analyzed, and is shown to be of relevance to typical reactor parameter region. The saturation is achieved through the generation of high-frequency quasi-mode due to nonlinear coupling of two RSAEs, which is then damped due to coupling with the shear Alfvén continuum, and leads to the nonlinear saturation of the primary RSAEs. An estimation of the nonlinear damping rate is also provided.
Energetic particles (EPs) including fusion alpha particles are of crucial importance in magnetically confined fusion plasmas due to their contribution to plasma heating and potentially current drive [1, 2]. A key aspect of EP confinement is related to the shear Alfvén wave (SAW) instabilities [3] resonantly excited by EPs [4–8]. In magnetic confinement devices, SAW can be excited as various EP continuum modes (EPMs) [7] or discrete Alfvén modes (AEs) [9–11] inside the frequency gaps of the SAW continuum induced by equilibrium magnetic geometry and plasma nonuniformity. These SAW instabilities can then induce significant EPs anomalous transport loss across the magnetic surfaces, leading to plasma performance degradation and even damage of plasma facing components [12, 13]. With the EPs transport rate determined by the saturation amplitude and spectrum of SAW instabilities [14, 15], it is crucial to understand the nonlinear dynamics resulting in their saturation. In the past decades, nonlinear saturation of SAW instabilities has been broadly investigated both numerically and theoretically [16–33], among which, one of the most important channel is nonlinear wave-wave coupling [34, 35], i.e. nonlinear spectrum evolution of SAW instabilities due to interacting with other collective electromagnetic oscillations.
In the advanced scenarios of future reactor burning plasmas, a large fraction of non-inductive (e.g. bootstrap) current will be maintained [36] off-axis, and the magnetic shear is reversed in the plasma core region, where large fraction of energetic fusion alpha particles are generated [37]. As a result, a specific Alfvén eigenmode, namely the reversed shear Alfvén eigenmode (RSAE, also known as Alfvén cascade due to its frequency sweeping character [38–41]) could be excited and play important roles in transport of fusion alpha particles. In particular, as multiple-n RSAEs can be strongly driven unstable simultaneously in reactors with machine size being much larger than fusion alpha particle characteristic orbit width, RSAEs can lead to strong alpha particle re-distribution and transport [37, 42]. RSAE is a branch of Alfvén eigenmodes localized around the SAW continuum extremum induced by the local minimum of the safety factor q-profile (labeled as qmin) to minimize the continuum damping, and is characterized by a radial width of ∼q/(r20q″ [41], with r_0 being the radial location of q_{\mathrm{min}} and q^{''}\equiv \partial^2_r q . RSAE was originally observed in the advanced operation experiments in JT-60U tokamak [43], and was then detected in numerous JET discharges [44]. In present day tokamaks, RSAEs are generally excited by large orbit EP during current ramp up stage where reversed shear q -profile is created by insufficient current penetration [45]. In most cases, with q_{\mathrm{min}} decreasing from, e.g., a rational value m/n to (m-1/2)/n , the RSAEs exhibit upward frequency sweeping from beta-induced Alfvén eigenmode (BAE) [46] to the toroidal Alfveń eigenmode (TAE) [9, 11] frequency ranges. Here, m and n stand for the poloidal and toroidal mode numbers, respectively.
Due to the increasing importance in reactor burning plasmas operating at advanced scenarios, RSAE has drawn much attention in recent investigations. For instance, the resonant decay of RSAE into a generic low frequency Alfvén mode (LFAM) was investigated in reference [47], based on which, a potential alpha channelling mechanism [48] via the LFAM Landau damping was also proposed and analyzed. The modulational instability of a finite amplitude RSAE and excitation of the zero-frequency zonal structures were investigated in reference [49], where RSAE was saturated due to the modulation of SAW continuum and scattered into short-radial-wavelength stable domain. In particular, it is pointed out that, the generation of zonal current around the q_{\mathrm{min}} region can be of particular importance, due to the sensitive dependence of RSAE on reversed shear profiles. Further numerical investigations of RSAE nonlinear dynamics can also be found in, e.g., references [50, 51], where RSAE nonlinear saturation due to wave-particle radial decoupling and zonal flow generation were investigated, respectively. It is also noteworthy that, in reference [52], a nonlinear saturation channel of RSAEs via nonlinear harmonic generation was investigated, where quasi-modes with double and/or triple toroidal mode numbers of the primary linearly unstable RSAE were generated due to kinetic electron contribution via “magnetic fluttering”, and led to RSAE nonlinear saturation. The setting of the simulation seems though, to some extend “artificial”, as only few toroidal mode numbers are kept in the simulation, it provides the important information of RSAE dissipation via nonlinear harmonic generation.
Motivated by reference [52], in this work, we present a potential nonlinear saturation mechanism for RSAE via nonlinear quasi-mode generation. This nonlinear mode coupling is achieved through the non-adiabatic responses to electrons, corresponding to the magnetic fluttering nonlinearity as addressed in reference [52]. Meanwhile, the mode coupling is generalized from RSAE self-coupling in the simulation [52] to include also the interaction between two RSAEs with different toroidal mode numbers, for which the nonlinear coupling could be much stronger. Generally, this quasi-mode could experience significant continuum or radiative damping, and provide a channel for primary RSAE energy dissipation. Using nonlinear gyrokinetic theory, the parametric dispersion relation of this nonlinear mode coupling process is derived. Focusing on the continuum damping of the quasi-mode, the resultant nonlinear damping to RSAE is then analyzed and estimated.
The remainder of this paper is arranged as follows. In section 2, the theoretical model is given. In section 3, the nonlinear dispersion relation describing the RSAE nonlinear evolution due to interaction with another background RSAE is derived. It is then used in section 4 to investigate the continuum damping of the quasi-mode. An estimation of the resulting damping to RSAE is also presented.
