Processing math: 13%
Advanced Search+
Penghao ZHANG, Liang YU, Dazhao HE, Xinyu TANG, Shuo CHEN, Shoulong DONG, Chenguo YAO. All-organic modification coating prepared with large-scale atmospheric-pressure plasma for mitigating surface charge accumulation[J]. Plasma Science and Technology, 2023, 25(8): 084001. DOI: 10.1088/2058-6272/aca8ec
Citation: Penghao ZHANG, Liang YU, Dazhao HE, Xinyu TANG, Shuo CHEN, Shoulong DONG, Chenguo YAO. All-organic modification coating prepared with large-scale atmospheric-pressure plasma for mitigating surface charge accumulation[J]. Plasma Science and Technology, 2023, 25(8): 084001. DOI: 10.1088/2058-6272/aca8ec

All-organic modification coating prepared with large-scale atmospheric-pressure plasma for mitigating surface charge accumulation

More Information
  • Corresponding author:

    Liang YU, E-mail: yu_liang@cqu.edu.cn

  • Received Date: September 13, 2022
  • Revised Date: December 01, 2022
  • Accepted Date: December 04, 2022
  • Available Online: December 05, 2023
  • Published Date: April 27, 2023
  • The surface charge accumulation on polymers often leads to surface flashover. Current solutions are mainly based on the introduction of inorganic fillers. The high-cost process and low compatibility remain formidable challenges. Moreover, existing researches on all-organic insulation focus on capturing electrons, contrary to alleviating charge accumulation. Here, an all-organic modification coating was prepared on polystyrene (PS) with the large-scale atmospheric-pressure plasma, which exhibits outperformed function in mitigating surface charge accumulation. The surface charge dissipation rate and surface conductivity are promoted by about 1.37 and 9.45 times, respectively. Simulation and experimental results show that this all-organic modification coating has a smaller electron affinity potential compared with PS. The decrease of electron affinity potential may result in accelerated surface charge decay of PS, which has never been involved in previous works. Moreover, this coating also has good reliability in a repeated surface flashover. This facile and large-scale approach brings up a novel idea for surface charge regulation and the manufacture of advanced dielectric polymers.

  • 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 [24], and the influences of EPs on the linear stabilities of internal kink mode [5, 6] and resistive tearing modes (TMs) [711]. 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 nhn (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, 1722], covering the plasma density ρ, thermal pressure p, plasma velocity v, magnetic field B, electric field E, and current density J,

    tρ·(ρv)+·[D(ρρ0)], (1)
    tpv·pΓp·v+·[κ(pp0)], (2)
    tvv·v+(J×Bp)/ρ+·[ν(vv0)], (3)
    tB×E, (4)

    with

    Ev×B+η(JJ0), (5)
    J×B, (6)

    All variables are nondimensionalized as B/B00B, x/ax, ρ/ρ00ρ, v/vAv, t/τAt, γandω/ωAγandω, p/(B2axis0/μ0)p, J/(B00/μ0a)J, E/(vABaxis0)E, η/(μ0vAa)η, and D,κ,andν/(vAa)D,κ,andν, where a is the poloidal minor radius; vABaxis0/μ0ρaxis0 is the Alfvén speed; and τAa/vA and ωAvA/a are the Alfvén time and Alfvén frequency, respectively. The variable with subscript 'axis0' corresponds to its initial value at the magnetic axis. Γ is the adiabatic index (5/3).

    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 RZ 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 Jh or pressure tensor Ph in the hybrid kinetic–MHD model [1], i.e.

    tvv·v+[(JJh)×Bp]/ρ+·[ν(vv0)], (7)
    tvv·v+[J×Bp(·Ph)]/ρ+·[ν(vv0)], (8)

    where the subscript 'h' denotes the EPs. The term ZhnheE (where Zh is the charge number of EPs, nh is the particle density of EPs, and e is the unit charge) counteracts the E×B drift in the EP current and is eliminated in equation (7).

    To push EPs, we adopt the guiding-center equations of motion [23], i.e.

    dXdt1B|| (9)
    dv||dtZheMB||B·E, (10)

    where M is the EP mass. The modified fields B and E are expressed in terms of the new effective electrostatic potential Φ and the vector potential A, i.e. B×A, B||B·b, (bB/B), and EΦtA. Φ and A are defined as

    ΦΦ+1ZheμB, (11)
    AA+mv||Zheb, (12)

    where μ, Φ, and A are, respectively, the magnetic moment, electric potential, and magnetic vector potential. The parallel and perpendicular components are defined based on the unit vector of the magnetic field b, e.g. v||v·bb and vb×v×b.

