
Citation: | Na WANG, Zhimin LIU, Yahong XIE, Jianglong WEI, Caichao JIANG, Wei LIU, Xufeng PENG, Guojian SU, Junwei XIE. An analysis and preliminary experiment of the discharge characteristics of RF ion source with electromagnetic shielding[J]. Plasma Science and Technology, 2023, 25(4): 045601. DOI: 10.1088/2058-6272/aca1fb |
Combined with two-dimensional (2D) and three-dimensional (3D) finite element analysis and preliminary experimental tests, the effects of size and placement of the electromagnetic shield of the radio-frequency (RF) ion source with two drivers on plasma parameters and RF power transfer efficiency are analyzed. It is found that the same input direction of the current is better for the RF ion source with multiple drivers. The electromagnetic shield (EMS) should be placed symmetrically around the drivers, which is beneficial for the plasma to distribute uniformly and symmetrically in both drivers. Furthermore, the bigger the EMS shield radius is the better generating a higher electron density. These results will be of guiding significance to the design of electromagnetic shielding for RF ion sources with a multi-driver.
The hybrid kinetic–magnetohydrodynamic (MHD) simulation scheme [1], in comparison with the pure MHD and full particle-in-cell (PIC) simulation methods, has definite advantages that include the wave–particle energy exchange mechanism through the (inverse) Landau damping process and high efficiency in solving the evolution of background plasma. Especially, the hybrid kinetic–MHD simulation model has been widely adopted in simulations for studying the interaction between energetic particles (EPs) with various instabilities in tokamaks, such as the destabilizations of Alfvén eigenmodes and energetic particle modes (EPMs) by EPs through inverse Landau damping [2–4], and the influences of EPs on the linear stabilities of internal kink mode [5, 6] and resistive tearing modes (TMs) [7–11]. Over the past few decades, several hybrid codes have been developed, such as NIMROD [12], M3D-K [13], HMGC [14], MEGA [15], CLT-K [3], and the latest M3D-C1-K code [16].
It is well known that there are two coupling schemes in the hybrid kinetic–MHD model, i.e. pressure coupling and current coupling [1]. In the pressure coupling scheme, the perpendicular momentum evolution of EPs is ignored due to the assumption of nh ≪ n (where nh and n are the densities of EPs and background ions, respectively). In the current coupling scheme, all momentum components of EPs and background plasma are included self-consistently. In the previous version of the CLT-K code, the current coupling is adopted in the simulations [3, 4]. In this work, we bring the pressure coupling scheme into the CLT-K code. Numerical equivalences between these two coupling schemes have been strictly verified under different approximations in nonlinear simulations.
Using the hybrid simulation model, the influences of EPs on the linear stability of TM have been extensively studied through numerical simulations. However, contradictory results were reported in [8] and [11], especially for the situation of co-passing EPs. According to [8], the co-passing EPs play a stabilization role on the m/n = 2/1 TM (where m and n represent the poloidal and toroidal mode numbers, respectively). But in [11], the co-passing EPs play a strong destabilization role on the m/n = 2/1 TM. Based on numerical verifications of the equivalences of the current and pressure coupling schemes, the CLT-K code is adopted to reinvestigate the influences of the co-/counter-passing and trapped EPs on the linear stability of the m/n = 2/1 TM. The co-passing and trapped EPs are found to stabilize the TM, but the counter-passing EPs have a destabilization effect on the TM. After exceeding the critical betas of EPs, the same branch of a high-frequency mode is excited by the co-/counter-passing and trapped EPs, which is identified as the m/n = 2/1 EPM.
The outline of the present paper is as follows: section 2 introduces the simulation model used in the CLT-K code; section 3 presents the verification of the equivalences between the current and pressure coupling schemes; the simulation results about the influences of the co-/counter-passing and trapped EPs on the linear stability of the m/n = 2/1 TM are presented in section 4; the resonant excitations of the m/n = 2/1 EPM are discussed in section 5; finally, the summary and discussion of the simulation results are given in section 6.
