Brought to you by:
Paper

6D phase space collective modes in Vlasov-Maxwell system

and

Published 6 June 2022 © 2022 IOP Publishing Ltd
, , Citation H Lin and C P Liu 2022 Plasma Res. Express 4 025005 DOI 10.1088/2516-1067/ac743e

2516-1067/4/2/025005

Abstract

Plasma electromagnetic (EM) kinetic simulation faces two unavoidable difficulties. One arises from the fact that Maxwell equations determine a simultaneity relation between transverse electric field ${\vec{E}}_{{Tr}}$ and local growth rates of probability distribution function (PDF) $f$ in Vlasov-Maxwell (V-M) simulation in Eulerian approach, so does that between ${E}_{{Tr}}$ and Lagrangian particles' time-varying rates which refers to displacement of $f$-element in $6D$ phase space in V-M simulation in Lagrangian approach, and macroparticles' accelerating rates in Particle-in-Cell (PIC) simulation. These simultaneous with ${E}_{{Tr}}$ are termed as bottom objects' growth rates (BOGRs) in this work. Directly solving the BOGRs needs to diagonalize a large full matrix, and hence often be approximated. The other arises from the fact that Lagrangian particles' time-histories should uniformly converge with respect to the time-step. This severe requirement is difficult to be satisfied and hence some of Lagrangian particles' time-histories lose fidelity. We propose a strict alternative method free from two difficulties. The initial-value problem of V-M system can be interpreted by $6D$ phase space allowed deformation of an initial $f$-profile. By virtue of a more compact description of $f,$ in which a conditional probability density function (C-PDF) well reflects some macroscopic conservation laws behind the V-M system, we can interpret the initial-value problem of $f$ in terms of $6D$ standing-wave oscillation in the C-PDF. A key subtle difference between microscopic scalar field and microscopic vector field can also lead to a similar scheme of particle simulation.

PACS: 52.65.-y.

Export citation and abstract BibTeX RIS

1. Introduction

Collective modes are common conceptions in plasma fluid theory. They usually exhibit as strict mathematic solutions of fluid equations with given initial conditions and $3D$ real space boundary conditions. For example, travelling-wave type electronic density oscillation is a familiar example of collective modes. In contrast, in plasma microscopic theory, similar conceptions are nearly unexploited and people tend to resort to kinetic simulation. No matter its detailed form is in Particle-in-Cell (PIC) code [116] or Vlasov code [1739], the kinetic simulation treats both Vlasov-Maxwell/Poisson system and Newton-Maxwell system as an initial-value problem, and hence often hires one of two mainstream methods of numerical experiments on partial differential equations (PDEs): Eulerian approach and Lagrangian one. The Eulerian approach is based on a fixed numerical mesh. Compared with Lagrangian one, a drawback of the mesh-based method is its significant increase in computational cost as the problem dimension increases even though it is of the advantage of high order accuracy. Therefore, the Lagrangian approach, which is based on tracing time-histories of Lagrangian particles, receives more attention. Here, the phrase 'Lagrangian particle' refers to macroparticles in the PIC simulation and $f$-elements, where $f$ stands for probability density distribution (PDF), of $6D$ phase space in the Vlasov-Maxwell (V-M) simulation in Lagrangian approach.

However, from basic mathematics perspective, both the Lagrangian approach and the Eulerian approach face two serious difficulties when applied to electromagnetic (EM) kinetic simulation. One is raised by simultaneity relation among physics fields. As expounded latter, the simultaneity relation between transverse electric field ${\vec{E}}_{{Tr}}$ and bottom objects' growth rates (BOGRs) which refer to ${\partial }_{t}f$-values in the V-M simulation in Eulerian approach, displacement of $f$-element in the V-M simulation in Lagrangian approach and macroparticles' accelerating rate in the PIC simulation, makes the calculation on these BOGRs involving in high-dimension matrix. Here, the transverse electric field ${\vec{E}}_{{Tr}}$ satisfies ${\rm{\nabla }}\cdot {\vec{E}}_{{Tr}}=0$ but ${\rm{\nabla }}\times {\vec{E}}_{{Tr}}\ne 0,$ and is a fraction of self-consistent electric field $\vec{E}.$ Usually, the dimension of such a matrix is so high that people have to resort to approximations. The common of these approximations is to ignore the simultaneity relation among ${\vec{E}}_{{Tr}}$ and BOGRs and hence avoid the difficulty in dealing with the high-dimension matrix. Clearly, such an ignorance is equal to ignoring some of Maxwell equations (MEs) which yield this simultaneity relation. It makes electromagnetic (EM) kinetic simulation to be approximated as electrostatic (ES) kinetic simulation and hence to be of very limited reliability in simulating hot plasmas even though such an approximation is good enough for simulating a cold plasma. On the other hand, many topics on light-matter interaction have urgent demands on the EM kinetic simulation. Even though suitable ansatz of $f$ might be applicable to some special issues [4042], its universal applicability cannot be warranted. Therefore, how to overcome this difficulty is of significant importance.