Considering two co-propagating RSAEs \Omega_0\equiv\Omega_0(\omega_0,{\boldsymbol{k}}_0) and \Omega_1\equiv\Omega_1(\omega_1,{\boldsymbol{k}}_1) coupling and generating a beat wave \Omega_\mathrm{b}\equiv\Omega_\mathrm{b}(\omega_\mathrm{b},{\boldsymbol{k}}_\mathrm{b}) , with the frequency and wavenumber of \Omega_\mathrm{b} determined by the matching condition \Omega_\mathrm{b} = \Omega_0 +\Omega_1 , the beat wave \Omega_\mathrm{b} is likely a high-frequency quasi-mode bearing significant continuum or radiative damping, as it may not satisfy the global RSAE dispersion relation with corresponding toroidal mode number n_\mathrm{b} = n_0+n_1 . Here, for “co-propagating”, we mean the two RSAEs propagate in the same direction along the magnetic field line. This nonlinear coupling provides the primary RSAEs an indirect damping mechanism, and may result in their saturation. Here, for simplicity of the discussion while focusing on the main physics picture, we focus on the continuum damping of the quasi-mode, and investigate the resultant nonlinear saturation of primary RSAEs. A sketched illustration of the proposed process is given in figure 1, where two RSAEs with n=3,\ 4 couple and generate an n = 7 high-frequency quasi-mode, which can be heavily damped due to coupling with the corresponding shear Alfvén continuum.
The nonlinear coupling of the two RSAEs and the resultant damping are investigated in a uniform low- \beta magnetized plasma using the standard nonlinear perturbation theory, with \beta\ll1 being the ratio between plasma and magnetic pressures. Here, for “uniform”, we mean the effects associated with diamagnetic effects are systematically neglected, while noting magnetically confined plasma is intrinsically nonuniform. Introducing the scalar potential \delta\phi and parallel component of vector potential \delta A_\parallel as the perturbed field variables, one then has \delta\phi = \delta\phi_0+\delta\phi_1+\delta\phi_\mathrm{b} with the subscripts 0 , 1 and \mathrm{b} denoting \Omega_0 , \Omega_1 and \Omega_\mathrm{b} , respectively. For convenience of investigation, \delta A_\parallel is replaced by \delta\psi\equiv \omega\delta A_\parallel/(ck_\parallel) , such that \delta\phi = \delta\psi can straightforwardly recover the ideal MHD limit, i.e. vanishing parallel electric field fluctuation \delta E_\parallel . Both electrons and ions are chacterised by Maxwellian equilibrium distributions F_\mathrm{M} .
For RSAEs typically dominated by single- n and single- m mode structures near q_{\mathrm{min}} , we take
\delta\phi_k = A_k(t) \Phi_k(x) \exp(-\mathrm{i}\omega_kt+\mathrm{i}n\xi-\mathrm{i}m\theta), |
with A_k(t) being the slowly varying mode amplitude, \Phi_k(x) being the parallel mode structure localized around q_{\mathrm{min}} with x\equiv nq-m , and the normalization \int |\Phi_k|^2\mathrm{d}x = 1 can be adopted.
Nonlinear mode equations can be derived from charge quasi-neutrality condition
\frac{N_0e^2}{T_\mathrm{i}}\left(1+\frac{T_\mathrm{i}}{T_\mathrm{e}}\right)\delta\phi_k = \sum\limits_{j = \mathrm{e},\mathrm{i}}\langle qJ_k\delta H_k\rangle_j, | (1) |
and nonlinear gyrokinetic vorticity equation [53]
\begin{split}& \frac{c^2}{4 {\text{π}}\omega_k^2}B\frac{\partial}{\partial l}\frac{k^2_\bot}{B}\frac{\partial}{\partial l}\delta\psi_k+\frac{e^2}{T_\mathrm{i}} \langle(1-J_k^2)F_\mathrm{M}\rangle\delta\phi_k -\sum\limits_{j = \mathrm{e},\mathrm{i}}\left \langle qJ_k\frac{\omega_\mathrm{d}}{\omega_k}\delta H_k\right \rangle_j \\ &= -\frac{\mathrm{i}}{\omega_k}\sum\limits_{{\boldsymbol{k}} = {\boldsymbol{k}}'+{\boldsymbol{k}}''}\Lambda_{k''}^{k'}{\Bigg [}\langle e(J_k J_{k'}-J_{k''})\delta L_{k'}\delta H_{k''}\rangle \\& \;\;\;\;\;\; +\frac{c^2}{4 {\text{π}}}k^{''2}_\bot\frac{\partial_l\delta\psi_{k'}\partial_l\delta\psi_{k''}}{\omega_{k'}\omega_{k''}}{\Bigg ]}. \end{split} | (2) |
Here, the terms on the left hand side of equation (2) are field line bending, inertia and curvature-pressure coupling terms, respectively, whereas the terms on the right hand side represent Reynolds and Maxwell stresses dominating in short wavelength limit. N_0 is the equilibrium particle density, q_j is the electric charge, the angular brackets \langle \dots\rangle denote velocity space integration, \partial_l is the spatial derivative along the equilibrium magnetic field, k_\bot = \sqrt{k^2_r+k^2_{\theta}} is the perpendicular wavenumber, J_k\equiv J_0(k_\bot \rho) with J_0 being the Bessel function of zero index accounting for finite Larmor radius effects, and \rho = v_\bot/\Omega_\mathrm{c} is the Larmor radius with \Omega_\mathrm{c} being the cyclotron frequency. Furthermore, \omega_\mathrm{d} = (v^2_\bot+2v^2_\parallel)/ (2\Omega_\mathrm{c}R_0)(k_r \sin\theta+k_\theta\cos\theta) is the magnetic drift frequency, \Lambda_{k''}^{k'} = (c/B_0){\hat{\boldsymbol{b}}}\cdot{\boldsymbol{ k}}''\times {\boldsymbol{k}}' accounts for perpendicular coupling with the constraint of frequency and wavevector matching conditions, and \delta L_k\equiv\delta \phi_k-k_\parallel v_\parallel\delta\psi_k/\omega_k is the scalar potential in the frame moving with guiding center. The non-adiabatic particle response \delta H_k is derived from the nonlinear gyrokinetic equation [54]:
\begin{split} & (-\mathrm{i}\omega+v_{\parallel}\partial_l+\mathrm{i}\omega_{\mathrm{d}})\delta H_k \\ &=-\mathrm{i}\frac{q}{T_j}\omega_kF_{\mathrm{M}}J_k\delta Lk-\sum\limits_{\boldsymbol{k}=\boldsymbol{k}'+\boldsymbol{k}''}^{ }\Lambda_{k''}^{k'}J_{k'}\delta L_{k'}\delta H_{k''}. \end{split} | (3) |
In this section, the coupled nonlinear equations for RSAE and the high-frequency quasi-mode are derived in sections 3.1 and 3.2, which are then combined and give the parametric dispersion relation in section 3.3.