    Ignoring polarization current [3, 4], Jh in equation (7) mainly includes the current of guiding-center motion JGC and magnetization current JMAG,

    JhJGC+JMAG, (13)

    where

    JGCZhe(vcurvature+vB+vB)fdv3, (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 vcurvature, magnetic drift vB, and parallel motion vB) are

    vcurvatureMv||2ZheB||×b, (16)
    vBμZheB||b×B, (17)
    vBv||BB||. (18)

    By introducing the following intermediate variables, i.e. particle density nh, particle flux density nhVh||, perpendicular pressure component Ph, and parallel pressure component Ph||,

    nhfdv3, (19)
    nhVh||v||fdv3, (20)
    PhμBfdv3, (21)
    Ph||Mv||2fdv3, (22)

    Jh in equation (13) can be expressed as

    JhZhenhVh||b+1B(Ph||Ph)×b+1Bb×Ph. (23)

    Meanwhile, Ph is written in the Chew–Goldberger–Low (CGL) form [24]

    PhPhI+(Ph||Ph)bb. (24)

    The δf method [3, 25, 26] is used in CLT-K to solve EP distribution function. At present, the contribution of EPs to equilibrium is also ignored.

    In the hybrid kinetic–MHD model, the additional term introduced by EPs in the momentum equation is the Jh×B in equation (7) or (·Ph) in equation (8). According to equations (23) and (24), the analytical equivalence between the two additional terms can be easily proved as

    Jh×B(Ph||Ph)×b+b×Ph×bb×·PhI+(Ph||Ph)bb×b(·Ph) (25)

    The parallel components of Jh and ·Ph do not contribute to the momentum evolution of EPs. It can also be found that with the CGL pressure formulation as equation (24), the variation of the perpendicular momentum of the EPs is completely ignored due to the perfect cancellation between Jh×B and (·Ph). Consequently, the current coupling and pressure coupling schemes for equations (7) and (8) are exactly the same when the full distribution function of EPs is used to calculate the total Ph and Jh as equations (23) and (24).

    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

    tvv·v+J0×δB+δJ×Bδpδ·Ph/ρ+·νvv0, (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. J0×B0p0·Ph0b00 and J0Jh0×B0p00, equations (26)–(30) will be completely equivalent to the total forms of the momentum equations of equations (7), (8), (23), and (24). Thus, there are two parts of contributions in the perturbed EP current and pressure tensor. The black perturbation terms without underline in equations (26)–(30) are caused by the variation of the EP distribution function δf, while the underlined terms in equations (26)–(30) mainly result from the magnetic field perturbation δB. In the following part, it will be proved that the current and pressure coupling schemes are equivalent when only considering the δf contribution or both terms resulted from δf and δB, respectively.

    First, we ignore the variation of magnetic field δB, i.e. only the black terms without underline related to δf in equations (26)–(30) are included. Then, equations (27) and (29) become

    δJhZhδnhV||b+1BδPh||δPh×b+1Bb×δPh, (31)
    δ·Ph·δPh·δPhI+δPh||δPhbb. (32)

    In addition, Jh0×δB in equation (26) should also be ignored. The momentum equations become

    tvv·v+J0×δB+δJδJh×Bδp/ρ+·νvv0, (33)

    and

    tvv·v+J0×δB+δJ×Bδp·δPh/ρ+·νvv0. (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 nhV||, Ph||, and Ph are replaced by their variations as δnhV||, δPh||, and δPh.

    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 (R, φ, Z) is adopted. About 40 markers are loaded in each grid cell. These parameters guarantee numerical convergence. The initial equilibrium and simulation parameters used are the same as those in previous work [27]. Here, we use a low resistivity as η107. The nonlinear evolutions of TAE are shown in figure 1. The results obtained with the current and pressure coupling schemes are exactly the same for the n = 1 TAE branch and the n ≠ 1 nonlinear sidebands. The eigenmode structure and frequency obtained from CLT-K can be found in figure 5 of [27] and will not be presented here again.