For background plasma, CLT-K solves the compressible MHD equations [3, 17–22], covering the plasma density
∂tρ=−▽·(ρv)+▽·[D▽(ρ−ρ0)], | (1) |
∂tp=−v·▽p−Γp▽·v+▽·[κ▽(p−p0)], | (2) |
∂tv=−v·▽v+(J×B−▽p)/ρ+▽·[ν▽(v−v0)], | (3) |
∂tB=−▽×E, | (4) |
with
E=−v×B+η(J−J0), | (5) |
J=▽×B, | (6) |
All variables are nondimensionalized as
For the MHD portion of the CLT-K code, i.e. the CLT code, the cylindrical coordinate system (R, φ, Z) is adopted, where R is the major radius, φ is the toroidal angle, and Z is along the vertical direction. The uniform and rectangular grid is used in the R–Z directions. For the discretization in space, the 4th-order finite difference method is employed in the R and Z directions, while in the φ direction either the 4th-order finite difference or pseudo-spectrum method can be used. As for the time advance, the 4th-order Runge–Kutta scheme is applied.
The contributions of EPs are taken into consideration in momentum equation with the EPs' current
∂tv=−v·▽v+[(J−Jh)×B−▽p]/ρ+▽·[ν▽(v−v0)], | (7) |
∂tv=−v·▽v+[J×B−▽p−(▽·Ph)⊥]/ρ+▽·[ν▽(v−v0)], | (8) |
where the subscript 'h' denotes the EPs. The term
To push EPs, we adopt the guiding-center equations of motion [23], i.e.
dXdt=1B|| | (9) |
(10) |
where
(11) |
(12) |
where
Ignoring polarization current [3, 4],
(13) |
where
(14) |
\mathbf{J}_{\mathrm{MAG}}=\nabla \times \mathcal{M}=-\nabla \times \int \mu \mathbf{b} f \mathrm{~d} v^3. | (15) |
The effective guiding-center drift velocities (curvature drift
(16) |
(17) |
(18) |
By introducing the following intermediate variables, i.e. particle density
(19) |
(20) |
(21) |
(22) |
(23) |
Meanwhile,
(24) |
The
In the hybrid kinetic–MHD model, the additional term introduced by EPs in the momentum equation is the
(25) |
The parallel components of
The initial equilibria of the hybrid simulation are usually obtained by solving the Grad–Shafranov equation. However, the pressure obtained with this method without considering EPs is an isotropic scalar, which means the anisotropy of the EP pressure tensor is not included. The direct inclusion of EPs in simulation will result in an unbalance between the total Lorentz force and the pressures (tensor) of the background plasma and the EPs, i.e. \mathcal{F}=\mathbf{J}_0 \times \mathbf{B}_0-\nabla p_0-\left(\nabla \cdot \mathbf{P}_{\mathrm{h} 0}\right)_{\perp \mathbf{b}_0} \neq 0 . (The subscript means perpendicular to the initial magnetic field.) This force unbalance will significantly affect the early simulation result, especially in linear stages. To avoid this issue, the initial residual force \mathcal{F} is subtracted in equations (7) and (8).
For the current coupling scheme in equation (7), we subtract initial contributions of the background plasma, that is, \mathcal{F}=\left(\mathbf{J}_0-\mathbf{J}_{\mathrm{h} 0}\right) \times \mathbf{B}_0-\nabla p_0 \neq 0 Equation (7) becomes
\begin{aligned} \partial_t \mathbf{v}= & -\mathbf{v} \cdot \nabla \mathbf{v}+\left[\mathbf{J}_0 \times \delta \mathbf{B}\underline{-\mathbf{J}_{\mathrm{h} 0} \times \delta \mathbf{B}}+\left(\delta \mathbf{J}-\delta \mathbf{J}_{\mathrm{h}}\right)\right. \\ & \times \mathbf{B}-\nabla \delta p] / \rho+\nabla \cdot\left[\nu \nabla\left(\mathbf{v}-\mathbf{v}_0\right)\right], \end{aligned} | (26) |
where
\delta \mathbf{J}_{\mathrm{h}}=\mathbf{J}_{\mathrm{h}}-\mathbf{J}_{\mathrm{h} 0} \\ \quad=Z_{\mathrm{h}} \delta\left(n_{\mathrm{h}} V_{\|}\right) \mathbf{b}+\frac{1}{B}\left(\delta P_{\mathrm{h} \|}-\delta P_{\mathrm{h} \perp}\right) \nabla \times \mathbf{b}+\frac{1}{B} \mathbf{b} \times \nabla \delta P_{\mathrm{h} \perp} \\ \underline{+\frac{Z_{\mathrm{h}}\left(n_{\mathrm{h}} V_{\|}\right)_0 \delta \mathbf{b}+\frac{1}{B B_0}\left(P_{\mathrm{h} \| 0}-P_{\mathrm{h} \perp 0}\right)\left(B_0 \nabla \times \delta \mathbf{b}-\delta B \nabla \times \mathbf{b}_0\right)}{+\frac{1}{B B_0}\left(B_0 \delta \mathbf{b}-\delta B \mathbf{b}_0\right) \times \nabla P_{\mathrm{h} \perp 0} .}} | (27) |
