Processing math: 0%
Advanced Search+
Qi ZHONG, Yong XIAO. Solving Poisson equation with slowing-down equilibrium distribution for global gyrokinetic simulation[J]. Plasma Science and Technology, 2023, 25(2): 025101. DOI: 10.1088/2058-6272/ac89cf
Citation: Qi ZHONG, Yong XIAO. Solving Poisson equation with slowing-down equilibrium distribution for global gyrokinetic simulation[J]. Plasma Science and Technology, 2023, 25(2): 025101. DOI: 10.1088/2058-6272/ac89cf

Solving Poisson equation with slowing-down equilibrium distribution for global gyrokinetic simulation

More Information
  • Corresponding author:

    Yong XIAO, E-mail: yxiao@zju.edu.cn

  • Received Date: March 08, 2022
  • Revised Date: August 02, 2022
  • Accepted Date: August 14, 2022
  • Available Online: December 05, 2023
  • Published Date: January 05, 2023
  • Fusion-born alpha particles in burning plasmas are usually regarded as have a slowing-down distribution, which differs significantly from the Maxwellian distribution of thermal particles in velocity space. A generalized multi-point average method has been developed for gyrokinetic Poisson equation with slowing-down equilibrium distribution using optimization in Fourier space. Its accuracy is verified in both long and short wavelength limits. The influence of changing equilibrium distribution from Maxwellian to slowing-down on gyrokinetic Poisson equation is analyzed to illustrate the significance of the new method. The effect of critical speed in the slowing-down distribution on the field solver is also presented. This method forms an important basis for global gyrokinetic simulation of low-frequency drift Alfvénic turbulence in burning plasmas.

  • In gyrokinetic particle simulation, the difference between particle distribution and gyro-center distribution leads to the double gyro-average of potential field, which is manifested in polarization density [1, 2]. The polarization density, which is essential for the gyrokinetic Poisson equation, depends critically on the phase structure of the equilibrium particle distribution, especially for short wavelength modes. Global gyrokinetic simulations are crucial for studying many important physics issues in magnetic fusion plasmas, such as turbulent transport scaling and turbulence spreading [37]. With the advent of burning plasmas, alpha particles inevitably excite Alfvénic turbulence with a slowing-down distribution via electron–alpha collisions [814]. A more accurate global gyrokinetic Poisson solver with slowing-down background distribution is desirable for simulating alpha particle physics in burning plasmas and neutral beam injection heating scenarios.

    The multi-point average method [2, 15] is a key invention that enables global simulation of microturbulence, which has been widely used in many global gyrokinetic particle-in-cell codes [16, 17]. However, these global gyrokinetic particle simulations are based on Maxwellian background distribution. Some researchers have recently used so-called equivalent Maxwellian distribution, whose temperature or second order velocity moment stays the same as the slowing-down distribution [18], to simulate the turbulent transport of alpha particles, which may be valid for a number of physics scenarios that depend weakly on the phase space structure of the equilibrium distribution. Some other works [19], though correctly considering velocity space derivatives for alpha particles, have used the local flux-tube approximation to avoid tackling the global spatial dependence of the gyrokinetic Poisson equation.

    In this work, a novel global method based on the multi-point average [2, 15] to solve the gyrokinetic Poisson equation is developed to include slowing-down equilibrium distribution for alpha particles, and the accuracy of this new method is verified in the long and short wavelength limits. This method is useful for global gyrokinetic simulation to accurately investigate alpha particle physics in burning plasmas with the advent of International Thermonuclear Experimental Reactor (ITER) operation when the alpha particles accumulate to a non-negligible portion of the ion species.

    The rest of this article is structured as follows: section 2 describes how to derive polarization density in gyrokinetics with slowing-down particle distribution; section 3 shows our numerical scheme to solve the gyrokinetic Poisson equation with the slowing-down equilibrium distribution; numerical verification is provided in section 4 for the accuracy of this new scheme; and the numerical results are summarized in section 5.

    The gyrokinetic equation describes the time evolution of the gyrocenter distribution function ˉf(X,v|| in 5D phase space [1, 20, 21], where X is the spatial gyrocenter coordinates, v|| is parallel velocity and μ is the magnetic moment with μ=mv2/2B. Suppose that the gyrocenter distribution function f¯ can be decomposed into an equilibrium component f¯0 and a perturbed component δf¯, i.e. f¯=f¯0+δf¯, where the overbar denotes a function of the gyrocenter coordinates. Following the usual gyrokinetic ordering, the perturbed distribution function δf¯ is determined by the perturbed gyrokinetic equation [2224]:

    \begin{aligned} & \left(\frac{\partial}{\partial t}+\dot{\boldsymbol{X}} \cdot \nabla+\dot{v}_{\|} \frac{\partial}{\partial v_{\|}}\right) \delta \overline{f_s}=-\boldsymbol{v}_{\mathrm{E}} \cdot \nabla f_{0 s} \\ & \quad+\frac{q_s \boldsymbol{B}^{\star}}{m_s B_0} \cdot\left(\nabla\left\langle\delta \phi_{\mathrm{gc}}\right\rangle+\boldsymbol{b} \partial_t\left\langle\delta A_{\| \mathrm{gc}}\right\rangle\right) \frac{\partial}{\partial v_{\|}} f_{0 s}, \end{aligned} (1)

    where the subscript s=i,e,α stands for particle species, i.e. ion, electron and alpha particle respectively; B0 is equilibrium magnetic field with magnitude B0 and direction b=B0/B0. B=B0+msv||×b/qs+δB; qs is particle charge; and ms is particle mass.