The nonlinear equation for quasi-mode \Omega_\mathrm{b} generation can be derived from the quasi-neutrality condition and the nonlinear gyrokinetic vorticity equation. The nonlinear non-adiabatic particle response of \Omega_\mathrm{b} can be derived from the nonlinear component of equation (3) noting the k_\parallel v_\mathrm{e}\gg\omega\gg k_\parallel v_\mathrm{i}\gtrsim\omega_\mathrm{d} ordering, and one obtains
\delta H_{\mathrm{bi}}^{\mathrm{NL}} = 0, | (4) |
\delta H_{\mathrm{be}}^{\mathrm{NL}} = \mathrm{i}\Lambda_0^1 \frac{e}{T_\mathrm{e}}F_\mathrm{M} \frac{1}{k_{\parallel \mathrm{b}}}\left(\frac{k_{\parallel 1}}{\omega_1}-\frac{k_{\parallel 0}}{\omega_0}\right)\delta\psi_0\delta\psi_1. | (5) |
In deriving equations (4) and (5), the linear particle responses \delta H_{k\mathrm{i}}^\mathrm{L} = (e/T_\mathrm{i})F_{\mathrm{M}}J_k\delta\phi_k and \delta H_{k\mathrm{e}}^\mathrm{L} = -(e/T_\mathrm{e})F_{\mathrm{M}}\delta\psi_k are used (interested readers may refer to the Appendix for the detailed derivation). With \delta H_{\mathrm{be}}^{\mathrm{NL}} representing the coupling between \Omega_0 and \Omega_1 due to nonlinear electron contribution, it corresponds to the magnetic fluttering nonlinearity investigated in [52]. Substituting equations (4) and (5) into the quasi-neutrality condition, one obtains
\delta\psi_{\mathrm{b}}=\delta\phi_{\mathrm{b}}+\mathrm{i}\Lambda_0^1\frac{1}{k_{\parallel\mathrm{b}}}\left(\frac{k_{\parallel1}}{\omega_1}-\frac{k_{\parallel0}}{\omega_0}\right)\delta\psi_0\delta\psi_1, | (6) |
i.e. breaking of ideal MHD constraint due to nonlinear mode coupling, while finite parallel electric field associated with linear FLR effects is not included here for simplicity [55]. This is also consistent with the b_k\ll1 ordering for linear unstable RSAEs with typically k^{-1}_{\perp} comparable to EP characteristic orbit width. Substituting the particle responses into the nonlinear gyrokinetic vorticity equation, one obtains
\begin{split} & b_{\mathrm{b}}\left(\delta\phi_{\mathrm{b}}-\frac{k_{\parallel\mathrm{b}}^2v_{\mathrm{A}}^2}{\omega_{\mathrm{b}}^2}\delta\psi_{\mathrm{b}}-\frac{\omega_{\mathrm{G}}^2}{\omega_{\mathrm{b}}^2}\delta\phi_{\mathrm{b}}\right) \\ &\qquad=-\mathrm{i}\frac{\Lambda_0^1}{\omega_{\mathrm{b}}}(b_0-b_1)\left(1-\frac{k_{\parallel0}k_{\parallel1}v_{\mathrm{A}}^2}{\omega_0\omega_1}\right)\delta\phi_0\delta\phi_1, \end{split} | (7) |
with b_k = k_\bot^2\rho_\mathrm{i}^2/2 , v_\mathrm{A} being the Alfvén speed, \omega_\mathrm{G}\equiv \sqrt{7/4+\tau}v_\mathrm{i}/R_0 being the leading order geodesic acoustic mode frequency [56, 57] and \tau\equiv T_\mathrm{e}/T_\mathrm{i} . Combining equations (6) and (7), one obtains
b_\mathrm{b} \varepsilon_{\mathrm{Ab}}\delta\phi_\mathrm{b} = \mathrm{i}\frac{\Lambda_0^1}{\omega_\mathrm{b}}\beta_\mathrm{b}\delta\phi_0\delta\phi_1. | (8) |
Equation (8) is the desired nonlinear equation describing the high-frequency quasi-mode \Omega_\mathrm{b} generation due to \Omega_0 and \Omega_1 coupling, with the \Omega_\mathrm{b} dielectric function \varepsilon_{\mathrm{Ab}} defined as
\varepsilon_{\mathrm{Ab}}\equiv 1-\frac{k_{\parallel \mathrm{b}}^2v_\mathrm{A}^2}{\omega_\mathrm{b}^2}-\frac{\omega_\mathrm{G}^2}{\omega_\mathrm{b}^2}, |
and the nonlinear coupling coefficient \beta_\mathrm{b} given by
\beta_\mathrm{b} =b_\mathrm{b} \frac{k_{\parallel \mathrm{b}}v_\mathrm{A}}{\omega_\mathrm{b}}\left(\frac{k_{\parallel 1}v_\mathrm{A}}{\omega_1}-\frac{k_{\parallel 0}v_\mathrm{A}}{\omega_0}\right) \nonumber -(b_0-b_1)\left(1-\frac{k_{\parallel 0}k_{\parallel 1}v^2_\mathrm{A}}{\omega_0\omega_1}\right). |
It is worth mentioning that, the \Omega_\mathrm{b} dielectric function, \varepsilon_{\mathrm{Ab}} , may not satisfy the global linear RSAE dispersion relation for toroidal mode number n_\mathrm{b} , and \Omega_\mathrm{b} could be a quasi-mode experiencing heavy damping, leading to the dissipation of both itself and the primary RSAEs, as shown later.