For the pressure coupling scheme in equation (8), we subtract contributions of the total plasma, that is, \mathcal{F}=\mathbf{J}_0 \times \mathbf{B}_0-\nabla p_0-\left(\nabla \cdot \mathbf{P}_{\mathrm{h} 0}\right)_{\perp \mathbf{b}_0} \neq 0 Equation (8) becomes
(28) |
where
\begin{aligned} & \delta\left[\left(\nabla \cdot \mathbf{P}_{\mathrm{h}}\right)_{\perp}\right]=\left(\nabla \cdot \mathbf{P}_{\mathrm{h}}\right)_{\perp}-\left(\nabla \cdot \mathbf{P}_{\mathrm{h} 0}\right)_{\perp \mathbf{b}_0} \\ & \quad=\left(\nabla \cdot \delta \mathbf{P}_{\mathrm{h}}\right)_{\perp}\underline{+\left(\nabla \cdot \mathbf{P}_{\mathrm{h} 0}\right)_{\perp}-\left(\nabla \cdot \mathbf{P}_{\mathrm{h} 0}\right)_{\perp \mathbf{b}_0}, } \end{aligned} | (29) |
and
\begin{aligned} & \delta \mathbf{P}_{\mathrm{h}}=\mathbf{P}_{\mathrm{h}}-\mathbf{P}_{\mathrm{h} 0}=\delta P_{\mathrm{h} \perp} \mathbf{I}+\left(\delta P_{\mathrm{h} \|}-\delta P_{\mathrm{h}_{\perp}}\right) \mathbf{b b} \\ & \quad+\underline{\left(P_{\mathrm{h} \| 0}-P_{\mathrm{h} \perp 0}\right)\left(\mathbf{b}_0 \delta \mathbf{b}+\delta \mathbf{b b}\right)} . \end{aligned} | (30) |
If the initial force balance condition is satisfied in an equilibrium considering both contributions of the background plasmas and the EPs, i.e.
First, we ignore the variation of magnetic field
(31) |
(32) |
In addition,
(33) |
and
(34) |
It can be found that the equivalence of the current and pressure coupling schemes based on equations (31)–(34) can be proved similarly following equation (25). In comparison with equations (23) and (24), the only difference that appears in equations (31) and (32) is that the scalar variables of
For most hybrid kinetic–MHD codes, the simplified formulae of equations (31)–(34) are widely adopted. Representatively, equation (31) in the current coupled momentum equation is used in MEGA [15] and the early version of the CLT-K code [3], and equation (32) is used in the pressure tensor coupled momentum equation in NIMROD [12], M3D-K [13], and HMGC [14].
Recently, we have introduced the current and pressure coupling schemes in the CLT-K code [27]. With equations (31)–(34), we have conducted a set of fully nonlinear simulations for the n = 1 TAE to demonstrate the numerical equivalence of these two coupling schemes. A mesh with 200 × 16 × 200 grids in (
Another more complicated situation is to include the contributions of both
Table 1 briefly summarizes the equations and the features of three forms of the current or pressure coupling schemes, i.e. (1) complete form; (2) perturbed form, including terms contributed from both
Items | Complete form | Perturbed form, including and | Perturbed form, including only |
Current coupling | Equations (7) and (23) | Equations (26) and (27) | Equations (31) and (33) |
Pressure coupling | Equations (8) and (24) | Equations (28)–(30) | Equations (32) and (34) |
Additional simplifications | None | The initial force unbalance \mathcal{F} caused by EPs is ignored | The underlined terms in equations (26)–(30) related to and the initial force unbalance \mathcal{F} caused by EPs are ignored |
TAE results | Not available | Larger saturation amplitudes of high-n branches | Smaller saturation amplitudes of high-n branches |
The complete forms of the coupling equations are strictly derived from the momentum equations of EPs and the background plasma. However, these complete coupling equations are unfriendly to the present equilibrium that cannot handle the anisotropic contribution of EPs. As a result, the initial force unbalance \left[\mathcal{F}=\mathbf{J}_0 \times \mathbf{B}_0-\nabla p_0-\left(\nabla \cdot \mathbf{P}_{\mathrm{h} 0}\right)_{\perp \mathbf{b}_0} \neq 0\right] will cause a large perturbation at the initial stage of the simulation.