    The equations of motion for the gyrocenter are given by

    Ẋ=v||b+vd+vE, (2)
    \dot{v}_{\|}=-\frac{\boldsymbol{B}^{\star}}{m_s B_0} \cdot\left(\mu_s \nabla B_0+q_s \nabla\left\langle\delta \phi_{\mathrm{gc}}\right\rangle+q_s \boldsymbol{b} \partial_t\left\langle\delta A_{\| \mathrm{gc}}\right\rangle\right), (3)

    where is magnetic drift for the guiding centers with representing magnetic field curvature, \boldsymbol{v}_{\mathrm{E}}=-\nabla\left\langle\delta \phi_{\mathrm{gc}}\right\rangle / B_0 is the gyro-averaged drift, and \langle\ldots\rangle represents gyrophase average. The gyrocenter perturbed electromagnetic fields are given by pushing forward transform from particle coordinates to guiding center coordinates

    δϕgcX;μ,ζ,texpρ·δϕ, (4)
    δA||gcX;μ,ζ,texpρ·δA||, (5)

    where the gyroradius ρ is defined as ρ=b×v/Ωs with the gyrofrequency Ωs=qsB/m and ζ is the gyrophase. The perturbed vector potential can be calculated by the parallel Ampère's law:

    2δA||=-μ0s=i,e,αqsδu||s (6)

    where the perturbed parallel fluid velocity δu||s can be calculated by δu||s=d3vv||exp-ρ·δf¯s. The perturbed electrostatic potential δϕ is calculated by the quasi-neutrality condition: qiδni+qαδnα=eδne. Due to the finite Larmor radius, the perturbed response of ions or alpha particles can be divided into two parts: the contribution from perturbed gyrocenter distribution and the contribution from particle polarization due to fluctuating electric fields:

    -j=i,αqjδnpol,j=qiδni,gc+qαδnα,gc-eδne, (7)

    where the guiding center density δns,gc is defined as δni,gc=d3vexp-ρ·δf¯s. In many cases, the adiabatic response can be conveniently assumed for electrons due to their fast parallel motion, i.e. δne=en0δϕ/Te with Te the electron temperature. However, the perturbed density δne actually contains contributions not only from adiabatic electrons but also from non-adiabatic electrons, i.e. δne=en0δϕTe+δneNA. We also note that simply employing equations (1), (6), (7) does not provide a stable numerical algorithm for nonlinear electromagnetic simulation. In [23], a fluid–kinetic hybrid electron simulation model is invented to resolve this difficulty.

    In the preceding equation, the polarization density for thermal ions is given by [1]

    δnpol,i=-qin0iTiδϕ-δϕ~,

    where Ti is the ion temperature and the double gyro-average of electric potential is defined as

    \widetilde{\delta \phi}=\int \mathrm{d}^3 v \exp (-\boldsymbol{\rho} \cdot \nabla)\left\langle\delta \phi_{\mathrm{gc}}\right\rangle f_{\mathrm{M}}\left(\mu, v_{\|}, \boldsymbol{X}\right)

    where Maxwellian background distribution fM is assumed for the thermal ions, and δϕ~ has a complicated form in real space but a neat form in Fourier space:

    δϕ~kΓ0kρiδϕkexpik·x (8)

    with the function Γ0=I0be-b, b=k2ρi2, where ρi=Ti/mi and I0 is a modified Bessel function of the first kind. In the gyrokinetic particle simulation, δϕ~ can be calculated by a multi-point average method in the real space, which enables global gyrokinetic simulation [15, 25].

    However, the calculation of polarization density for alpha particles with the slowing-down background distribution remains to be unsolved for gyrokinetic particles simulation. The slowing-down distribution fsld is the steady-state solution to the collisional scattering for an isotropic particle source with a large birth speed v0, i.e. the alpha particle source produced by the thermal nuclear fusion, where the birth energy of alpha particle 12mαv02=3.5MeV for typical D–T fusion. It is discovered that fsld can be explicitly expressed as [26, 27]:

    fsldv=n0α4πI1Hv0-vv3+vc3, (9)

    where vc3πme4neiqα2n0αe2mα1/3vth,e is the critical speed at which the electron drag is comparable to the thermal ion drag, and I1=13ln1+v03/vc3 is an auxiliary function for the purpose of normalization. The alpha temperature Tα can be defined as the second velocity moment of the slowing-down distribution function, which is similar to that of the Maxwellian:

    32n0αTα=d3v12mαv2fsldv12n0αmαvc2I2I1 (10)

    with I2=v022vc2-16π3-23arctan1-2v0/vc3-ln1+v0/vc31-v03/vc3. We note that for the fusion-born alpha particles in a 10 keV 50%–50% D-T plasma, vc/v00.3, Tα1.28msvc2. The integrations I1,I2 have also been defined in [18].