    Figure  1.  Time evolutions of the kinetic energy in the n = 1 TAE nonlinear simulations with the current coupling (solid lines) and pressure coupling (dotted lines) schemes. Only the variation of the EP distribution function δf is considered.

    Another more complicated situation is to include the contributions of both δf and δB as equations (26)–(30). The equivalence of the current and pressure coupling schemes with equations (26)–(30) is hard to be demonstrated analytically. We adopted equations (26)–(30) in the latest CLT-K code and carried out a set of nonlinear simulations for the n = 1 TAE with the same parameters, as shown in figure 1. According to the results shown in figure 2, the numerical equivalence between the two coupling schemes is well satisfied. In the linear stage of TAE, the results between the two methods, as shown in figures 1 and 2, are almost the same. However, the major difference made by the variation of the magnetic field appears significantly in the nonlinear saturation stage of TAE, especially for the n = 3 and 4 branches. A possible explanation is that by including the underlined terms related to the magnetic field perturbation in δJh×B and δ·Ph, the nonlinearity of the magnetic field will be taken into account self-consistently. With the underlined terms in equations (26)–(30), the nonlinearity of the magnetic field is of the order of OδB2. Otherwise, the nonlinearity included in the reduced δJh×B or δ·Ph (without the underlined terms) mainly comes from the perturbed EP distribution function δf and the magnetic field perturbation δB, i.e. OδfδB. The nonlinearity of the magnetic field tends to generate nonlinear branches with higher n, e.g. n = 3 and 4. On the other hand, in our simulations, we only considered the n = 1 component of δf. In consequence, the nonlinearity of δf and δB, i.e. OδfδB, will mainly generate the lower-n nonlinear branches, e.g. n = 0, 1, and 2.

    Figure  2.  Time evolutions of the kinetic energy in the n = 1 TAE nonlinear simulations with the current coupling (solid lines) and pressure coupling (dotted lines) schemes. Both the variations of the EP distribution function δf and the magnetic field δB are considered.

    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 δf and δB; and (3) perturbed form mainly considering the contribution of δf. For each form, the current and pressure coupling schemes are basically equivalent to each other.

    Table  1.  Comparison of different forms of the current or pressure coupling schemes.
    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
     | Show Table
    DownLoad: CSV

    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 δf will be more important. In this case, it is reasonable to ignore the underlined perturbed terms associated with δB, and we can get the second simplified perturbed form of the coupling equations. The last perturbed form of the coupling equations is also widely adopted by various hybrid codes, as listed in the introduction and in section 3.2.

    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 δWh caused by EPs. At TM's linear stage, because the magnetic field perturbation δB is small, we adopt the simplified coupling schemes (only including the contribution of δf). The results from the current and pressure coupling schemes are exactly the same.

    We use a mesh with 200 × 16 × 200 grids in (R, φ, Z) for all cases, and about 40 markers are loaded in each grid cell. These parameters guarantee the numerical convergence. The anisotropic slowing down distribution with the pitch angle ΛμB0/E is used for EPs

    f01v3+vc31+erfv0vΔvexpΛΛ0ΔΛ2×expψΔψ, (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 (βb = 0) tokamak equilibrium, which is obtained from the QSOLVER code [28] with shifted circular flux surfaces. The aspect ratio is R0/a = 5.76. The plasma density is uniform. The profile of the safety factor is plotted in figure 3(a). The m/n = 2/1 TM's instability parameter Δ' is 14.8 [29], which means the m/n = 2/1 TM is unstable. The saturated mode structure Eφ of the m/n = 2/1 TM without EPs is plotted in figure 3(b). We mainly discuss the influences of EPs on the m/n = 2/1 TM. The n = 1 perturbed δJh and δPh are included with only the contribution of δf as equations (31)–(34) in the simulations. In addition, δf associated with the EP response can be separated into the adiabatic and non-adiabatic parts [8, 30]. The adiabatic term ξ·f0 in δf is the fluidlike response of the EPs (where ξ is the plasma displacement), which will be subtracted in δf to investigate its role in the (de)stabilization of TMs. The non-adiabatic part of δf as δf+ξ·f0, mainly accounts for the resonance and the energy exchange between EPs and waves. To get comparable results, the initial EP distribution and tokamak equilibrium are the same as those adopted by Cai et al [8].