The other is raised by uniform convergence (with respect to the time-step ${\rm{\Delta }}t$) requirement on calculated Lagrangian particles' time-histories. If each electron is identified by its initial position and initial velocity $\left({r}_{s},{\upsilon }_{s}\right),$ where subscript $s$ means 'starting', the particle simulation is to deal with $N\equiv \displaystyle \iint d{r}_{s}d{\upsilon }_{s}$ mono-variable functions of $t.$ Each mono-variable function, or each particle's trajectory, obeys a relativistic Newton equation (RNE), a $2$-nd differential equation with respect to the variable $t.$ When updating these mono-variable functions, we face an unavoidable difficulty: how large the time-step ${\rm{\Delta }}t$ should be? If expressing each function as a power series of the variable $t,$ one can easily find that each power series, (identified by ${r}_{s},{\upsilon }_{s}$), has its own convergence radius, denoted as ${T}_{{CR}}\left({r}_{s},{\upsilon }_{s}\right),$ and hence naturally define a scalar field ${T}_{{CR}}$ over the initial phase space $\left({r}_{s},{\upsilon }_{s}\right).$ Each calculated trajectory, (identified by ${r}_{s},{\upsilon }_{s}$), can be viewed as loyal if ${\rm{\Delta }}t\lt {T}_{{CR}}\left({r}_{s},{\upsilon }_{s}\right)$ otherwise as losing fidelity. Thus, a loyal updating all particles requires ${\rm{\Delta }}t\lt {T}_{{CR}}\left({r}_{s},{\upsilon }_{s}\right)\forall {r}_{s},{\upsilon }_{s},$ or ${\rm{\Delta }}t\lt {T}_{{CR}}^{{\min }}$ where ${T}_{{CR}}^{{\min }}$ is defined as the minimum of all ${T}_{{CR}}\left({r}_{s},{\upsilon }_{s}\right).$ We call this requirement as uniform convergence requirement. Clearly, this is a quite severe requirement (because ${T}_{{CR}}^{{\min }}=0$ is easily to occur). Of course, ignoring this uniform convergence requirement implies the updating of all particles at the cost of some calculated trajectories losing fidelity. This might be an important reason for many PIC simulations violating MEs [17, 4350]. For Eulerian approach, similar uniform convergence requirement for ensuring $f\geqslant 0$ anywhere is also required.

Initial-value problem can be expressed/interpreted in a more suitable form free from two difficulties. In plasma microscopic theory, a conception similar to the collective mode in plasma fluid theory is $6D$ phase space time-dependent deformation of given initial $f$-profile. This conception can be used to express/interpret the initial-value problem. In plasma fluid theory, detailed mathematic form of a collective mode is derived from conservation laws behind fluid equation, $3D$ real space boundary condition and initial condition. Likewise, in plasma microscopic theory, conservation laws behind the V-M system, $6D$ phase space boundary condition and initial condition can also be used to figure out the detailed mathematic form of time-dependent deformation of given initial $f$-profile. By virtue of a new compact expression of the $f,$ the initial-value problem of $f$ is expressed/interpreted as consistent evolutions of $5$ functions constituting this new expression. The new compact expression fully utilizes geometric/toplogical nature of the $f$-profile due to some global conservation laws, and local conservation laws which reflects some fixed points (of $f$) in $6D$ phase space existing. Its necessity, advantage over other methods, uniqueness and technical details will be presented in following sections. A subtle difference between microscopic scalar field and microscopic vector field a similar scheme of particle simulation.

2. Theory and Method

For outstanding the emergency of overcoming two difficulties mentioned before, we first expound them in section 2.1. Then, we explain, in section 2.2, why others methods fail to overcome them. These explanations point out the importance of a suitable expression of the $f$ in overcoming two difficulties. Technical details of a new EM kinetic theory/simulation method free from two difficulties are presented in section 2.3 and 2.4. A numerical example of the application of such a new method is presented in section 2.5. A comparison with other methods ignoring two difficulties is presented in section 2.6.

2.1. Crucial difficulty in EM kinetic simulation

According to MEs, scalar potential $\phi $ and vector potential $\vec{A}$ satisfy wave equations with different 'source' terms: $n\equiv \displaystyle \int f{d}^{3}\upsilon $ for $\phi $ and $\vec{j}\equiv \displaystyle \int \upsilon f{d}^{3}\upsilon $ for $\vec{A},$ where $f$ is the PDF. Lorentz gauge condition ${\partial }_{t}\phi +{\rm{\nabla }}\cdot \vec{A}=0$ relates $\phi $ and $\vec{A}$ and agrees with continuity equation ${\partial }_{t}n+{\rm{\nabla }}\cdot \vec{j}=0$ reflecting conservation of total particles number. According to standard partial differential equation (PDE) theory, $\phi $ and $\vec{A}$ solutions can be expressed in a general form [51]:

Equation (1)

where $F$ stands for $\phi $ or $\vec{A},$ ${FS}$ stands for its source term charge density $n$ or current density $\vec{j},$ and ${dS}$ is a sphere surface which is centered at $r$ and of a radius ${ct}.$ This general form indicates that $F$ depends on its two initial conditions: $F{{\rm{| }}}_{t=0}$ and ${\partial }_{t}F{{\rm{| }}}_{t=0}.$ Due to Lorentz gauge condition, ${\partial }_{t}\phi {{\rm{| }}}_{t=0}$ can be simultaneous with $\vec{A}{{\rm{| }}}_{t=0},$ and further with $f{{\rm{| }}}_{t=0}.$ But ${\partial }_{t}\vec{A}{{\rm{| }}}_{t=0},$ or initial transverse electric field profile ${\vec{E}}_{{Tr}}{{\rm{| }}}_{t=0},$ has a different simultaneous partner. Now that $\vec{A}$ satisfies a wave equation with a source $\vec{j},$ ${\partial }_{t}\vec{A}$ will satisfy a wave equation with a source ${\partial }_{t}\vec{j}.$ Thus, ${\vec{E}}_{{Tr}}{{\rm{| }}}_{t=0}\equiv -{\partial }_{t}\vec{A}{{\rm{| }}}_{t=0}$ is simultaneous with ${\partial }_{t}f{{\rm{| }}}_{t=0}.$ Moreover, when ${FS}$ in equation (1) obeys an equation dependent on ${\partial }_{t}F{{\rm{| }}}_{t=0},$ such as Vlasov equation (VE) containing ${E}_{{Tr}}\equiv {\partial }_{t}A,$ by setting the result of the operator ${\partial }_{t}$ on the formula equation (1) of $F\equiv A$ at $t=0,$ we can easily obtain a complicated self-consistent equation of ${\partial }_{t}A{{\rm{| }}}_{t=0}$ (because ${\partial }_{t}{FS}$ depends on ${\partial }_{t}A{{\rm{| }}}_{t=0}$). Namely, it is not an easy task to solve an initial condition ${\partial }_{t}A{{\rm{| }}}_{t=0}$ consistent with others such as $f{{\rm{| }}}_{t=0},$ $A{{\rm{| }}}_{t=0}$ etc

$\phi $ and $\vec{A}$ appearing in the VE of $f,$ which formally is a PDE defined over a $6D$ phase space plus a time coordinate, can be completely expressed, through above formula, in terms of $f.$ This makes the VE finally to become a differential-integral equation of $f,$ or an integral equation of local time growth-rate ${\partial }_{t}f$ because ${\vec{E}}_{{Tr}}$ is simultaneous with ${\partial }_{t}f:$

Equation (2)

where $e$ is electronic charge, and ${\vec{E}}_{{Tr}}\left[\int \upsilon {\partial }_{t}f{d}^{3}\upsilon \right]$ represents ${\vec{E}}_{{Tr}}$ being a functional of $\int \upsilon {\partial }_{t}f{d}^{3}\upsilon .$ Moreover, all other terms in the VE are represented by a functional of $f,$ which is denoted as $O\left[f\right]$ and $O$ is the initial letter of ${Others}.$ Its detailed form is easily written out but somewhat long for being presented here. Here, magnetic field $\vec{B}$ and longitudinal electric field ${\vec{E}}_{{Lo}}$ are simultaneous with $f$ and hence contained in ${Others}$-functional.