    Next, we show how to derive the form of polarization density for the alpha particles, i.e. δnpol,α in equation (7), based on a gyrokinetic approach. Employing the Lie transform method [28], we can calculate the alpha density nα by integrating over the velocity space on the distribution function, generated by pulling back the gyrocenter alpha distribution function fα¯=fsld+δf¯α,

    n_{0 \alpha}+\delta n_\alpha=\int \mathrm{d}^3 v \exp (-\rho \cdot \nabla)\left(1+\frac{q_\alpha}{B} \delta \tilde{\phi}_{\mathrm{gc}} \frac{\partial}{\partial \mu}\right)\left(f_{\mathrm{sld}}+\delta \bar{f}_\alpha\right)

    where \delta \tilde{\phi}_{\mathrm{gc}} \equiv \delta \phi_{\mathrm{gc}}-\left\langle\delta \phi_{\mathrm{gc}}\right\rangle is the gyrophase-dependent part of As is discussed, the perturbed alpha density can be separated into a perturbed gyrocenter center density and a polarization density, i.e. and the polarization density is associated with the fluctuating electric field through

    δnpol,α=d3vexp-ρ·qαBδϕ~gcμfsld. (11)

    We note that in this expression, the nonlinear term in δnα,pol is ignored for simplicity.

    To solve the gyrokinetic Poisson equation, a proper numerical algorithm is needed to deal with the guiding center transformation exp-ρ· and gyrophase average in equation (11). In the local gyrokinetic simulation, this operation is carried out in Fourier space, and the associated phase angle and velocity space integration can be done analytically, which greatly simplifies numerical calculation. For global gyrokinetic simulation, the four-point or multi-point gyro-average method has been invented to calculate the ion polarization density in real space when the equilibrium distribution f0i is Maxwellian [15]. Here we modify the original four-point average method to accommodate a slowing-down equilibrium distribution in order to investigate the self-consistent turbulent transport physics involving alpha particles. Considering the Fourier representation of the perturbed potential δϕ=kδϕkexpik·x in equation (11), then we can obtain:

    δnpol,α=-kd3v1-exp-ik·ρJ0kρ×δϕkλfsldexpik·x=-kcα-Γαkρcn0αTαqαδϕkexpik·x, (12)

    where and J_0= J_0\left(k_{\perp} \rho_{\mathrm{i}}\right)=\langle\exp (\mathrm{i} \boldsymbol{k} \cdot \boldsymbol{\rho})\rangle is the zeroth-order Bessel function, and is defined as

    Γαkρcd3vexp-ik·ρJ0kρλfsld=dvvJ02kv/Ωif~sldv, (13)

    which can be considered as the expectation of λ weighted by the equilibrium distribution fsld after double gyro-averaging due to the back and forth transformation between particle position and gyrocenter position, where f~sldv=dv||λfsld. Unlike the Maxwellian case, Γα does not have a simple analytic expression in Fourier space and has to be evaluated numerically. In principle, one can solve the quasi-neutrality equation, i.e. equation (7), using equations (12) and (13) in Fourier space. Figure 1 shows the polarization term cα-Γαkρc in Fourier space between Maxwellian (red solid line) and slowing-down (blue dashed line) with identical temperature but different v0/vc. From bottom to top, we have v0/vc ranging from 2 to 5 with a 0.5 step size. The polarization calculated from slowing-down and that from the equivalent Maxwellian are the same for long wavelengths but differ while kρc1.

    Figure  1.  Polarization comparison between slowing-down distribution and equivalent Maxwellian in k-space. The blue dashed line is for the slowing-down distribution and the red solid line is for the equivalent Maxwellian distribution. The lines from bottom to top correspond to various ratios of v0/vc ranging from 2 to 5 with a step size of 0.5.

    With these newly defined functions, the gyrokinetic Poisson equation can be written in a dimensionless form

    n0in0qi2Tiδϕ-δϕ~+n0αn0qα2Tαcαδϕ-δϕα~=qiδni,gcn0+qαδnα,gcn0-eδne,gcn0, (14)

    where δϕα~ is defined by the following Fourier representation:

    δϕα~kΓαkρiδϕkexpik·x. (15)

    For the typical D–T plasmas parameters mentioned before in the definition in temperature of slowing-down distribution, the polarization density of alpha particles is enhanced by the mass ratio mα/mi for long wavelengths if the equilibrium density ratio is given. However, the ratio nα,pol/ni,pol goes down as k become larger, as shown in figure 2.

    Figure  2.  Polarization density ratio between alpha particles and thermal fuel ions varying in k space, normalized by their equilibrium densities respectively.