    Figure  3.  (a) Initial profile of equilibrium's safety factor q. (b) The mode structure of the m/n = 2/1 TM without EPs in the nonlinear saturated stage. The initial q = 2 surface is denoted by the dashed line.

    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 βhc of co-passing EPs is plotted in figure 4(a). When including the total response of EPs, as shown by red circles in figure 4(a), the growth rate of the TM decreases slightly with the increase of βhc, which means that EPs, in general, play a weak stabilization effect on the TM. The frequency of the TM is very close to the analytical diamagnetic drift frequency ωh* of EPs at the q = 2 surface; see figure 4(b). The negative frequency indicates that the mode rotates in the direction of the ion-diamagnetic drift. The mode frequency in figure 4(b) is missing when βhc>1.0%. This is because a high-frequency m/n = 2/1 EPM is excited, which results in difficulty in the identification of the TM frequency (if TM exists).

    Figure  4.  (a) Linear growth rates and (b) mode frequencies (at the q = 2 surface) of the m/n = 2/1 mode versus βhc of co-passing EPs. The results with (without) the adiabatic term are indicated by red circles (blue squares). The diamagnetic drift frequency of the analytical EPs at the q = 2 surface is also plotted by green stars in (b).

    Next, we subtract the adiabatic contribution in δf of EPs, that is, only the non-adiabatic response of EPs is included in simulations. As shown by blue squares in figure 4(a), without the adiabatic term, the non-adiabatic response of co-passing EPs plays a strong effective role in the stabilization of the TM. For 0.40% βhc 1.0%, the TM is fully stabilized by the non-adiabatic response of EPs. The mode frequency can still be measured because we initially add an m/n = 2/1 TM-like perturbation around the q = 2 rational surface to trigger the fast growth of the m/n = 2/1 TM. With the adiabatic response of co-passing EPs, the growth rate of TM is much larger than that without the adiabatic response. Consequently, the adiabatic response of co-passing EPs has a strong destabilization effect on the TM. For βhc<0.40%, the mode frequencies of the TM without adiabatic term are also quite close to those in the presence of both the adiabatic and non-adiabatic terms.

    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 βhc (<1.15%), the dominant mode in the system is the TM with low diamagnetic drift frequency. The total response of counter-passing EPs has a small destabilization effect on the TM, as indicated by red circles in figure 5(a). Meanwhile, the non-adiabatic response of counter-passing EPs significantly destabilizes the TM with the growth rate even increased by a factor of 3. Thus, the adiabatic response of counter-passing EPs greatly stabilizes the TM. The frequency of the TM resulting from counter-passing EPs is also comparable with the diamagnetic drift frequency of the analytical EPs, as shown in figure 5(b).

    Figure  5.  (a) Linear growth rates and (b) mode frequencies (at the q = 2 surface) of the m/n = 2/1 TM versus βhc of counter-passing EPs. The results with (without) the adiabatic term are indicated by red circles (blue squares). The diamagnetic drift frequency of the analytical EPs at the q = 2 surface is also plotted by green stars in (b).

    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 \%.

    Figure  6.  (a) Linear growth rates and (b) mode frequencies (at the q = 2 surface) of the m/n = 2/1 TM versus βhc of trapped EPs. The results with (without) the adiabatic term are indicated by red circles (blue squares). The diamagnetic drift frequency of the analytical EPs at the q = 2 surface is also plotted by green stars in (b).

    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 δJ|| in the ideal region, which results in the modifications of Δ' and the linear stability of the TM. In our hybrid simulation, the numerical noise caused by PIC simulation makes δJ|| difficult to be analyzed. In the following part, to understand the influence of EPs on the stability of the TM, we will analyze the potential energy δWh caused by EPs with different distribution functions. By taking the pressure coupling as an example, δWh can be written as

    δWh12dx3ξ*··δPh12dx3ξ*·κδPh||δPh+ξ*·δPh, (36)

    where κb·b is the magnetic field curvature. The same formula of δWh can be obtained by replacing ·Ph in equation (36) by Jh×B for the current coupling scheme.