Once specifying the value of ${\vec{E}}_{{Tr}}{{\rm{| }}}_{t=0}\equiv -{\partial }_{t}\vec{A}{{\rm{| }}}_{t=0},$ we have implicitly specified the value of ${\partial }_{t}f{{\rm{| }}}_{t=0}$ to some extent. Thus, when conducting the V-M simulation in Eulerian approach, substituting this ${\vec{E}}_{{Tr}}{{\rm{| }}}_{t=0},$ (i.e., the specified ${\partial }_{t}f{{\rm{| }}}_{t=0},$) into the VE which is treated as an 'algebra' equation of ${\partial }_{t}f{{\rm{| }}}_{t=0},$ it is necessary to examine whether the implicitly-specified ${\partial }_{t}f{{\rm{| }}}_{t=0}$ agrees with the calculated ${\partial }_{t}f{{\rm{| }}}_{t=0}.$ For brevity, we denote them as ${\partial }_{t}f{{\rm{| }}}_{t=0}^{{MEs}}$ and ${\partial }_{t}f{{\rm{| }}}_{t=0}^{{VE}}$ (where upper index 'MEs' means it being determined by MEs and 'VE' by VE), respectively. If they are different or ${\rm{| }}{\partial }_{t}f{{\rm{| }}}_{t=0}^{{VE}}-{\partial }_{t}f{{\rm{| }}}_{t=0}^{{MEs}}{\rm{| }}$ is not small enough, using the ${\partial }_{t}f{{\rm{| }}}_{t=0}^{{VE}}$ to calculate $f{{\rm{| }}}_{t={\rm{\Delta }}t}$ is futile. (In other words, because we can construct, according to the simultaneity relation above-mentioned, a ${\vec{E}}_{{Tr}}{{\rm{| }}}_{t=0}^{{VE}}$ from ${\partial }_{t}f{{\rm{| }}}_{t=0}^{{VE}},$ we should examine whether ${\vec{E}}_{{Tr}}{{\rm{| }}}_{t=0}^{{VE}}$ agreeing with the specified value of ${\vec{E}}_{{Tr}}{{\rm{| }}}_{t=0},$ or ${\vec{E}}_{{Tr}}{{\rm{| }}}_{t=0}^{{MEs}}.$) Once disagreement occur, we have to try a new trial of ${\vec{E}}_{{Tr}}{{\rm{| }}}_{t=0},$ as well as a new trial of ${\partial }_{t}f{{\rm{| }}}_{t=0}^{{MEs}}$-profile, and calculate its yielded ${\partial }_{t}f{{\rm{| }}}_{t=0}^{{VE}}$-profile, and then compare them. Repeating such a cycle until ${\rm{| }}{\partial }_{t}f{{\rm{| }}}_{t=0}^{{VE}}-{\partial }_{t}f{{\rm{| }}}_{t=0}^{{MEs}}{\rm{| }}$ being small enough (if can). For example, $\frac{{\rm{| }}{\partial }_{t}f{{\rm{| }}}_{t=0}^{{VE}}-{\partial }_{t}f{{\rm{| }}}_{t=0}^{{MEs}}{\rm{| }}}{{\rm{| }}{\partial }_{t}f{{\rm{| }}}_{t=0}^{{MEs}}{\rm{| }}}\lt 10 \% $ is a tolerable level of the relative error. Clearly, such iterative manner is very time-consuming in computer experiment but strictly speaking unable to be omitted.

Of course, if not specifying the ${\partial }_{t}f{{\rm{| }}}_{t=0}$ but taking the VE as an integral equation of ${\partial }_{t}f,$ discrete expression of the integral $\int \upsilon {\partial }_{t}f{d}^{3}\upsilon $ in the VE will lead to a high-dimension matrix, or those values of ${\partial }_{t}f$ at different $\left(t,r,\upsilon \right)$-coordinates will form, (because discretization of the integral with respect to $\upsilon $ needs $\upsilon $-points in sufficient amount), a ${\rm{\infty }}\times {\rm{\infty }}$ matrix equation nearly impossible to-be-solved. Moreover, singularities associated with $\frac{1}{{\rm{| | }}\vec{r}-{\vec{r}}^{{\prime} }{\rm{| }}-{ct}+c{t}^{{\prime} }{\rm{| }}}$ and $\frac{1}{{\rm{| | }}\vec{r}-{\vec{r}}^{{\prime} }{\rm{| }}-{ct}+c{t}^{{\prime} }{{\rm{| }}}^{2}}$ also make the calculation more complicated.

The V-M simulation in Lagrangian approach [4350, 5254] is to use displacements of $f$-elements in $6D$ phase space to describe time-evolution of $f.$ Here, the phrase 'displacement' represents a $1D$ 'coordinate' along a phase space contour defined by $f\left({r}_{s}-{\upsilon }_{s}{\rm{\Delta }}t,{\upsilon }_{s}-a\left({r}_{s},{\upsilon }_{s}\right){\rm{\Delta }}t\right)=f\left({r}_{s},{\upsilon }_{s}\right),$ where $a\left({r}_{s},{\upsilon }_{s}\right)$ stands for accelerating rate and hence depends on $\left({E}_{{Lo}},{E}_{{Tr}},B\right){{\rm{| }}}_{\left(r,t\right)=\left({r}_{s},0\right)}$ and relativistic factor ${\rm{\Gamma }}{{\rm{| }}}_{\upsilon ={\upsilon }_{s}}.$ These displacements also need to be examined whether their yielded ${E}_{{Tr}}$ (through MEs) agrees with ${E}_{{Tr}}$ in the accelerating rate $a\left({r}_{s},{\upsilon }_{s}\right)=a\left[\left({E}_{{Lo}},{E}_{{Tr}},B\right){{\rm{| }}}_{\left(r,t\right)=\left({r}_{s},0\right)},{\rm{\Gamma }}{{\rm{| }}}_{\upsilon ={\upsilon }_{s}}\right].$ Here $\left({E}_{{Lo}},B\right){{\rm{| }}}_{\left(r,t\right)=\left({r}_{s},0\right)}$ are functions of $\int f\left({r}_{s},{\upsilon }_{s}\right){d}^{3}{\upsilon }_{s}$ but ${E}_{{Tr}}{{\rm{| }}}_{\left(r,t\right)=\left({r}_{s},0\right)}$ is a function of $\int {\partial }_{t}f\left({r}_{s},{\upsilon }_{s}\right){d}^{3}{\upsilon }_{s}.$