The nonlinear coupling equation for the test RSAE \Omega_0 can be derived following a similar procedure. However, noting that \Omega_\mathrm{b} is a quasi-mode, one needs to keep both the linear and nonlinear particle responses since they could be of the same order. The resultant nonlinear non-adiabatic particle responses of \Omega_0 are respectively
\delta H_{0\mathrm{i}}^{\mathrm{NL}} = 0, | (9) |
\begin{split} \delta H_{0\mathrm{e}}^{\mathrm{NL}} =\;& -\mathrm{i}\Lambda_0^1 \frac{e}{T_\mathrm{e}}F_\mathrm{M} \frac{1}{k_{\parallel 0}}\left(\frac{k_{\parallel 1}}{\omega_1}-\frac{k_{\parallel \mathrm{b}}}{\omega_\mathrm{b}}\right)\delta\psi_1^{*}\delta\psi_\mathrm{b} \\& -(\Lambda_0^1)^2\frac{e}{T_\mathrm{e}}F_\mathrm{M} \frac{k_{\parallel 1}}{k_{\parallel 0}k_{\parallel \mathrm{b}}\omega_1}\left(\frac{k_{\parallel 1}}{\omega_1}-\frac{k_{\parallel 0}}{\omega_0}\right)|\delta\psi_1|^2\delta\psi_0. \end{split} | (10) |
Substituting equations (9) and (10) into the quasi-neutrality condition, one obtains
\begin{split} \delta\psi_0 =\;& \delta\phi_0-\mathrm{i}\Lambda_0^1 \frac{1}{k_{\parallel 0}}\left(\frac{k_{\parallel 1}}{\omega_1}-\frac{k_{\parallel \mathrm{b}}}{\omega_\mathrm{b}}\right)\delta\psi_1^{*}\delta\psi_\mathrm{b} \\ &- (\Lambda_0^1)^2\frac{k_{\parallel 1}}{k_{\parallel 0}k_{\parallel \mathrm{b}}\omega_1}\left(\frac{k_{\parallel 1}}{\omega_1}-\frac{k_{\parallel 0}}{\omega_0}\right)|\delta\psi_1|^2\delta\psi_0. \end{split} | (11) |
On the other hand, the nonlinear gyrokinetic vorticity equation yields
\begin{split} & b_0\left(\delta\phi_0-\frac{k_{\parallel0}^2v_{\mathrm{A}}^2}{\omega_0^2}\delta\psi_0-\frac{\omega_{\mathrm{G}}^2}{\omega_0^2}\delta\phi_0\right) \\ &\qquad=\mathrm{i}\frac{\Lambda_0^1}{\omega_0}(b_{\mathrm{b}}-b_1)\left(1-\frac{k_{\parallel\mathrm{b}}k_{\parallel1}v_{\mathrm{A}}^2}{\omega_{\mathrm{b}}\omega_1}\right)\delta\phi_1^*\delta\phi_{\mathrm{b}}. \end{split} | (12) |
Combining equations (11) and (12), one obtains the nonlinear equation of \Omega_0
b_0 (\varepsilon_{\mathrm{A}0}+\varepsilon_{\mathrm{A}0}^{\mathrm{NL}})\delta\phi_0 = -\mathrm{i}\frac{\Lambda_0^1}{\omega_0}\beta_0\delta\phi_1^{*}\delta\phi_\mathrm{b}, | (13) |
with the linear \Omega_0 dielectric function in the WKB limit given by
\varepsilon_{\mathrm{A}0}\equiv 1-\frac{k_{\parallel 0}^2v_\mathrm{A}^2}{\omega_0^2}-\frac{\omega_\mathrm{G}^2}{\omega_0^2}, |
the nonlinear coupling coefficient \beta_0 given by
\beta_0 = b_0 \frac{k_{\parallel 0}v_\mathrm{A}}{\omega_0}\left(\frac{k_{\parallel 1}v_\mathrm{A}}{\omega_1}-\frac{k_{\parallel \mathrm{b}}v_\mathrm{A}}{\omega_\mathrm{b}}\right) -(b_\mathrm{b}-b_1)\left(1-\frac{k_{\parallel \mathrm{b}}k_{\parallel 1}v^2_\mathrm{A}}{\omega_\mathrm{b}\omega_1}\right), | (14) |
and \varepsilon_{\mathrm{A}0}^{\mathrm{NL}} due to nonlinear particle contribution to \Omega_\mathrm{b} being
\varepsilon_{\mathrm{A}0}^{\mathrm{NL}} = \frac{(\Lambda_0^1)^2}{\omega_0\omega_\mathrm{b}}\frac{k_{\parallel0}k_{\parallel1}v^2_\mathrm{A}}{\omega_0\omega_1}\frac{\omega_\mathrm{b}}{k_{\parallel \mathrm{b}}}\left(\frac{k_{\parallel 1}}{\omega_1}-\frac{k_{\parallel 0}}{\omega_0}\right)|\delta\phi_1|^2. |
Equation (13) is the nonlinear equation for the test RSAE \Omega_0 evolution due to the feedback of the quasi-mode \Omega_\mathrm{b} , and can be coupled with equation (8) to yield the nonlinear dispersion relation for \Omega_0 regulation via the high-frequency quasi-mode generation.