    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 ξ··Phadiabatic at the φ0 cross-section is plotted in figure 7. Co-passing and trapped EPs have stronger adiabatic responses in the low field side (LFS), while the main response of counter-passing EPs appears in the high field side (HFS). The adiabatic response of EPs peaks at the X and O points of the TM, where the plasma displacement ξ is the strongest. The poloidal asymmetry of ξ··Phadiabatic is mainly caused by the orbit shifts of EPs. The (de)stabilization roles of the adiabatic responses of the EPs can be understood based on the sign of ξ··Phadiabatic. Specifically, in figure 7(a), the adiabatic response of co-passing EPs is negative at the X and O points of the TM, which indicates a destabilization effect on the TM. On the contrary, the adiabatic response of counter-passing EPs mainly stabilizes the TM since ξ··Phadiabatic is positive in figure 7(b). For trapped EPs in figure 7(c), the adiabatic response dominantly plays a destabilization role in the LFS but a weak stabilization role in the HFS. Overall, the adiabatic response of trapped EPs destabilizes the TM. The adiabatic responses of EPs are well consistent with the linear simulation results shown in figures 46.

    Figure  7.  The 2D distribution of ξ··Phadiabatic in the presence of total response of EPs (βhc0.57%; t2991.2τA): (a) co-passing EPs, (b) counter-passing EPs, and (c) trapped EPs. The dashed line indicates the q = 2 surface. The Poincaré plots of magnetic field lines are presented by dotted lines. Note that the perturbed magnetic fields are enlarged to clearly show the locations of the X and O points of magnetic islands.

    Next, we analyze δWh generated by the total response of EPs so as to understand the behind (de)stabilization mechanisms of the TM by EPs. The integral term ξ··Ph contributed by the total δf contains high numerical noise. So, we integrate ξ··Ph along the toroidal direction and plot the 2D distribution of 12ξ··PhRdφ in figure 8. The value of δWh is also calculated by performing 3D volume integral of ξ··Ph. The total response of co- and counter-passing EPs dominates in the LFS and HFS, respectively, while the response of trapped EPs has a quasi-symmetric poloidal distribution. The different distributions of δWh are related to the various features of EP orbits. The orbits of co- (counter)-passing EPs are significantly shifted toward LFS (HFS). The m/n = 2/1 TM is located around the q = 2 rational surface. As a result, the dominant interaction of co- (counter)-passing EPs and TM will naturally take place in the LFS (HFS) of the q = 2 rational surface. However, for the case with trapped EPs, we used a moderate pitch-angle parameter Λ01.0, and particles with a large pitch angle will also be randomly loaded in the HFS initially, which together result in a non-negligible portion of barely passing EPs in addition to the trapped EPs. In our simulations, it is difficult to separate or remove the barely passing EPs since their pitch angle is also large. For trapped EPs, they are mainly located in the LFS due to their orbits, which dominantly contributes to δWh in the LFS. But, for the barely passing EPs, they will spend much longer time due to their small v|| in the HFS. As a result, the interaction of the barely passing EPs and the m/n = 2/1 TM will be obvious in the HFS. Thus, the distribution of δWh for the case with trapped EPs is quasi-symmetric in the poloidal section. On the other hand, inside (outside) the q = 2 surface, EPs with different distributions play stabilizing (destabilizing) roles on the TM. The antisymmetric distribution of 12ξ··PhRdφ around the q = 2 surface is related to the odd symmetric mode structure of the m/n = 2/1 TM. The integrated values of δWh in figure 8 show that co-passing and trapped EPs play stabilizing roles on the TM (δWh>0), while counter-passing EPs destabilize the TM (δWh<0). The analyses of δWh also agree with the linear results of figures 46.

    Figure  8.  The 2D distribution of 12ξ··PhRdφ in the presence of the total response of EPs (βhc0.57%; t2991.2τA): (a) co-passing EPs, (b) counter-passing EPs, and (c) trapped EPs. The integrated values of δWh are marked out. The dashed line indicates the q = 2 surface.

    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 βhc of EPs are increased to higher values.

    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.

    Figure  9.  (a) Linear growth rates and (b) mode frequencies (at the q = 2 surface) of the m/n = 2/1 mode versus βhc of co-passing EPs. The results with (without) the adiabatic term are indicated with red circles (blue squares). The diamagnetic drift frequency of the analytical EPs is plotted by green stars in (b).
    Figure  10.  (a) Mode-frequency spectrum and (b) mode structure of the m/n = 2/1 poloidal electric field Eθ in the presence of co-passing EPs with βhc1.43%. The n = 1 SAW continuum is plotted with black solid lines (a). The dashed line in (b) indicates the q = 2 surface.