    This Fourier representation is neat in form but it mixes up the configuration space and velocity space dependences through the J0 term. In reality, the background magnetic field and perpendicular temperature can vary in real space, and then Γα will gain global spatial dependence. Besides, the Fourier transform approach makes it more difficult to deal with realistic tokamak geometry, where no periodicity exists in the radial direction and on many occasions the global effects have to be considered seriously [3]. For the Maxwellian background distribution, the multi-point/four-point gyro-average method has been developed to solve this gyrokinetic Poisson equation in real space [1, 2]. Here we modify this method by including the slowing-down background distribution fsld as the equilibrium distribution in the gyrokinetic Poisson equation.

    The crucial part of implementing this multi-point average method for a gyrokinetic Poisson solver is to represent δϕα~ in equation (15) by the values of δϕ at various field points in real space. By numerical interpolation, we note that δϕ~ can be expressed as a linear combination of the δϕ values on a number of nearby grid points and consequently equation (14) is transformed into a discrete matrix form such as A·x=b, which can then be solved by many known matrix inversion algorithms, such as the Krylov subspace solver, provided by an existing parallel library like PETSc [29].

    Starting from the integral form of δϕα~ instead of the Fourier form in equation (15), one finds that

    \widetilde{\delta \phi_\alpha}=\int_0^{\infty} \mathrm{d} v_{\perp} v_{\perp}\left\langle\exp (-\rho \cdot \nabla)\left\langle\delta \phi_{\mathrm{gc}}\right\rangle\right\rangle \tilde{f}_{\mathrm{sld}}\left(v_{\perp}\right). (16)

    To calculate at a grid point for a specific one needs to evaluate the gyro-averaged function \left\langle\exp (-\rho \cdot \nabla)\left\langle\delta \phi_{\mathrm{gc}}\right\rangle\right\rangle, , which is the average value of \left\langle\delta \phi_{\mathrm{gc}}\right\rangle on a ring with radius around as shown by the dotted circle in figure 3. The gyro-averaged quantity \left\langle\delta \phi_{\mathrm{gc}}\right\rangle can also be calculated by this ring average method, e.g., the value of \left\langle\delta \phi_{\mathrm{gc}}\right\rangle at the black triangle in figure 3 can be calculated by the average value on a solid circle. It is not necessary to actually integrate numerically along the whole ring to compute the gyrophase average, which would make the gyro-average process rather time-consuming and expensive. According to [1, 2], a selection of four points uniformly distributed on the ring (four-point average method) is sufficient to compute the gyro-average for wavelengths up to Thus, nine neighboring points are required to compute on the grid point, as shown by the eight red points and the central blue diamond in figure 3. This scheme can be modified to adopt 4N (N = 1, 2 ···) points on the ring straightforwardly. In more general geometry, these points required for the gyro-average computation may not lie exactly on the grids, but their values can be acquired by a linear interpolation of the nearby grid points. Finally, a few rings are summed with different values with the weight function and the relationship between and on each grid point is found.

    Figure  3.  Scheme for calculating δϕ~ at grid point i,j.

    The remaining issue in evaluating equation (16) is how to discretize the v integral with the weight function f~sldv. Here we approximate the integral by a weighted summation by choosing a few sampling grid points along the v coordinate. From the definition of δϕ~, one can tell that it is equivalent to the approximate equation (13) by:

    \begin{aligned} \Gamma_\alpha\left(k_{\perp} \rho_{\mathrm{c}}\right) & =\int \mathrm{d} v_{\perp} J_0^2\left(k_{\perp} v_{\perp} / \Omega_{\mathrm{i}}\right) \tilde{f}_{\mathrm{sld}}\left(v_{\perp}\right) \\ & ≊ \sum\limits_j c_j J_0^2\left(k_{\perp} \rho_{\mathrm{c}} \frac{v_{\perp j}}{v_{\mathrm{c}}}\right), \end{aligned} (17)

    where cj are the summing weights due to f~v and vj are the sampling grid points. The value pairs of cj,vj are chosen by minimizing the following error function:

    ϵ=0aΓαx-jcjJ02xvjvc2dx (18)

    Here a is the maximum value of kρc that we are interested in. Since low-frequency micro-turbulence usually peaks around kρc<1, it is required that this approximation has a better accuracy for long wavelengths or kρc1. Considering the Taylor expansion for J0x and Γαx around x~0, one finds that J0x=1-x2/4+Ox4 and Γαx=c0-Tαmvc2x2+Ox4. Let the first two terms equal each other:

    jcj=c0=I2I33I12, (19)
    jcjvj2vc2=2Tαmvc2=2I23I1. (20)