Similar difficulty also exists for Newton-Maxwell system. In the PIC code [115], ${\vec{E}}_{{Tr}}$ on any macroparticle is of a simultaneity relation with 'accelerating rate' of other macroparticles within adjacent cells (because two interpolation functions, one for from particle to grid and the other for from grid to particle, are taken as only dependent on space coordinates). If using this ${\vec{E}}_{{Tr}}$ to calculate the 'accelerating rate' of a macroparticle, we will have a similar equation (where ${m}_{e}$ is static electronic mass and a macroparticle shares a same charge/mass ratio with an electron)

Equation (3)

and hence need to diagonalize a matrix whose dimension is the number of macroparticles within adjacent cells.

This consistency-examination of ${\vec{E}}_{{Tr}}$ is absent in various algorithms. For example, in the PIC code [116], although particles' information and fields' information are alternatively updated, such a consistency-examination should be made per two time-steps. Unfortunately, it is omitted. This omission might be an important factor causing various 'violations of physics laws' [17]. Moreover, it will also affect implicit EM PIC scheme which utilizes particle enslavement technique [55] because iteratively solving the root of enslaved residual $G\left({E}^{k}\right),$ where $k$ is iteration number, will encounter a difficulty arising from uncertain initial form of $G\left({E}^{0}\right)$ due to ${\vec{E}}_{{Tr}}.$

Because ${E}_{{tr}}$ is taken as playing less role in these issues [439], it is often not taken into account and hence EM kinetic simulations in these works are actually approximated [439]. For example, in some issues on plasmas in magnetic fusion device involving in kinetic simulation on ionic PDF, the self-consistent magnetic field generated by plasma currents, (which is mainly determined by electronic PDF), is neglected and hence ${E}_{{tr}}$ is indeed not taken into account [23, 2830, 33, 34]. Some kinetic simulations are based on Vlasov-Poisson (VP) code to study electronic PDF [3942] and ionic PDF [35, 38]. Some simulations are on 1D model and hence not take into account $B$ and ${E}_{{tr}}$ [38]. Some kinetic simulations on electronic PDF through Eulerian Vlasov code ignored ${E}_{{tr}}$ [18, 21], so did Semi-Lagrangian (SL) Vlasov code on discharge plasmas [31, 32] and PIC code for hot electron [22]. In some $3D$ EM PIC simulations on kinetic-scale plasma, even though fluctuating magnetic field is taken into account, ${E}_{{tr}}$ is still not strictly taken into account [15, 16].

Even though many authors have been aware of some inherent drawbacks of current popular simulation schemes, such as hiring truncation approximation [20], statistical noise [35, 38] etc, the most urgent task, i.e. strictly taking ${E}_{{Tr}}$ into account, have not win sufficient attention. This makes these simulation schemes under the risk of violating MEs [17] and hence discounts their effort in improving these simulation schemes [19, 23, 25]. For example, usually the kinetic simulation is taken as an initial problem evolution of an initial PDF $f{{\rm{| }}}_{t=0},$ but people often choose $f{{\rm{| }}}_{t=0}$ as having a non-zero vortex of the current density or ${\rm{\nabla }}\times \vec{j}{{\rm{| }}}_{t=0}\ne 0.$ For example, $f{{\rm{| }}}_{t=0}={\exp }\left(-\frac{{x}^{2}+{y}^{2}}{{R}^{2}}\right){\exp }\left(-\frac{1-0.01{\upsilon }_{z}/c}{\sqrt{1-{\rm{| }}\upsilon {{\rm{| }}}^{2}/{c}^{2}}}\right)$ means the initial PDF having a Gaussian space transverse shape and is a typical example of ${\rm{\nabla }}\times \vec{j}{{\rm{| }}}_{t=0}\ne 0,$ which will inevitably, according to MEs, trigger ${E}_{{Tr}}$ to be present. Namely, the choice of the initial PDF is inconsistent with hiring the ES approximation or using the VP code.

The V-M simulation in Eulerian approach is finally inevitably frustrated by the simultaneity relation between ${\vec{E}}_{{Tr}}$ and ${\partial }_{t}f.$ Ignoring this simultaneity relation can make the simulation going-on and speed-up but the reliability of yielded results is limited. Therefore, even though this Eulerian approach is applicable for electrostatic (ES) kinetic simulation [3039, 41, 42], its suitability to EM kinetic simulation is still uncertain. This simultaneity relation also have profound effect on the V-M simulation in Lagrangian approach.