    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 2% for counter-passing EPs and 1.5% for trapped EPs. The frequency spectra of the EPM in the presence of counter-passing and trapped EPs are very similar to those in figure 10. Meanwhile, as shown in figures 9 and 11, the adiabatic responses of co- and counter-passing EPs have very limited influences on the excitation of EPM. However, the adiabatic response of trapped EPs greatly promotes the excitation of EPM, which is similar to the result reported in [8]. This should be caused by the wide banana orbit of trapped EPs so that the trapped EPs have a much larger adiabatic response on EPM than passing EPs. In the recent theoretical work by Zhang et al [31], the excitation of the m/n = 2/1 EPM by the non-adiabatic response of trapped EPs has been systematically discussed. Nonetheless, for passing EPs, they assumed the orbit width is small, so that the leading order of non-adiabatic response from passing EPs can be ignored and the m/n = 2/1 EPM is hard to be excited [31]. According to our simulations, co-passing EPs have the smallest threshold for the excitation of the EPM, indicating that the finite orbit width effect of passing EPs is non-negligible.

    Figure  11.  (a) Linear growth rates and (b) mode frequencies (at the q = 2 surface) of the m/n = 2/1 mode versus βhc of counter-passing EPs. The results with (without) the adiabatic term are indicated by red circles (blue squares). The diamagnetic drift frequency of the analytical EPs is plotted by green stars in (b).
    Figure  12.  (a) Linear growth rates and (b) mode frequencies (at the q = 2 surface) of the m/n = 2/1 mode versus βhc of trapped EPs. The results with (without) the adiabatic term are indicated by red circles (blue squares). The diamagnetic drift frequency of the analytical EPs is plotted by green stars in (b).

    In the end, we present resonant conditions of EPs with the m/n = 2/1 EPM. The resonance condition can be written as ω+pωθnωφ0, where ωθ and ωφ are the poloidal and toroidal orbit frequencies, respectively; p is an integer with pl+m, while l is the Fourier component of the poloidal motion of EPs. The δf structures of EPs with different distribution functions are plotted in figure 13. For co-passing EPs (Λ0.25) in figure 13(a), the dominant resonance condition is p1. The p1 resonance condition is different from previous report [32], where the p2 resonance condition was found by M3D-K simulations. This difference should be caused by the different energies of co-passing EPs adopted in simulations of CLT-K (v00.5vA) and M3D-K (v0>vA). On the other hand, with our simulation parameters and initial equilibrium, p = 2, n = 1, ω0.020ωA resonance condition of co-passing EPs, and the m/n = 2/1 EPM is not covered by the phase space of EPs. For counter-passing EPs, the p3 resonance condition is found to be dominant, which also results in a significant radial transport of counter-passing EPs; see figure 13(b). The large increase of δf near the plasma boundary, i.e. the upper boundary in figure 13(b), indicates a significant loss of counter-passing EPs caused by the EPM. Meanwhile, the precession frequency of trapped EPs matches the EPM frequency. Thus, the dominant resonance condition is p0 for trapped EPs. Also, higher resonance conditions are observed in figure 13(c), such as p1 and 2.

    Figure  13.  The δf structures of EPs for cases in the presence of (a) co-passing (Λ0.25), (b) counter-passing (Λ0.25), and (c) trapped (Λ1.0) EPs. The different resonance conditions between EPs and TM are plotted with different lines.

    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 δf is considered, which captures the most important wave–particle resonance and the adiabatic response of EPs. However, if the magnetic perturbation caused by instabilities grows to a large amplitude, the contribution of δB in the perturbed pressure tensor or the current of EPs should also be retained. The significant difference by the δB contribution can be found in the nonlinear stage of TAE, as discussed in section 3. To describe the evolution of plasma in the presence of EPs more accurately, a self-consistent equilibrium is necessary to include the initial contribution of EPs properly, e.g. the anisotropic pressure equilibrium, including parallel and perpendicular pressure components. The optimization of such an equilibrium is complicated but worth pursuing further, because it directly influences the widely used Grad–Shafranov model and the closure of the MHD equations.