Combining equations (8) and (13), one obtains
b_0 b_\mathrm{b} (\varepsilon_{\mathrm{A}0}+\varepsilon_{\mathrm{A}0}^{\mathrm{NL}})\varepsilon_{\mathrm{Ab}}\delta \phi_0 = \frac{(\Lambda_0^1)^2}{\omega_0 \omega_\mathrm{b}}\beta_0 \beta_\mathrm{b}|\delta\phi_1|^2\delta\phi_0. | (15) |
Equation (15) describes the evolution of the test RSAE \Omega_0 due to the nonlinear interaction with another RSAE \Omega_1 , which can also be considered as the “parametric decay dispersion relation” of \Omega_1 decaying into \Omega_0 and \Omega_\mathrm{b} . Noting that \varepsilon_{\mathrm{A}0}^{\mathrm{NL}} related term contributes only to the nonlinear frequency shift, re-organizing equation (15) and taking the imaginary part, one obtains
\frac{2\gamma_{\mathrm{ND}}}{\omega_{0\mathrm{r}}}b_0 \delta\phi_0 = {\text{π}}\frac{(\Lambda_0^1)^2}{\omega_0 \omega_\mathrm{b}}\frac{\beta_0 \beta_\mathrm{b}}{b_\mathrm{b}}\delta(\varepsilon_{\mathrm{Ab}})|\delta\phi_1|^2\delta\phi_0. | (16) |
In deriving equation (16), we have expanded \varepsilon_{\mathrm{A}0}\simeq \mathrm{i}\partial_{\omega_{0\mathrm{r}}}\varepsilon_{\mathrm{A0r}}(\partial_t-\gamma_{0})\simeq -(2\mathrm{i}/\omega_{0\mathrm{r}})\gamma_{\mathrm{ND}} , with \gamma_0 being the linear growth rate of \Omega_0 and \gamma_{\mathrm{ND}} being its damping rate due to scattering by \Omega_\mathrm{b} , respectively. It is also noteworthy that, as \Omega_\mathrm{b} is a quasi-mode with the imaginary part of \varepsilon_{\mathrm{Ab}} being comparable to the real part, no expansion is made to \varepsilon_{\mathrm{Ab}} . Meanwhile, for the continuum damping of interest, \mathrm{Im} (1/\varepsilon_{\mathrm{Ab}}) = - {\text{π}}\delta(\varepsilon_{\mathrm{Ab}}) is taken, corresponding to the absorption of the nonlinear generated quasi-mode \Omega_\mathrm{b} near the SAW continuum resonance layer [55].
An estimation of the nonlinear damping rate is made to quantify the contribution of this nonlinear process. Multiplying equation (16) with \Phi_0^{*} and averaging over radial mode structure, one obtains
\frac{2\gamma_{\mathrm{ND}}}{\omega_{0\mathrm{r}}}\langle \Phi_0^{*}b_0\Phi_0\rangle_x = {\text{π}}\left\langle \frac{(\Lambda_0^1)^2}{\omega_0 \omega_\mathrm{b}}\frac{\beta_0 \beta_\mathrm{b}}{b_\mathrm{b}}\delta(\varepsilon_{\mathrm{Ab}})|A_1|^2|\Phi_1|^2|\Phi_0|^2 \right\rangle _x. | (17) |
Here, \langle \dots\rangle_x\equiv \int\cdots \mathrm{d}x denotes the integration over x , with the weighting of |\Phi_0|^2 . To make analytical progress, the parallel mode structures for RSAEs are taken as \Phi_k\simeq \exp(-x^2/2\Delta_k^2)/( {\text{π}}^{1/4}\Delta_k^{1/2}) with \Delta_k being the characteristic radial width of the parallel mode structures and one typically has \Delta_0\sim \Delta_1\lesssim{\cal{O}}(1) .
Equation (17) gives the test RSAE \Omega_0 damping rate due to coupling to a “background” RSAE \Omega_1 . As multiple- n RSAEs could be driven unstable simultaneously [37] at the same location, all the background RSAEs interacting with \Omega_0 should be taken into account. Summation over all the RSAEs within strong or moderate coupling range to the test RSAE \Omega_0 , and assuming that the integrated electromagnetic fluctuation amplitude induced by RSAEs is of the same order as the background RSAEs, the nonlinear damping rate can be estimated as
\frac{\gamma_{\mathrm{ND}}}{\omega_{0\mathrm{r}}}\sim \frac{b}{\Delta_0\Delta_1} \left(\frac{qR_0}{\rho_\mathrm{i}}\right)^2 \left|\frac{\delta B_r}{B_0}\right|^2 \frac{x_0^3}{\varpi^6} \sim {\cal{O}}(10^{-3}-10^{-2}). | (18) |
In estimating \gamma_{\mathrm{ND}} , \delta(\varepsilon_{\mathrm{Ab}}) = \partial\varepsilon_{\mathrm{Ab}}/\partial x\sum\nolimits_{x_0}^{}\delta(x-x_0)\simeq -2x/\varpi_\mathrm{b}^2\sum\nolimits_{x_0}^{}\delta(x-x_0) is taken, with \varpi\equiv \omega/\omega_\mathrm{A} , \omega_\mathrm{A}\equiv v_\mathrm{A}/(q_{\mathrm{min}}R_0) being the local Alfvén frequency and x_0 being the zero points of \varepsilon_{\mathrm{Ab}} . Other parameters are taken as T_\mathrm{i}/T_\mathrm{E}\sim {\cal{O}}(10^{-2}) , R_0/\rho_\mathrm{i}\sim {\cal{O}}(10^3) , |\delta B_{\mathrm{\mathit{r}}}/B_0|^2\sim{\cal{O}}(10^{-7}) , b\sim k^2_\theta\rho_\mathrm{i}^2\sim(T_\mathrm{i}/T_\mathrm{E})/q^2 for linearly unstable RSAEs with T_\mathrm{E} being the EP characteristic energy, \delta Br being the perturbed radial magnetic field, and B0 being the equilibrium magnetic field amplitude.
Equation (18) shows an appreciable nonlinear damping to the test RSAE \Omega_0 , which could make significant contribution to its nonlinear saturation. Note that in the present work, only the scattering to high-frequency quasi-mode is taken into account. Nevertheless, one can generalize the analysis to include other damping channels including other nonlinear mode coupling mechanism [47, 49, 58]. This is, however, beyond the scope of the present work, focusing on providing an interpretation to the simulation of reference [52], with the generalization to include coupling to background RSAEs with different toroidal mode numbers.