For illustrating the effect of the simultaneity relation described above on simulation, we display results based on strict treatment on the simultaneity relation and those based on approximate treatment in figures 1, 2. Here, the phrase 'strict treatment' refers to BOGRs being linked by a matrix, and the phrase 'approximate treatment' refers to simplifying this matrix as a diagonal one by approximating its non-diagonal elements as being of a different simultaneity relation (alike that between ${\vec{E}}_{{Lo}}$ and $f$) and hence merging them into the functional $O$ in equations (2), (3). Figure 1 is for the V-M simulation in Eulerian approach and figure 2 for the PIC simulation. In figure 1, $f$ starts from a familiar Gaussian initial profile $f{{\rm{| }}}_{t=0}={f}_{{\max }}^{0}{\exp }\left(-\frac{{x}^{2}+{y}^{2}}{{R}^{2}}\right){\exp }\left(-1/\sqrt{1-\frac{{\upsilon }_{x}^{2}+{\upsilon }_{y}^{2}}{{c}^{2}}}\right),$ and all discrete grid-points in the phase space are numbered. Likewise, in figure 2, all macroparticles are numbered. Moreover, the term 'driving strength' in figure 1(b) stands for $O(f)$ in equation (2) and ${F}_{x}$ in figure 2(b) for $x$-component of $O$ in equation (3). Parameters in figure 1 are chosen as $R=1{0}^{-6}m$ and ${f}_{{\max }}^{0}=1{0}^{25}{m}^{-3}{c}^{-3}$ (where $c$ is speed of light in vacuum).

Figure 1.

Figure 1. Examples of $2D2V$ Vlasov simulation in Eulerian approach. The profile of dimensionless probability density function $f/{f}_{{\max }}^{0}$ is represented by its values at discrete points of a grid in the dimensionless $4D$ phase space $\left(x/R,y/R,{\upsilon }_{x}/c,{\upsilon }_{y}/c\right),$ the projection of the grid on the $x---y$ plane is a square whose each side contains $6$ points, and that on the ${\upsilon }_{x}---{\upsilon }_{y}$ plane is a circle ${\upsilon }_{x}^{2}+{\upsilon }_{y}^{2}\leqslant {c}^{2}$ whose diameter contains $6$ points, where $c$ is speed of light in vacuum, $R$ is the characteristic length proportional to the value of the density $\int f{d}^{3}\upsilon $ and ${f}_{{\max }}^{0}$ represents the maximum of the initial $f$-profile $f{{\rm{| }}}_{t=0}.$ (a) is for values of these elements of the matrix in equation (2);(b) is for the driving at each grid-point; (c) is for 'local growth rate' $\equiv {\partial }_{t}f$ at each grid-point under strict treatment; (d) is the counterpart of (c) under approximate treatment.(e) is the amplification of the regime around the point $(350,350)$ in (a).

Standard image High-resolution image
Figure 2.

Figure 2. Examples of $2D2V$ PIC simulation. Initially, $2000$ macroparticles are assigned into a $20\times 20$ grid on the $x---y$ plane. Self-consistent magnetic field is taken as along $z$-direction. (a) is for values of these elements of the matrix in equation (3); (b) is for the driving force on each macroparticle; (c) is for macroparticles' accelerating rate according to strict treatment; (d) is the counterpart of (c) according to approximate treatment. (e) is the amplification of the regime around the point $(1000,1000)$ in (a).

Standard image High-resolution image

Clearly, marked differences exist between results based on different treatments. As shown in figures 1(c), (d), the difference is so large that results in figure 1(c) is higher several orders of magnitude than those in figure 1(d). Moreover, although the curve in figure 1(c) and that in figure 1(d) are in same unit, their shapes are obviously different. Using finer mesh in phase space still faces same problem. Even though results in figures 2(c), (d) are at a same order of magnitude, their relative error is nearly $\gt 100 \% .$ Figures 1, 2 display examples in $2D2V$ format. Similar marked differences between results based on different treatments exist universally for EM problem in all formats (from the simplest $1D2V$ format to $3D3V$ one) because there is always a high-dimension matrix (due to ${E}_{{Tr}}$) no matter which of these formats is studied. For example, a typical $1D2V$ initial PDF $f{{\rm{| }}}_{t=0}={f}_{{\max }}^{0}{\exp }\left(-\frac{{x}^{2}}{{R}^{2}}\right){\exp }\left(-\left(1-v\left(t=0\right){\upsilon }_{y}\right)/\sqrt{1-\frac{{\upsilon }_{x}^{2}+{\upsilon }_{y}^{2}}{{c}^{2}}}\right),$ where $v$ is space-independent, can be represented by a mesh in the $1D2V$ phase space, and the number of points in this mesh are usually chosen to be up to $1000$ or higher. Because $v\left(t=0\right)\ne 0$ implies $f{{\rm{| }}}_{t=0}$ being an asymmetric function of ${\upsilon }_{y}$ and only a space variable $x$ exists, there are initially a current vortex ${\partial }_{x}{j}_{y}\ne 0$ and hence a $z$-direction component of $B$ and a fast growing $y$-direction component of ${E}_{{Tr}}.$ As previously mentioned, the presence of ${E}_{{Tr}}$ means a high-dimension matrix to be treated when calculating subsequent time evolution of $f.$ According to figures 3(c), (d), strict treatment and approximated one yield obviously different results of the local growth rate. This is similar to what are displayed by figures 1(c),(d). Figure 3(b) displays ${E}_{{Lo}}$-profile after two time steps $2{\rm{\Delta }}t,$ which exhibits marked space oscillation and hence implies marked density wave.