    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 δf contribution of EPs. The contribution of δB in the perturbed pressure tensor or the current of EPs is ignored because of the small amplitude of δB in the linear stage. The CLT-K simulations show that non-adiabatic responses of co-passing and trapped EPs stabilize the TM, while that of counter-passing EPs has the opposite effect. Moreover, the adiabatic responses of EPs with different distributions tend to counteract the non-adiabatic (de)stabilization effects of co-/counter-passing and trapped EPs. The behind (de)stabilization mechanisms are numerically analyzed for the first time based on the perturbed potential energy δWh produced by EPs. The simulation results and numerical analyses both present a consistent conclusion as Cai et al [7, 8] for co-/counter-passing EPs.

    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.

    We acknowledged the Graduate Student Research Innovation Project of Chongqing (No. CYB22016), National Natural Science Foundation of China (Nos. 52237010, 52277135, 51907011). The authors also thank Xiaoyu Tang (Chongqing Normal University), Prof Tao Shao (Institute of Electrical Engineering, Chinese Academy of Sciences), and Dr Shilin Wu (Tsinghua University) for their support.

  • [1]
    Hannan M A et al 2020 IEEE Access 8 19113 doi: 10.1109/ACCESS.2020.2967345
    [2]
    Mejia-Ruiz G E et al 2022 IEEE Trans. Ind. Electron. 69 8949 doi: 10.1109/TIE.2021.3116557
    [3]
    Dong S L et al 2020 IET Power Electron. 13 2342 doi: 10.1049/iet-pel.2019.1184
    [4]
    Li S T 2020 High Volt. 5 122 doi: 10.1049/hve.2020.0021
    [5]
    Wu S C et al 2022 Phys. Scr. 97 072001 doi: 10.1088/1402-4896/ac6f25
    [6]
    Du B X, Liang H C and Li J 2019 IEEE Trans. Dielectr. Electr. Insul. 26 801 doi: 10.1109/TDEI.2018.007761
    [7]
    Li C Y et al 2022 IEEE Electr. Insul. Mag. 38 6 doi: 10.1109/MEI.2022.9757853
    [8]
    Zhang P H et al 2021 Nanotechnology 32 125703 doi: 10.1088/1361-6528/abcfe9
    [9]
    Li S T et al 2013 IEEE Trans. Dielectr. Electr. Insul. 20 1934 doi: 10.1109/TDEI.2013.6633727
    [10]
    Kong F et al 2020 Appl. Surf. Sci. 527 146826 doi: 10.1016/j.apsusc.2020.146826
    [11]
    Mao J L et al 2021 Langmuir 37 3903 doi: 10.1021/acs.langmuir.1c00118
    [12]
    Yuan M Q et al 2021 Nanotechnology 32 262001 doi: 10.1088/1361-6528/abe9e3
    [13]
    Ren C Y et al 2020 Plasma Sci. Technol. 22 044002 doi: 10.1088/2058-6272/ab580d
    [14]
    Liu L et al 2019 IEEE Trans. Dielectr. Electr. Insul. 26 1588 doi: 10.1109/TDEI.2019.008194
    [15]
    Zhang B Y et al 2019 Nanoscale 11 18046 doi: 10.1039/C9NR06827A
    [16]
    Xue J Y et al 2019 J. Phys. D: Appl. Phys. 52 405502 doi: 10.1088/1361-6463/ab3001
    [17]
    Du B X, Liang H C and Li J 2019 IEEE Trans. Dielectr. Electr. Insul. 26 706 doi: 10.1109/TDEI.2019.8726015
    [18]
    Zhu M X et al 2021 Chem. Eng. J. 404 126476 doi: 10.1016/j.cej.2020.126476
    [19]
    Yang X et al 2022 Appl. Surf. Sci. 592 153221 doi: 10.1016/j.apsusc.2022.153221
    [20]
    Zhang Z S et al 2021 Nanotechnology 32 122001 doi: 10.1088/1361-6528/abccea
    [21]
    Zhang P H et al 2022 High Volt. accepted doi: 10.1049/hve2.12261
    [22]