Motivated by recent simulation study [52], a novel mechanism for RSAE nonlinear saturation is proposed and analyzed, which is achieved through generation of a high-frequency quasi-mode by the nonlinear mode coupling of two RSAEs. This high-frequency quasi-mode can be significantly damped due to coupling to the corresponding SAW continuum, thus leads to a nonlinear damping effect to the RSAEs, and promotes their nonlinear saturation. The nonlinear dispersion relation describing this nonlinear coupling process is derived based on the nonlinear gyrokinetic theory. To estimate the relevance of this nonlinear saturation mechanism to RSAE, an estimation of the nonlinear damping rate to the test RSAE is given by \gamma_{\mathrm{ND}}/\omega_{0\mathrm{r}}\sim {\cal{O}}(10^{-3}-10^{-2}) under typical parameters of the future burning plasmas. This result could be comparable with the typical RSAE linear growth rate excited by resonant EPs, and thus, demonstrate the significance of the nonlinear saturation mechanism proposed here.
The nonlinear coupling coefficient derived in this work is complicated, and depends on various conditions including the frequency, wavenumber, radial mode structure of the RSAEs and the structure of Alfvén continuum corresponding to the mode number of quasi-mode. This study seeks to estimate the relevance of the nonlinear saturation mechanism and has not done a thorough investigation on the optimised parameter regimes for this process to occur and dominate. For more detailed analysis, interested readers may refer to [47] with the nonlinear coupling coefficient having similar features.
As a final remark, the high-frequency quasi-mode discussed here, is damped due to the coupling to local Alfvén continuum only, whereas other damping effects (e.g. radiative damping and Landau damping due to frequency mismatch) are not included, and inclusion of which could yield an enhanced regulation effect to RSAE. Besides, the final nonlinear saturation of RSAE may require other channels including the self-consistent evolution of EPs distribution function [59], spontaneous decay into LFAM [47], zonal field generation [49] and geodesic acoustic mode (GAM) generation [58]. Further comprehensive and detailed investigations, particularly through large scale nonlinear gyrokinetic simulations, are required to assess the saturation level of RSAE and the energetic particle transport rate.
In this appendix, the nonlinear particle responses to \Omega_\mathrm{b} are derived, based on the nonlinear components of equation (3). Specifically, for ion, noting \omega\gg k_\parallel v_\mathrm{i}\gtrsim\omega_\mathrm{d} , one has
\begin{split} -\mathrm{i}\omega\delta H_{\mathrm{bi}}^{\mathrm{NL}} \simeq &-\Lambda_0^1\left( J_1\delta L_1 \delta H_{0\mathrm{i}}^\mathrm{L}- J_0\delta L_0\delta H_{1\mathrm{i}}^\mathrm{L} \right) \\ \simeq& -\Lambda_0^1\left( J_1\delta\phi_1 \delta H_{0\mathrm{i}}^\mathrm{L}- J_0\delta\phi_0\delta H_{1\mathrm{i}}^\mathrm{L} \right)\\ = &\,\, 0. \end{split} | (A.1) |
Note in this derivation, \delta L\equiv \delta\phi-(k_\parallel v_\parallel/\omega)\delta\psi\simeq \delta\phi for ion with \omega\gg k_\parallel v_\mathrm{i} , and \delta H_{k\mathrm{i}}^\mathrm{L} = (e/T_\mathrm{i})F_\mathrm{M}J_k\delta\phi_k is used.
For electron with k_\parallel v_\mathrm{e}\gg\omega \gg\omega_\mathrm{d} and negligible FLR effects, only the k_\parallel v_\parallel term is kept, and the nonlinear gyrokinetic equation for electron can be written as
\begin{split} \mathrm{i}k_{\parallel \mathrm{b}} v_\parallel\delta H_{\mathrm{be}}^{\mathrm{NL}}\simeq&-\Lambda_0^1\left( \delta L_1 \delta H_{0\mathrm{e}}^\mathrm{L}- \delta L_0\delta H_{1\mathrm{e}}^\mathrm{L} \right) \\ \simeq&-\Lambda_0^1\frac{e}{T_\mathrm{e}}F_\mathrm{M}\left( \frac{k_{\parallel1} v_\parallel}{\omega_1}-\frac{k_{\parallel0} v_\parallel}{\omega_0} \right)\delta\psi_0\delta\psi_1. \end{split} | (A.2) |
One obtains, the nonlinear electron response to \Omega_\mathrm{b}
\delta H_{\mathrm{be}}^{\mathrm{NL}} = \mathrm{i}\Lambda_0^1\frac{e}{T_\mathrm{e}}F_\mathrm{M}\frac{1}{k_{\parallel \mathrm{b}}}\left(\frac{k_{\parallel1}}{\omega_1}-\frac{k_{\parallel0}}{\omega_0}\right)\delta\psi_0\delta\psi_1. | (A.3) |
Note in this derivation, \delta L\equiv \delta\phi-(k_\parallel v_\parallel/\omega)\delta\psi\simeq -(k_\parallel v_\parallel/\omega)\delta\psi for electron with \omega\ll k_\parallel v_\mathrm{e} , and \delta H_{k\mathrm{e}}^\mathrm{L} = -(e/T_\mathrm{e})F_\mathrm{M}\delta\psi_k is used. The nonlinear non-adiabatic responses to \Omega_0 can be derived similarly.
This work was supported by the Strategic Priority Research Program of Chinese Academy of Sciences (No. XDB0790000), the Collaborative Innovation Program of Hefei Science Center, CAS (No. 2022HSC-CIP008), and National Natural Science Foundation of China (Nos. 12275236 and 12261131622).