After we subtracted the initial force unbalance \mathcal{F} in the momentum equations, we obtained the first perturbed form of the coupling equations. This form of equations is exactly the same as the complete form when the initial equilibrium condition, including the contribution of EPs, is satisfied, i.e. \mathcal{F}=\mathbf{J}_0 \times \mathbf{B}_0-\nabla p_0-\left(\nabla \cdot \mathbf{P}_{\mathrm{h} 0}\right)_{\perp \mathbf{b}_0}=0 The expressions of and are complicated because they include perturbations from both of EPs and the perturbed magnetic field
In general, we focus on wave–particle resonance problems, especially for the linear stage of modes. Then, the contribution from
Simulation results from these two forms of perturbed coupling equations are rather similar for TAE, especially for the linear stage. The major difference is found in the saturation levels of high-n branches, which might be resulted from different nonlinear terms included in the equations, as discussed in section 3.3.
In this section, we will systematically discuss the influences of EPs with different distribution functions on the linear stability of the m/n = 2/1 TM. Furthermore, to understand the results, we will analyze the perturbed potential energy
We use a mesh with 200 × 16 × 200 grids in (
(35) |
with 0 (passing EPs) and 1 (trapped EPs), and is the poloidal flux being averaged over the EP orbit [3], where and are, respectively, the poloidal fluxes at the magnetic axis and the boundary. The EP's Larmor radius \varrho_{\mathrm{h}} with the perpendicular speed of is 0.07 Nevertheless, the finite Larmor radius effect is not included in CLT-K because we employ the drift-kinetic model to describe the guiding-center orbits of EPs. The of EPs is scanned with its value at the magnetic axis varying from 0.154% to 3% while the EP pressure profile is fixed. The uniform resistivity used in the simulations is fixed to be Other dissipation coefficients are
Here, we adopt a zero beta (
We first present the simulation results of the m/n = 2/1 TM in the presence of co-passing EPs. The linear growth rate of the m/n = 2/1 TM versus
Next, we subtract the adiabatic contribution in
The influence of counter-passing EPs on the linear stability of the m/n = 2/1 TM is also investigated. As shown in figure 5, within the scanned range of
Here, we briefly study the role of trapped EPs () on the linear stability of the m/n = 2/1 TM. Cai et al adopt a large value of () [8], which results in a quasi-isotropic EP distribution instead of a trapped one. However, here, we adopt a narrow distribution function in pitch angel space () so that most of EPs belong to trapped particles. The results from CLT-K are plotted in figure 6. Qualitatively, the trapped EPs have a similar influence on the stability of the TM as the co-passing EPs. Overall, the total response of trapped EPs slightly stabilizes the TM, but the non-adiabatic response dramatically stabilizes the TM. Nevertheless, the stabilization effect from the non-adiabatic response of trapped EPs is strong so that the TM is completely stabilized when \beta_{\mathrm{h}}^{\mathrm{c}} \gtrsim 0.3 \%.
Similar to the situations of passing EPs, trapped EPs also result in a rotation of the TM in the frequency of , as shown in figure 6(b). However, when including only the non-adiabatic response of trapped EPs and \beta_{\mathrm{h}}^{\mathrm{c}} \gtrsim 0.8 \%, the mode frequency disappears. This is because the mode structure of the TM is unobservable, which indicates the full stabilization of the TM by trapped EPs.