    These two constraints are then used to reduce the degrees of freedom. In order to minimize ϵ with respect to cj,vj, we use the Nelder–Mead method [30], which is a gradient-free iterative optimization algorithm. I1,2,3 are functions of vc/v0, which is chosen to be 0.3 here to show the numerical result of cj,vj. In the one-velocity-node case, we find that c=1.226 with the velocity node v/vc=1.443 and the relative error is 3.6% for kρc<0.5. When using two velocity nodes, we find that c1=0.9347 and c2=0.2910 with the velocity nodes v1/vc=0.8778 and v2/vc=2.510, and the relative error is about 3.6% for kρc<1.5. In the three-velocity-node case, we find that c1,c2,c3=0.1186,0.3881,0.7190 with v1/vc,v2/vc,v3/vc=0.7016,1.716,2.984, and the relative error is only 0.46% for kρc<2. The three-velocity-node approximation is compared with the exact value from direct numerical integration, as shown in figure 4. Satisfactory accuracy is achieved with a relative error less than 0.46% for kρc<2, which is sufficient to include the most interesting finite Larmor radius effects due to the slowing-down alpha particles. We find that using more velocity nodes can achieve better accuracy for larger kρi, e.g., using four velocity nodes can make the Γα function from the four-point-average method sufficiently accurate up to kρi~4. Notice that vc can be varied at equilibrium scale and the weights ci and velocity nodes vi will change as vc/v0 changes. We can compute a series of weights and velocity nodes of different vc/v0, then generate a spline with these discrete values. When constructing the matrix corresponding to Γα, the local values of weights and velocity nodes can be obtained from the pre-calculated spline. In this way, the four-point average method can capture the non-uniformity of the equilibrium quantities. As a contrast, this non-uniformity affects only ρ but not the coefficients cj,vj for the Maxwellian case [15].

    Figure  4.  Exact Γα function (blue solid line) and its numerical approximations vary with perpendicular wavelength kρc: 4-point average method with three velocity nodes in the integration (red dotted line), and Padé approximation (black dashed line).

    We also test the widely used Padé approximation for the thermal ions, and find that it can introduce a 10% relative error near kρc~1.5 comparing to the exact solution. Figure 5 shows the comparison between the four-point average method and the Padé approximation with the following form:

    Γαkρc=cα1+1cαρc2k2. (21)
    Figure  5.  Comparison of analytic expression and numeric solutions using four-point average approximation and Padé approximation for the gyrokinetic Poisson equation in the long wavelength limit with krρi=0.11 and m=6.

    We note that the Padé approximation with the Maxwellian distribution has a similar form with cα=1. In the long wavelength limit, these approximations are both very close to the exact value, as shown in figure 5, and it can be further verified by the numerical benchmarks shown in the next section.

    Although the method demonstrated here is applied to the isotropic background distribution, it can also be generalized for more complex anisotropic background distributions. To obtain the Γα function in equation (13) for the polarization density term, we just need to perform /μ and integrate it over velocity space with the anisotropic distribution function, which can be performed even for the numerical distribution functions. This will be carried out in our future work.

    To verify our global algorithm for the slowing-down background distribution, we shall solve the gyrokinetic Poisson equation without electron response in a large-aspect-ratio toroidal geometry tokamak with circular cross section as a sample problem. We have implemented this multi-point method for the gyrokinetic Poisson solver in Gyrokinetic Toroidal Code (GTC) by loading slowing-down marker distribution, modifying weighting coefficients and using its embedded PETSc Library for the gyrokinetic Poisson solver. The simulation domain is chosen to be ra0,a1 on the poloidal plane, with a0=0.142a, a1=0.938a and ρα/a2.51×10-2. In the GTC, the accuracy of gyrokinetic Poisson solver has been verified for thermal ions and electrons [31]. Thus, we can verify the accuracy of gyrokinetic Poisson solver for alpha particles by ignoring contributions from the thermal ion and electron in equation (14). In the long wavelength limit equation (14) can be reduced to the following form using the Taylor expansion of Γα:

    qαmαΩα22δϕ=-δnα,gcn0. (22)

    In the large safety factor limit, the toroidal effect can be ignored and 2=1rrrr+1r22θ2 in polar coordinates on the poloidal plane since the difference between the perpendicular plane and poloidal plane can be ignored in the large-aspect-ratio limit. After these simplifications, equation (22) is just a normal Poisson equation and we can choose δnα,gc to be the eigenfunction of the Laplacian operator to ensure an analytic solution. Let δni,gc/ni=Jmk0r-Ymk0rJmk0a1/Ymk0a1cosmθ in which k0 satisfies Jmk0a0Ymk0a1-Ymk0a0Jmk0a1=0. Here Jm and Ym are mth-order Bessel functions of the first and second kinds, respectively. Then the solution of this Poisson equation is just δϕ=k0-2Jmk0r-Ymk0rJmk0a1/Ymk0a1cosmθ with zero boundary condition on r=a0,a1. With m=6 as an example, the comparison between the analytic solution and numeric solution along the line θ=0 is shown in figure 5. The relative differences are very small at 0.6% between the analytic and four-point average and 1.3% between the analytic and Padé approximation. More generally, the poloidal cross section contour for the solution is shown in figure 6. The difference is negligibly small between the numerical solution from the four-point average and that from the Padé approximation, as shown in figure 6(c). Thus, in the long wavelength limit, our four-point average method works perfectly for the slowing-down equilibrium distribution.

    Figure  6.  2D poloidal contours for solutions to different operators to the gyrokinetic Poisson equation in the long wavelength limit: (a) four-point average operator, (b) Padé approximation operator, where krρi=0.11 and m=6. The differences between these two solutions are shown in (c).