    Park Y T et al 2015 Adv. Funct. Mater. 25 575 doi: 10.1002/adfm.201402553
    [23]
    Yuan C et al 2020 Nat. Commun. 11 3919 doi: 10.1038/s41467-020-17760-x
    [24]
    Wang R et al 2022 Energy Stor. Mater. 49 339 doi: 10.1016/j.ensm.2022.04.028
    [25]
    Wang C et al 2020 Appl. Surf. Sci. 505 144432 doi: 10.1016/j.apsusc.2019.144432
    [26]
    Kango S et al 2013 Prog. Polym. Sci. 38 1232 doi: 10.1016/j.progpolymsci.2013.02.003
    [27]
    Teoh G H et al 2021 J. Membr. Sci. Res. 7 29
    [28]
    Jarvid M et al 2015 J. Mater. Chem. A 3 7273 doi: 10.1039/C4TA04956J
    [29]
    Shao T et al 2018 High Volt. 3 14 doi: 10.1049/hve.2016.0014
    [30]
    Zhang S et al 2022 Green Chem. 24 1534 doi: 10.1039/D1GC03859A
    [31]
    Zhang S et al 2022 High Volt. 7 718 doi: 10.1049/hve2.12201
    [32]
    Zhang P H et al 2020 Surf. Coat. Technol. 387 125511 doi: 10.1016/j.surfcoat.2020.125511
    [33]
    Rohr N et al 2020 J. Clin. Med. 9 2776 doi: 10.3390/jcm9092776
    [34]
    Friedrich J 2011 Plasma Proc. Polym. 8 783 doi: 10.1002/ppap.201100038
    [35]
    Wu F Y et al 2023 IEEE Trans. Ind. Electron. 70 2431 doi: 10.1109/TIE.2022.3165285
    [36]
    Takuma T, Yashima M and Kawamoto T 1998 IEEE Trans. Dielectr. Electr. Insul. 5 497 doi: 10.1109/94.708264
    [37]
    Zhang B Y, Qi Z and Zhang G X 2017 IEEE Trans. Dielectr. Electr. Insul. 24 1229 doi: 10.1109/TDEI.2017.006067
    [38]
    Liu Y Q et al 2019 Plasma Sci. Technol. 21 055501 doi: 10.1088/2058-6272/aafdf7
    [39]
    Xie Q et al 2018 Plasma Sci. Technol. 20 025504 doi: 10.1088/2058-6272/aa97d0
    [40]
    Yu H J et al 2017 Adv. Mater. 29 1605148 doi: 10.1002/adma.201605148
    [41]
    Zou Y et al 2018 Org. Electron. 56 125 doi: 10.1016/j.orgel.2018.02.008
    [42]
    Hae K J et al 2010 Europhysics Let. 92 37005 doi: 10.1209/0295-5075/92/37005
    [43]
    Markus T Z et al 2009 J. Phys. Chem. C 113 14200 doi: 10.1021/jp9041167
    [44]
    Carlson B et al 2008 J. Phys. Chem. C 112 8419 doi: 10.1021/jp7113434
    [45]
    Li X B et al 2021 Nano Energy 81 105671 doi: 10.1016/j.nanoen.2020.105671
    [46]
    Feng Q K et al 2022 Chem. Rev. 122 3820 doi: 10.1021/acs.chemrev.1c00793
    [47]
    Zhang B Y et al 2021 Adv. Mater. 33 2101374 doi: 10.1002/adma.202101374
    [48]
    Wu S L et al 2020 High Volt. 5 15 doi: 10.1049/hve.2019.0144
    [49]
    Zhou Y et al 2018 Adv. Mater. 30 1805672 doi: 10.1002/adma.201805672
    [50]
    Cheng S et al 2021 Energy Storage Mater. 42 445 doi: 10.1016/j.ensm.2021.07.018
    [51]
    Adrian L et al 2021 J. Electrochem. Soc. 168 056504 doi: 10.1149/1945-7111/abfb37
    [52]
    Zhang C et al 2019 Surf. Coat. Technol. 362 1 doi: 10.1016/j.surfcoat.2019.01.081
    [53]
    Xie Q et al 2016 Plasma Sci. Technol. 18 325 doi: 10.1088/1009-0630/18/3/18
    [54]
    Pan C et al 2017 Plasma Sci. Technol. 19 075503 doi: 10.1088/2058-6272/aa6436
  • Related Articles

    [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.
  • Cited by

    Periodical cited type(4)

    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

    Other cited types(0)

Catalog

    Figures(6)  /  Tables(1)

    Article views (65) PDF downloads (116) Cited by(4)

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return