The linear results about passing EPs on the stability of the m/n = 2/1 TM by CLT-K support the analytical and numerical results of Cai et al [7, 8]. For trapped EPs, the distribution function in our simulations greatly differs from that of [8], so the comparison is of lesser significance. According to the theoretical work by Cai et al [7], the non-adiabatic responses of passing EPs mainly influence the perturbed parallel current
(36) |
where
We first analyze the adiabatic response of EPs. For simulation cases in the presence of total response of EPs, the poloidal distribution of the integral term
Next, we analyze
In this section, we will briefly discuss the excitation of the m/n = 2/1 EPM by EPs with different distribution functions. The simulation parameters are the same as in section 4, but the
For co-passing EPs, the growth rates and frequencies of the m/n = 2/1 mode are plotted in figure 9. When the dominant mode is the low-frequency \left(\omega \simeq \omega_{\mathrm{h}}^*\right) m / n=2 / 1 TM. However, when exceeds the threshold an m/n = 2/1 mode with \omega \simeq-0.025 \omega_{\mathrm{A}} is excited by co-passing EPs. The frequency of this m/n = 2/1 mode is much larger than and clearly intersects with the m = 2 component of the shear Alfvén wave (SAW) continuum, as shown in figure 10(a). Correspondingly, the mode structure of the m/n = 2/1 poloidal electric field is plotted in figure 10(b). The mode structure of this high-frequency mode extends toward the magnetic axis and is totally different from that of the m/n = 2/1 TM, as shown in figure 3(b). Consequently, this high-frequency mode is identified as the m/n = 2/1 EPM. In figure 9(a), with the increasing the growth rate of the TM first decreases and then that of EPM increases obviously.
The same branch of EPM (with similar mode frequencies and mode structures as in figure 10) is also observed in cases with counter-passing and trapped EPs. Respectively shown in figures 11 and 12, the excitation thresholds of the m/n = 2/1 EPM are about
In the end, we present resonant conditions of EPs with the m/n = 2/1 EPM. The resonance condition can be written as
In this paper, we strictly verified the equivalences of the current and pressure coupling schemes under different approximations. For most of hybrid MHD–kinetic simulations, only the contribution of
Then, influences of co-/counter-passing and trapped EPs on the linear stability of the m/n = 2/1 TM are carefully reinvestigated by adopting the first approximation method in CLT-K, i.e. considering only the
Finally, we studied the general excitations of the high-frequency m/n = 2/1 EPM by EPs. The transition of TM toward EPM occurs when the beta of EPs is large enough to overcome damping mechanisms, e.g. continuum damping. In our simulations, the beta thresholds for exciting the EPM by different distributed EPs are in the order as follows: the co-passing EPs have the smallest beta threshold, then the trapped EPs in second, and the beta threshold of counter-passing EPs is the largest. The simulation results partially confirm the analytical research by Zhang et al [31], which focused on the trapped EPs. In addition, the excitation behaviors of EPM by co-/counter-passing EPs via different resonance conditions [32] are also systematically discussed.
This work was supported by the Comprehensive Research Facility for Fusion Technology Program of China (No. 2018-000052-73-01-001228), National Natural Science Foundation of China (No. 11975263) and the National Key R & D Program of China (No. 2017YFE0300101).