    In order to simulate short wavelength modes, we need to verify the validity of our algorithm in the short wavelength limit. We solve equation (14) with only slowing-down alpha particles in a natural unit Tα=Te=B0=qα=e=1, with Tα,Te and B0 being the quantity at the magnetic axis. Using eigenfunctions of the Laplacian operator, δϕr,θ can be expressed as a series:

    δϕr,θ=m,iδϕm,iJmkm,ir-Jmkm,ia0Ymkm,ia0Ymkm,ireimθ (23)

    Here, is the ith positive value satisfying and the boundary condition is fulfilled. The gyro-averaged potential \left\langle\delta \phi_{\mathrm{gc}}\right\rangle can now be expanded as

    \begin{aligned} & \left\langle\delta \phi_{\mathrm{gc}}\right\rangle(r, \theta)=\sum\limits_{m, i} J_0\left(k_{m, i} \rho\right) \delta \phi_{m, i} \\ & \quad \times\left(J_m\left(k_{m, i} r\right)-\frac{J_m\left(k_{m, i} a_0\right)}{Y_m\left(k_{m, i} a_0\right)} Y_m\left(k_{m, i} r\right)\right) \mathrm{e}^{\mathrm{i} m \theta} \end{aligned} (24)

    Then, following the same procedure of how equation (12) is derived, the double gyro-averaged term is

    δϕ~αr,θ=m,iΓαρckm,iδϕm,i×Jmkm,ir-Jmkm,ia0Ymkm,ia0Ymkm,ireimθ (25)

    In order to compare this analytic solution with that from the four-point average method, we keep the magnetic field and temperature profile constant and thus ρc is spatially uniform. If the density is given by a summation of the aforementioned eigenfunctions, then equation (14) can be solved analytically. Here we sum up three eigenfunctions with m=6, and different radial wavelengths at krρc=1,1.5,2, and corresponding amplitudes to 1,0.5 and 0.2 respectively. Figures 7 and 8 show the 1D and 2D plots for the numerical solutions from the four-point average method and Padé approximation, respectively. The globally averaged differences between the numerical methods and the analytic solution are about 1% and can be ignored. Figure 7 also shows the solution from the original four-point average method with an equivalent Maxwellian with the same temperature as the slowing-down distribution, i.e. the orange dotted-and-dashed line. This numerical solution deviates from the analytic solution with an overall error of more than 6%. The amplitude of the solution using the four-point average method is slightly larger, which can be ascribed to the fact that the operator of the four-point average is larger than the Padé approximation in k space, as shown in figure 4. Though not so precise in k space, the Padé approximation still gives a solution as good as that from the four-point average method in this case. This suggests that we may not need to be very accurate at large kρα since the polarization density is nearly identical to the adiabatic response in the short wavelength limit, regardless of the particle's velocity distribution.

    Figure  7.  Comparison of analytic solution (blue solid line) and numeric solutions using the new four-point average method with slowing-down equilibrium (red X marker), original four-point average method with equivalent Maxwellian equilibrium (orange dotted-and-dashed line) and Padé approximation (black long dashed line) for the gyrokinetic Poisson equation in the short wavelength limit.
    Figure  8.  2D contour of the solution to gyrokinetic Poisson equation in the short wavelength limit on the poloidal plane. The numerical methods used in solving the Poisson equation are the (a) four-point average method and (b) Padé approximation. The difference between them is shown in (c).

    Finally, we investigate how the fraction of energetic particle affects the gyrokinetic Poisson solver. Here we assume that the main ions are a 50%–50% mixture of D and T with a temperature of 10 keV, and the energetic particles are the fusion-born alpha particles with a slowing-down distribution. The density term is the same as that in the above short wavelength verification where krρc~1 and in this case krρi1. The solutions are shown in figure 9 with different energetic particle fractions from 5% to 20%. The difference could be significant when nα/ne increases above 5%, as demonstrated in the figure.

    Figure  9.  Solutions to an arbitrary density source with a dominant wavelength krρc~1 with different energetic alpha particle fraction from 5% to 20%.

    A real space gyrokinetic Poisson solver for slowing-down equilibrium distribution has been developed based on the multi-point average method [2, 15] and verified for its accuracy in the long and short wavelength limits. The discovery process for this method is shown in detail and it can be further modified to accommodate more complicated equilibrium distributions. This method can be incorporated into the global gyrokinetic particle simulation to study the crucial alpha particle physics in the burning plasmas, i.e. to simulate the drift Alfvénic turbulence accurately in the presence of slowing-down alpha particle distribution.

    This work is supported by the National Magnetic Confinement Fusion Program of China (No. 2015GB110000), and by National Natural Science Foundation of China (No. 11975201).