[1] |
Fasoli A et al 2007 Nucl. Fusion 47 S264 doi: 10.1088/0029-5515/47/6/S05
|
[2] |
Chen L and Zonca F 2016 Rev. Mod. Phys. 88 015008 doi: 10.1103/RevModPhys.88.015008
|
[3] |
Alfvén H 1942 Nature 150 405
|
[4] |
Kolesnichenko Y I 1967 At. Energ. 23 289
|
[5] |
Mikhailovskii A B 1975 Zh. Eksp. Teor. Fiz. 68 25
|
[6] |
Rosenbluth M N and Rutherford P H 1975 Phys. Rev. Lett. 34 1428 doi: 10.1103/PhysRevLett.34.1428
|
[7] |
Chen L 1994 Phys. Plasmas 1 1519 doi: 10.1063/1.870702
|
[8] |
Chen L and Zonca F 2007 Nucl. Fusion 47 886 doi: 10.1088/0029-5515/47/8/021
|
[9] |
Cheng C Z, Chen L and Chance M S 1985 Ann. Phys. 161 21 doi: 10.1016/0003-4916(85)90335-5
|
[10] |
Chen L 1989 On resonant excitations of high-n magnetohydrodynamic modes by energetic/alpha particles in tokamaks Princeton: Princeton Plasma Physics Lab.
|
[11] |
Fu G Y and Van Dam J W 1989 Phys. Fluids B 1 1949
|
[12] |
ITER Physics Expert Group on Energetic Particles, Heating and Current Drive and ITER Physics Basis Editors 1999 Nucl. Fusion 39 2471 doi: 10.1088/0029-5515/39/12/305
|
[13] |
Ding R et al 2015 Nucl. Fusion 55 023013 doi: 10.1088/0029-5515/55/2/023013
|
[14] |
Chen L 1999 J. Geophys. Res.: Space Phys. 104 2421 doi: 10.1029/1998JA900051
|
[15] |
Falessi M V and Zonca F 2019 Phys. Plasmas 26 022305 doi: 10.1063/1.5063874
|
[16] |
Berk H L and Breizman B N 1990 Phys. Fluids B 2 2246 doi: 10.1063/1.859406
|
[17] |
Todo Y et al 1995 Phys. Plasmas 2 2711 doi: 10.1063/1.871235
|
[18] |
Lang J Y, Fu G Y and Chen Y 2010 Phys. Plasmas 17 042309 doi: 10.1063/1.3394702
|
[19] |
Briguglio S et al 2014 Phys. Plasmas 21 112301 doi: 10.1063/1.4901028
|
[20] |
Zhu J, Fu G Y and Ma Z W 2013 Phys. Plasmas 20 072508 doi: 10.1063/1.4816950
|
[21] |
Zhang H W et al 2022 Nucl. Fusion 62 026047 doi: 10.1088/1741-4326/ac40c3
|
[22] |
Spong D A, Carreras B A and Hedrick C L 1994 Phys. Plasmas 1 1503 doi: 10.1063/1.870700
|
[23] |
Hahm T S and Chen L 1995 Phys. Rev. Lett. 74 266 doi: 10.1103/PhysRevLett.74.266
|
[24] |
Zonca F et al 1995 Phys. Rev. Lett. 74 698 doi: 10.1103/PhysRevLett.74.698
|
[25] |
Chen L et al 1998 Plasma Phys. Control. Fusion 40 1823 doi: 10.1088/0741-3335/40/11/001
|
[26] |
Todo Y, Berk H L and Breizman B N 2010 Nucl. Fusion 50 084016 doi: 10.1088/0029-5515/50/8/084016
|
[27] |
Chen L and Zonca F 2012 Phys. Rev. Lett. 109 145002 doi: 10.1103/PhysRevLett.109.145002
|
[28] |
Qiu Z Y, Chen L and Zonca F 2016 Phys. Plasmas 23 090702 doi: 10.1063/1.4962997
|
[29] |
Qiu Z Y, Chen L and Zonca F 2017 Nucl. Fusion 57 056017 doi: 10.1088/1741-4326/aa6413
|
[30] |
Qiu Z Y et al 2018 Phys. Rev. Lett. 120 135001 doi: 10.1103/PhysRevLett.120.135001
|
[31] |
Qiu Z Y, Chen L and Zonca F 2019 Nucl. Fusion 59 066024 doi: 10.1088/1741-4326/ab1693
|
[32] |
Chen L, Qiu Z Y and Zonca F 2022 Nucl. Fusion 62 094001 doi: 10.1088/1741-4326/ac7cf9
|
[33] |
Chen L, Qiu Z Y and Zonca F 2023 Nucl. Fusion 63 106016 doi: 10.1088/1741-4326/acf230
|
[34] |
Sagdeev R Z and Galeev A A 1969 Nonlinear Plasma Theory (New York: W. A. Benjamin
|
[35] |
Qiu Z Y, Chen L and Zonca F 2023 Rev. Mod. Plasma Phys. 7 28 doi: 10.1007/s41614-023-00130-7
|
[36] |
Gormezano C et al 2007 Nucl. Fusion 47 S285 doi: 10.1088/0029-5515/47/6/S06
|
[37] |
Wang T et al 2018 Phys. Plasmas 25 062509 doi: 10.1063/1.5026652
|
[38] |
Shinohara K et al 2001 Nucl. Fusion 41 603 doi: 10.1088/0029-5515/41/5/313
|
[39] |
Sharapov S E et al 2006 Nucl. Fusion 46 S868 doi: 10.1088/0029-5515/46/10/S02
|
[40] |
Berk H L et al 2001 Phys. Rev. Lett. 87 185002 doi: 10.1103/PhysRevLett.87.185002
|
[41] |
Zonca F et al 2002 Phys. Plasmas 9 4939 doi: 10.1063/1.1519241
|
[42] |
Wang T et al 2019 Phys. Plasmas 26 012504 doi: 10.1063/1.5064863
|
[43] |
Kimura H et al 1998 Nucl. Fusion 38 1303 doi: 10.1088/0029-5515/38/9/304
|
[44] |
Sharapov S E et al 2001 Phys. Lett. A 289 127 doi: 10.1016/S0375-9601(01)00588-6
|
[45] |
Huang J et al 2020 Nucl. Fusion 60 126007 doi: 10.1088/1741-4326/abaf33
|
[46] |