[1] |
Xie Y H et al 2021 Plasma Sci. Technol. 23 012001 doi: 10.1088/2058-6272/abc46b
|
[2] |
Wang N et al 2022 Fusion Eng. Des. 183 113272 doi: 10.1016/j.fusengdes.2022.113272
|
[3] |
Xie Y H et al 2021 Fusion Eng. Des. 167 112377 doi: 10.1016/j.fusengdes.2021.112377
|
[4] |
Wei J L et al 2021 Fusion Eng. Des. 169 112482 doi: 10.1016/j.fusengdes.2021.112482
|
[5] |
Wünderlich D et al 2021 Nucl. Fusion 61 096023 doi: 10.1088/1741-4326/ac1758
|
[6] |
Wünderlich D et al 2019 Nucl. Fusion 59 084001 doi: 10.1088/1741-4326/ab246c
|
[7] |
Kraus W et al 2008 Rev. Sci. Instrum. 79 02C108 doi: 10.1063/1.2804917
|
[8] |
Maistrello A et al 2021 Fusion Eng. Des. 167 112337 doi: 10.1016/j.fusengdes.2021.112337
|
[9] |
Kraus W et al 2012 Rev. Sci. Instrum. 83 02B104 doi: 10.1063/1.3662957
|
[10] |
Jain P et al 2022 Plasma Phys. Control. Fusion 64 095018 doi: 10.1088/1361-6587/ac8617
|
[11] |
Kraus W et al 2018 Rev. Sci. Instrum. 89 052102 doi: 10.1063/1.5012591
|
[12] |
Xie Y H et al 2019 Rev. Sci. Instrum. 90 113319 doi: 10.1063/1.5128258
|
[13] |
Hopwood J 1994 Plasma Sources Sci. Technol. 3 460 doi: 10.1088/0963-0252/3/4/002
|
[14] |
Rauner D et al 2022 Plasma 5 280 doi: 10.3390/plasma5030022
|
[15] |
Fantz U et al 2007 Plasma Phys. Control. Fusion 49 B563 doi: 10.1088/0741-3335/49/12B/S53
|
[16] |
Kraus W et al 2015 Fusion Eng. Des. 91 16 doi: 10.1016/j.fusengdes.2014.11.015
|
[17] |
Heinemann B et al 2017 New J. Phys. 19 015001 doi: 10.1088/1367-2630/aa520c
|
[18] |
Bandyopadhyay M et al 2011 Two RF driver based negative ion source for fusion R&D IEEE/NPSS 24th Symp. on Fusion Engineering (Chicago) (IEEE) 1
|
[19] |
Jain P et al 2018 Plasma Phys. Control. Fusion 60 045007 doi: 10.1088/1361-6587/aaab19
|
[20] |
Janev R K, Reiter D and Samm U 2003 Collision Processes in
Low-Temperature Hydrogen Plasmas Report Jül-4105
Forschungszentrum Jülich
|
[21] |
Wang Y J et al 2021 Chin. Phys. B 30 095205 doi: 10.1088/1674-1056/ac0e21
|
[22] |
Zielke D 2021 Development of a predictive self-consistent fluid model for optimizing inductive RF coupling of powerful negative hydrogen ion sources PhD Thesis University of Augsburg, Augsburg, Germany
|
[23] |
Pitchford L C et al 2017 Plasma Process. Polym. 14 1600098 doi: 10.1002/ppap.201600098
|
[24] |
Grudiev A et al 2014 Rev. Sci. Instrum. 85 02B134 doi: 10.1063/1.4842317
|
[25] |
Hagelaar G J M and Pitchford L C 2005 Plasma Sources Sci. Technol. 14 722 doi: 10.1088/0963-0252/14/4/011
|
[26] |
Gogolides E and Sawin H H 1992 J. Appl. Phys. 72 3971 doi: 10.1063/1.352250
|
[27] |
Dutton J 1975 J. Phys. Chem. Ref. Data 4 577 doi: 10.1063/1.555525
|
[1] | Yifei ZHAO, Yueqiang LIU, Guangzhou HAO, Zhengxiong WANG, Guanqi DONG, Shuo WANG, Chunyu LI, Guanming YANG, Yutian MIAO, Yongqin WANG. Loss of energetic particles due to feedback control of resistive wall mode in HL-3[J]. Plasma Science and Technology, 2024, 26(10): 104002. DOI: 10.1088/2058-6272/ad547e |
[2] | Jixing YANG, Guoyong FU, Wei SHEN, Minyou YE. Linear hybrid simulations of low-frequency fishbone instability driven by energetic passing particles in tokamak plasmas[J]. Plasma Science and Technology, 2022, 24(6): 065101. DOI: 10.1088/2058-6272/ac5972 |
[3] | Weikang TANG (汤炜康), Lai WEI (魏来), Zhengxiong WANG (王正汹), Jialei WANG (王佳磊), Tong LIU (刘桐), Shu ZHENG (郑殊). Effects of resonant magnetic perturbation on locked mode of neoclassical tearing modes[J]. Plasma Science and Technology, 2019, 21(6): 65103-065103. DOI: 10.1088/2058-6272/ab0a18 |