  • [1]
    Lee W W 1983 Phys. Fluids 26 556 doi: 10.1063/1.864140
    [2]
    Lee W W 1987 J. Comput. Phys. 72 243 doi: 10.1016/0021-9991(87)90080-5
    [3]
    Lin Z et al 2002 Phys. Rev. Lett. 88 195004 doi: 10.1103/PhysRevLett.88.195004
    [4]
    Wang W X et al 2006 Phys. Plasmas 13 092505 doi: 10.1063/1.2338775
    [5]
    Lang J Y, Chen Y and Parker S E 2007 Phys. Plasmas 14 082315 doi: 10.1063/1.2771141
    [6]
    Qi L et al 2019 Nucl. Fusion 59 026013 doi: 10.1088/1741-4326/aaf5fd
    [7]
    Hahm T S et al 2004 Plasma Phys. Control. Fusion 46 A323 doi: 10.1088/0741-3335/46/5A/036
    [8]
    Wang X, Zonca F and Chen L 2010 Plasma Phys. Control. Fusion 52 115005 doi: 10.1088/0741-3335/52/11/115005
    [9]
    Bierwage A, Chen L and Zonca F 2010 Plasma Phys. Control. Fusion 52 015005 doi: 10.1088/0741-3335/52/1/015005
    [10]
    Chen L and Zonca F 2016 Rev. Mod. Phys. 88 015008 doi: 10.1103/RevModPhys.88.015008
    [11]
    Heidbrink W W 2008 Phys. Plasmas 15 055501 doi: 10.1063/1.2838239
    [12]
    Heidbrink W W et al 2016 Nucl. Fusion 56 056005 doi: 10.1088/0029-5515/56/5/056005
    [13]
    Shen W et al 2017 Nucl. Fusion 57 116035 doi: 10.1088/1741-4326/aa7f9c
    [14]
    Sheng H and Waltz R E 2016 Nucl. Fusion 56 056004 doi: 10.1088/0029-5515/56/5/056004
    [15]
    Lin Z and Lee W W 1995 Phys. Rev. E 52 5646 doi: 10.1103/PhysRevE.52.5646
    [16]
    Wang W X et al 2015 Phys. Plasmas 22 102509 doi: 10.1063/1.4933216
    [17]
    Ku S, Chang C S and Diamond P H 2009 Nucl. Fusion 49 115021 doi: 10.1088/0029-5515/49/11/115021
    [18]
    Estrada-Mila C, Candy J and Waltz R E 2006 Phys. Plasmas 13 112303 doi: 10.1063/1.2364149
    [19]
    Angioni C and Peeters A G 2008 Phys. Plasmas 15 052307 doi: 10.1063/1.2913610
    [20]
    Catto P J 1978 Plasma Phys. 20 719 doi: 10.1088/0032-1028/20/7/011
    [21]
    Frieman E A and Chen L 1982 Phys. Fluids 25 502 doi: 10.1063/1.863762
    [22]
    Hahm T S, Lee W W and Brizard A 1998 Phys. Fluids 31 1940 doi: 10.1063/1.866641
    [23]
    Lin Z H and Chen L A 2001 Phys. Plasmas 8 1447 doi: 10.1063/1.1356438
    [24]
    Holod I et al 2009 Phys. Plasmas 16 122307 doi: 10.1063/1.3273070
    [25]
    Lin Z et al 1998 Science 281 1835 doi: 10.1126/science.281.5384.1835
    [26]
    Cordey J G and Core W G F 1974 Phys. Fluids 17 1626 doi: 10.1063/1.1694943
    [27]
    Gaffey J DJr 1976 J. Plasma Phys. 16 149 doi: 10.1017/S0022377800020134
    [28]
    Littlejohn R G 1982 J. Math. Phys. 23 742 doi: 10.1063/1.525429
    [29]
    Balay S et al 1997 Efficient Management of Parallelism in Object-Oriented Numerical Software Libraries (Boston, MA: Springer) 163 doi: 10.1007/978-1-4612-1986-6_8
    [30]
    Nelder J A and Mead R 1965 Comput. J. 7 308 doi: 10.1093/comjnl/7.4.308
    [31]
    Rewoldt G, Lin Z and Idomura Y 2007 Comput. Phys. Commun. 177 775 doi: 10.1016/j.cpc.2007.06.017
  • Related Articles