Zonca F, Chen L and Santoro R A 1996 Plasma Phys. Control. Fusion 38 2011 doi: 10.1088/0741-3335/38/11/011
|
[47] |
Wei S Z et al 2022 Nucl. Fusion 62 126038 doi: 10.1088/1741-4326/ac968f
|
[48] |
Fisch N J and Herrmann M C 1994 Nucl. Fusion 34 1541 doi: 10.1088/0029-5515/34/12/I01
|
[49] |
Wei S Z et al 2021 J. Plasma Phys. 87 905870505 doi: 10.1017/S0022377821000908
|
[50] |
Wang T et al 2024 Plasma Sci. Technol. 26 053001 doi: 10.1088/2058-6272/ad15e0
|
[51] |
Liu P F et al 2023 Rev. Mod. Plasma Phys. 7 15 doi: 10.1007/s41614-023-00117-4
|
[52] |
Ye L, Chen Y and Fu G Y 2023 Nucl. Fusion 63 026004 doi: 10.1088/1741-4326/aca9df
|
[53] |
Chen L and Hasegawa A 1991 J. Geophys. Res.: Space Phys. 96 1503 doi: 10.1029/90JA02346
|
[54] |
Frieman E A and Chen L 1982 Phys. Fluids 25 502 doi: 10.1063/1.863762
|
[55] |
Hasegawa A and Chen L 1976 Phys. Fluids 19 1924 doi: 10.1063/1.861427
|
[56] |
Winsor N, Johnson J L and Dawson J M 1968 Phys. Fluids 11 2448 doi: 10.1063/1.1691835
|
[57] |
Zonca F and Chen L 2008 Europhys. Lett. 83 35001 doi: 10.1209/0295-5075/83/35001
|
[58] |
Wang Y H et al 2022 Plasma Sci. Technol. 24 025105 doi: 10.1088/2058-6272/ac42ba
|
[59] |
Wang T et al 2020 Nucl. Fusion 60 126032 doi: 10.1088/1741-4326/abb2d4
|
[1] | Tao WANG, Shizhao WEI, Sergio BRIGUGLIO, Gregorio VLAD, Fulvio ZONCA, Zhiyong QIU. Nonlinear dynamics of the reversed shear Alfvén eigenmode in burning plasmas[J]. Plasma Science and Technology, 2024, 26(5): 053001. DOI: 10.1088/2058-6272/ad15e0 |
[2] | Yahui WANG, Tao WANG, Shizhao WEI, Zhiyong QIU. Nonlinear excitation of a geodesic acoustic mode by reversed shear Alfvén eignemodes[J]. Plasma Science and Technology, 2022, 24(2): 025105. DOI: 10.1088/2058-6272/ac42ba |
[3] | Peifeng FAN (范培锋), Hong QIN (秦宏), Jianyuan XIAO (肖建元). Discovering exact, gauge-invariant, local energy–momentum conservation laws for the electromagnetic gyrokinetic system by high-order field theory on heterogeneous manifolds[J]. Plasma Science and Technology, 2021, 23(10): 105103. DOI: 10.1088/2058-6272/ac18ba |
[4] | Hongyu WANG (王虹宇), Pengfei LIU (刘鹏飞), Zhihong LIN (林志宏), Wenlu ZHANG (张文禄). Linear gyrokinetic simulations of reversed shear Alfvén eigenmodes and ion temperature gradient modes in DIII-D tokamak[J]. Plasma Science and Technology, 2021, 23(1): 15101-015101. DOI: 10.1088/2058-6272/abc871 |
[5] | Liu CHEN (陈骝), Fulvio ZONCA, Haotian CHEN (陈昊天). Unexpanded nonlinear electromagnetic gyrokinetic equations for magnetized plasmas[J]. Plasma Science and Technology, 2020, 22(10): 102001. DOI: 10.1088/2058-6272/aba187 |
[6] | Tao ZHANG (张涛), Haiqing LIU (刘海庆), Guoqiang LI (李国强), Long ZENG (曾龙), Yao YANG (杨曜), Tingfeng MING (明廷凤), Xiang GAO (高翔), Hui LIAN (连辉), Kai LI (李凯), Yong LIU (刘永), Yingying LI (李颖颖), Tonghui SHI (石同辉), Xiang HAN (韩翔), the EAST team. Experimental observation of reverse- sheared Alfvén eigenmodes (RSAEs) in ELMy H-mode plasma on the EAST tokamak[J]. Plasma Science and Technology, 2018, 20(11): 115101. DOI: 10.1088/2058-6272/aac9b5 |
[7] | Zhiyong QIU, Liu CHEN, Fulvio ZONCA. Kinetic theory of geodesic acoustic modes in toroidal plasmas: a brief review[J]. Plasma Science and Technology, 2018, 20(9): 94004-094004. DOI: 10.1088/2058-6272/aab4f0 |
[8] | Wei WANG (王玮), Zhengxiong WANG (王正汹), Jiquan LI (李继全), Yasuaki KISHIMOTO, Jiaqi DONG (董家齐), Shu ZHENG (郑殊). Magnetic-island-induced ion temperature gradient mode: Landau damping, equilibrium magnetic shear and pressure flattening effects[J]. Plasma Science and Technology, 2018, 20(7): 75101-075101. DOI: 10.1088/2058-6272/aab48f |
[9] | REN Yanqiu (仁艳秋), LI Gun (李滚), DUAN Wenshan (段文山). Damping Solitary Wave in a Three-Dimensional Rectangular Geometry Plasma[J]. Plasma Science and Technology, 2016, 18(2): 108-113. DOI: 10.1088/1009-0630/18/2/02 |
[10] | T. S. HAHM. Ion Heating from Nonlinear Landau Damping of High Mode Number Toroidal Alfvén Eigenmodes[J]. Plasma Science and Technology, 2015, 17(7): 534-538. DOI: 10.1088/1009-0630/17/7/02 |