[4] | Yun YUAN (袁赟), Xingqiang LU (路兴强), Jiaqi DONG (董家齐), Zhixiong HE (何志雄), Ruibo ZHANG (张睿博), Shijia CHEN (陈诗佳), Xueyu GONG (龚学余), Yun YUAN (袁赟), Xingqiang LU (路兴强), Jiaqi DONG (董家齐), Zhixiong HE (何志雄), Ruibo ZHANG (张睿博), Shijia CHEN (陈诗佳), Xueyu GONG (龚学余). Influence of stationary driven helical current on the m=2/n=1 resistive tearing mode[J]. Plasma Science and Technology, 2019, 21(5): 55101-055101. DOI: 10.1088/2058-6272/aafdc7 |
[5] | X T DING (丁玄同), W CHEN (陈伟). Review of the experiments for energetic particle physics on HL-2A[J]. Plasma Science and Technology, 2018, 20(9): 94008-094008. DOI: 10.1088/2058-6272/aad27a |
[6] | Ding LI (李定), Wen YANG (杨文), Huishan CAI (蔡辉山). On theoretical research for nonlinear tearing mode[J]. Plasma Science and Technology, 2018, 20(9): 94002-094002. DOI: 10.1088/2058-6272/aabde4 |
[7] | Guiliang SONG (宋桂良), Huishan CAI (蔡辉山). Linear tearing modes in an electron-positron plasma[J]. Plasma Science and Technology, 2017, 19(4): 45002-045002. DOI: 10.1088/2058-6272/aa5801 |
[8] | LI Dehui(李德徽), XIANG Nong(项农), LIN Yu(林郁), WANG Xueyi(汪学毅), YANG Cheng(杨程), MA Jun(马骏). Benchmark Simulations of Gyro-Kinetic Electron and Fully-Kinetic Ion Model for Lower Hybrid Waves in Linear Region[J]. Plasma Science and Technology, 2014, 16(9): 821-825. DOI: 10.1088/1009-0630/16/9/03 |
[9] | WANG Jun (王军), HU Chundong (胡纯栋), HU Shuanghui (胡双辉), WU Bin (吴斌), et al.. Alfvén Instabilities Excited by Energetic Particles in a Parameter Regime Similar to EAST Operation[J]. Plasma Science and Technology, 2013, 15(8): 750-754. DOI: 10.1088/1009-0630/15/8/06 |
[10] | QIU Zhiyong, Fulvio ZONCA, CHEN Liu. Kinetic Theories of Geodesic Acoustic Modes: Radial Structure, Linear Excitation by Energetic Particles and Nonlinear Saturation[J]. Plasma Science and Technology, 2011, 13(3): 257-266. |
1. | Zhang, H.X., Zhang, H.W., Ma, Z.W. et al. Influences of δB contribution and parallel inertial term of energetic particles on MHD-kinetic hybrid simulations: a case study of the 1/1 internal kink mode. Plasma Physics and Controlled Fusion, 2025, 67(1): 015033. DOI:10.1088/1361-6587/ad994c |
2. | Wang, H., Jiang, S., Liu, T. et al. Effects of diamagnetic drift on nonlinear interaction between multi-helicity neoclassical tearing modes. Chinese Physics B, 2024, 33(6): 065202. DOI:10.1088/1674-1056/ad24d3 |
3. | Xu, L., Lin, S., Mai, C. et al. Observation of fast electron redistribution during saturated kink mode in high β p H-mode discharge with central heating in EAST tokamak. Nuclear Fusion, 2024, 64(4): 046024. DOI:10.1088/1741-4326/ad2d3a |
4. | Ma, Y., Zhang, R., Cai, H. Numerical calculation for instability criterion of tearing modes influenced by energetic ions. Plasma Physics and Controlled Fusion, 2023, 65(11): 115006. DOI:10.1088/1361-6587/acfff2 |
Items | Complete form | Perturbed form, including and | Perturbed form, including only |
Current coupling | Equations (7) and (23) | Equations (26) and (27) | Equations (31) and (33) |
Pressure coupling | Equations (8) and (24) | Equations (28)–(30) | Equations (32) and (34) |
Additional simplifications | None | The initial force unbalance \mathcal{F} caused by EPs is ignored | The underlined terms in equations (26)–(30) related to and the initial force unbalance \mathcal{F} caused by EPs are ignored |
TAE results | Not available | Larger saturation amplitudes of high-n branches | Smaller saturation amplitudes of high-n branches |