    [1]Xin AI, Qiuyue NIE, Zhonglin ZHANG, Peiqi CHEN, Shulei ZHENG, Changshi YAN, Guoqiang WEI. Effect of attack angle on the electromagnetic wave transmission characteristics in the hypersonic plasma sheath of a re-entry vehicle[J]. Plasma Science and Technology, 2024, 26(12): 125301. DOI: 10.1088/2058-6272/ad7786
    [2]Lu MA (马璐), Xiaodong WANG (王晓东), Jian ZHU (祝健), Shun KANG (康顺). Effect of DBD plasma excitation characteristics on turbulent separation over a hump model[J]. Plasma Science and Technology, 2018, 20(10): 105503. DOI: 10.1088/2058-6272/aacdf0
    [3]Xiangyang LIU (刘向阳), Siyu WANG (王司宇), Yang ZHOU (周阳), Zhiwen WU (武志文), Kan XIE (谢侃), Ningfei WANG (王宁飞). Thermal radiation properties of PTFE plasma[J]. Plasma Science and Technology, 2017, 19(6): 64012-064012. DOI: 10.1088/2058-6272/aa65e8
    [4]ZHANG Weiwei (张卫卫), DENG Baiquan (邓柏权), ZUO Haoyi (左浩毅), ZHENG Xianjun (曾宪俊), CAO Xiaogang (曹小岗), XUE Xiaoyan (薛晓艳), OU Wei (欧巍), CAO Zhi (曹智), GOU Fujun (芶富均). Analysis of Power Model for Linear Plasma Device[J]. Plasma Science and Technology, 2016, 18(8): 844-847. DOI: 10.1088/1009-0630/18/8/09
    [5]LIU Zhiwei (刘智惟), BAO Weimin (包为民), LI Xiaoping (李小平), SHI Lei (石磊), LIU Donglin (刘东林). Influences of Turbulent Reentry Plasma Sheath on Wave Scattering and Propagation[J]. Plasma Science and Technology, 2016, 18(6): 617-626. DOI: 10.1088/1009-0630/18/6/07
    [6]LI Mei (李美), ZHANG Junpeng (张俊鹏), HU Yang (胡杨), ZHANG Hantian (张含天), WU Yifei (吴益飞). Simulation of Fault Arc Based on Different Radiation Models in a Closed Tank[J]. Plasma Science and Technology, 2016, 18(5): 549-553. DOI: 10.1088/1009-0630/18/5/18
    [7]ZHOU Suyun (周素云), CHEN Hui (陈辉), LI Yanfang (李艳芳). The Effect of Ion Motion on Laser-Driven Plasma Wake in Capillary[J]. Plasma Science and Technology, 2016, 18(1): 86-89. DOI: 10.1088/1009-0630/18/1/15
    [8]SHI Lei (石磊), ZHAO Lei (赵蕾), YAO Bo (姚博), LI Xiaoping (李小平). Telemetry Channel Capacity Assessment for Reentry Vehicles in Plasma Sheath Environment[J]. Plasma Science and Technology, 2015, 17(12): 1006-1012. DOI: 10.1088/1009-0630/17/12/05
    [9]JI Xiang (戢翔), SONG Yuntao (宋云涛), WU Songtao(武松涛), WANG Zhibin(王志滨), SHEN Guang (沈光), LIU Xufeng (刘旭峰), CAO Lei (曹磊), ZHOU zibo (周自波), PENG Xuebing(彭学兵), WANG Chenghao(王成昊). Electromagnetic Modeling of the Passive Stabilization Loop at EAST[J]. Plasma Science and Technology, 2012, 14(9): 855-858. DOI: 10.1088/1009-0630/14/9/16
    [10]XI Yanbin (奚衍斌), LIU Yue (刘悦). FDTD Simulation on Power Absorption of Terahertz Electromagnetic Waves in Dense Plasma[J]. Plasma Science and Technology, 2012, 14(1): 5-8. DOI: 10.1088/1009-0630/14/1/02
  • Cited by

    Periodical cited type(4)

    1. Esposito, S., Scarabosio, A., Vecchi, G. et al. Non-equilibrium plasma distribution in the wake of a slender blunted-nose cone in hypersonic flight and its effect on the radar cross section. Aerospace Science and Technology, 2024. DOI:10.1016/j.ast.2024.109699
    2. Bao, Y., He, X., Su, W. et al. Study on the generation of terahertz waves in collision plasma. Physics of Plasmas, 2024, 31(9): 093302. DOI:10.1063/5.0219947
    3. Tong, J., Li, H., Xu, B. et al. Inversion of electron densities in plasma wakes of hypersonic targets. Results in Physics, 2024. DOI:10.1016/j.rinp.2024.107714
    4. Zhang, H., Li, J., Qiu, C. et al. Electromagnetic scattering characteristics of a hypersonic vehicle with a microrough surface in the millimeter wave band. AIP Advances, 2023, 13(9): 095215. DOI:10.1063/5.0160916
    1. Esposito, S., Scarabosio, A., Vecchi, G. et al. Non-equilibrium plasma distribution in the wake of a slender blunted-nose cone in hypersonic flight and its effect on the radar cross section. Aerospace Science and Technology, 2024. DOI:10.1016/j.ast.2024.109699
    2. Bao, Y., He, X., Su, W. et al. Study on the generation of terahertz waves in collision plasma. Physics of Plasmas, 2024, 31(9): 093302. DOI:10.1063/5.0219947
    3. Tong, J., Li, H., Xu, B. et al. Inversion of electron densities in plasma wakes of hypersonic targets. Results in Physics, 2024. DOI:10.1016/j.rinp.2024.107714
    4. Zhang, H., Li, J., Qiu, C. et al. Electromagnetic scattering characteristics of a hypersonic vehicle with a microrough surface in the millimeter wave band. AIP Advances, 2023, 13(9): 095215. DOI:10.1063/5.0160916

    Other cited types(0)

Catalog

    Figures(9)

    Article views (87) PDF downloads (320) Cited by(4)

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return