Figure 3.

Figure 3. Examples of $1D2V$ Vlasov simulation in Eulerian approach. The profile of dimensionless probability density function $f/{f}_{{\max }}^{0}$ is represented by its values at discrete points of a grid in the dimensionless $3D$ phase space $\left(x/R,{\upsilon }_{x}/c,{\upsilon }_{y}/c\right),$ the projection of the grid on the $x$ axis is a line containing $36$ points, and that on the ${\upsilon }_{x}---{\upsilon }_{y}$ plane is a circle ${\upsilon }_{x}^{2}+{\upsilon }_{y}^{2}\leqslant {c}^{2}$ whose diameter contains $6$ points, where $c$ is speed of light in vacuum, $R$ is the characteristic length proportional to the value of the density $\int f{d}^{3}\upsilon $ and ${f}_{{\max }}^{0}$ represents the maximum of the initial $f$-profile $f{{\rm{| }}}_{t=0}.$ (a) is for values of these matrix elements;(b) is for ${E}_{{Lo}}{{\rm{| }}}_{t=2{\rm{\Delta }}t}$ calculated through strict treatment, where ${E}_{{Lo}}$ is in unit of $\frac{{m}_{e}{\omega }_{p}^{3}}{{ce}},$ ${\omega }_{p}\equiv \sqrt{{e}^{2}{n}_{{\max }}^{0}}{m}_{e}{\varepsilon }_{0}$ and ${n}_{{\max }}^{0}$ represents the maximum of initial density profile $\int {f}^{0}d{\upsilon }_{x}d{\upsilon }_{y};$ (c) is for 'local growth rate' $\equiv {\partial }_{t}f$ at each grid-point under strict treatment; (d) is the counterpart of (c) under approximate treatment.

Standard image High-resolution image

The root reason for limiting scientific validity of older approximated methods/approaches is to adopt approximation on the coefficient matrix needing diagonalization in discrete expression of the VE, which consists of ${\partial }_{t}f$-values at ${N}_{r}\times {N}_{\upsilon }$ mesh points of $6D$ phase space:

Equation (4)

where ${\rm{\Delta }}r$ and ${\rm{\Delta }}\upsilon $ are steps of the mesh/grid in the phase space, and detailed forms of the function ${coefg}$ and the functional ${FG}$ can be directly derived from MEs and the functional $O$ in equations (1), (2). After being aware of this root reason, we diagonalize this coefficient matrix, even though doing so is rather time-consuming, to obtain a set of equations, which are straightforward expressions of ${N}_{r}\times {N}_{\upsilon }$ functions ${\partial }_{t}f\left(r,\upsilon ,t\right)$ in terms of $f\left(r,\upsilon ,t\right):$

Equation (5)

where detailed forms of the coefficient function ${coeffd}$ and the functional ${FD}$ are easily written out but somewhat long for being presented here. Although this straightforward expression is now available, another difficulty still hinders calculating the $f{{\rm{| }}}_{t={\rm{\Delta }}t}$-profile from the $f{{\rm{| }}}_{t=0}$-profile and the ${\partial }_{t}f{{\rm{| }}}_{t=0}$-profile (because the time-step ${\rm{\Delta }}t$ should ensure ${f}_{t={\rm{\Delta }}t}\geqslant 0\forall r,\upsilon .$) Because the discrete expression is through ${N}_{r}\times {N}_{\upsilon }$ mesh points in $6D$ phase space, the allowed value of ${\rm{\Delta }}t$ should be less than the minimum of ${N}_{r}\times {N}_{\upsilon }$ values of ${\rm{| }}\frac{0-f}{{\partial }_{t}f}{{\rm{| }}}_{r,\upsilon }{\rm{| }}.$ Usually, a comprehensive discrete expression of a $f$-profile should contain those boundary mesh points where $f=0$ satisfies. Because not all ${N}_{r}\times {N}_{\upsilon }$ values of ${\partial }_{t}f{{\rm{| }}}_{t=0}$ are $\geqslant 0,$ once at some boundary mesh points the values of ${\partial }_{t}f{{\rm{| }}}_{t=0}$ are negative, the allowed value of ${\rm{\Delta }}t$ will be confined to be $\to 0.$ This difficulty is analogous to the uniform convergence requirement mentioned before.

For the V-M simulation, two difficulties drive us to seek for new ways of describing the evolution of $f.$ As shown latter, in a geometric expression of $f,$ ${N}_{r}\times {N}_{\upsilon }$ expression formulas of ${\partial }_{t}f$ are converted into those of ${\partial }_{t}{H}_{0}$ in a same amount. Unlike $f$-values mandatory to be $\geqslant 0,$ ${H}_{0}$-value are allowed to be negative. This overcomes the difficulty arising from the requirement that ${\rm{\Delta }}t$ should warrant $f{{\rm{| }}}_{t={\rm{\Delta }}t}$ being non-negative anywhere.

2.2. Main difficulty in V-M simulation and advantages of combination of different expressions of ${\boldsymbol{f}}$

A main difficulty in simulating the V-M system, both in Eulerian approach and Lagrangian one, is to find an efficient expression of $f$ free from being supplemented by unnecessary approximations. In the Eulerian approach, if $f$ is expressed by a power expansion formula

Equation (6)

the VE will act as a recurrence relation/formula defining a set of expansion coefficient functions $\{{c}_{i\geqslant 0}\left(r,t\right),i\geqslant 0\}.$ $f$ constructed from this $\{{c}_{{\rm{i}}}\}$-set obviously satisfy the VE. The recurrence relation/formula can lead to a general relation between two expansion coefficient functions with adjacent subindex

Equation (7)

which implies ${c}_{i}\sim \frac{1}{i!}{c}_{0}^{\left(0,i\right)}.$ Here, ${c}_{0}^{\left(0,i\right)}\equiv {d}_{{t}^{i}}^{i}{c}_{0}$ and ${c}_{0}^{\left(i,j\right)}\equiv {d}_{{r}^{i}}^{i}{d}_{{t}^{j}}^{j}{c}_{0}$ are common expressions of high-order derivatives of ${c}_{0}\left(r,t\right).$ The convergence condition ${li}{m}_{i\to {\rm{\infty }}}{\rm{| }}\frac{{c}_{i+1}{\left[\upsilon -u\right]}^{i+1}}{{c}_{i}{\left[\upsilon -\vec{u}\right]}^{i}}{\rm{| }}\lt 1,$ whose detailed form is ${\rm{| }}\upsilon -\vec{u}{\rm{| }}/c\lt {li}{m}_{i\to {\rm{\infty }}}\frac{{c}_{i}}{{c}_{i+1}}\sim {li}{m}_{i\to {\rm{\infty }}}\left(i+1\right){li}{m}_{i\to {\rm{\infty }}}\frac{{c}_{0}^{\left(0,i\right)}}{{c}_{0}^{\left(0,i+1\right)}}\lt {li}{m}_{i\to {\rm{\infty }}}\left(i+1\right),$ can be warranted if ${c}_{0}$ can satisfy $c\lt {li}{m}_{i\to {\rm{\infty }}}{\rm{| }}\frac{{c}_{0}^{\left(0,i\right)}}{{c}_{0}^{\left(0,i+1\right)}}{\rm{| }},$ where $c$ is speed of light in vacuum. Namely, once ${c}_{0}$ is convergent with respect to $t,$ it can lead to a set $\{{c}_{i\gt 0}\}$ warranting the expansion series ${\sum }_{i\gt 0}{c}_{i}{\rm{| }}\upsilon -\vec{u}{{\rm{| }}}^{i}$ convergent over the region ${\rm{| }}\upsilon -\vec{u}{\rm{| }}/c\lt {\rm{\infty }}\supset {\rm{| }}\upsilon -\vec{u}{\rm{| }}/c\lt 2.$

The power expansion can be converted into another form

Equation (8)

where Heviside function is defined as $H\left(x\gt 0\right)=1$ and $H\left(x\leqslant 0\right)=0$ and satisfies $2H\left({\rm{| }}x{\rm{| }}\right)=H\left(x\right)+H\left(-x\right),$ and the coefficient function $\{{b}_{i}\}$-set can be straightforward transformed from the $\{{c}_{i}\}$-set.

Equation (9)

Equation (10)

where $\frac{{\left(-1\right)}^{m}{q}^{2m}}{2m!}$ is from Taylor expansion of ${\cos }$ function. The non-negative-probability requirement demands ${b}_{1}\left(r,t\right)\equiv 0$ mandatory to be fulfilled [5658]. Thus, complete microscopic description, which consists of the VE and the non-negative-probability requirement, implies following relation:

Equation (11)

where $C\left({\rm{\Omega }}\right)$ represents the set of all continuous functions defined over $7$-D space ${\rm{\Omega }}$ whose coordinates are $\upsilon ,r,t.$

Note that when the operator $\hat{L}$ acts on ${b}_{i}\cdot {\left(\upsilon -\vec{u}\right)}^{i},$ relevant terms will be in three classes: rising-order terms $\sim {\left(\upsilon -\vec{u}\right)}^{i+1},$ keeping-order terms $\sim {\left(\upsilon -\vec{u}\right)}^{i},$ and decreasing-order ones $\sim {\left(\upsilon -\vec{u}\right)}^{i-1}.$ Thus, there will be a mono-variable recurrence relation/formula chain

Equation (12)

where the functional ${RF}$ represents the recurrence formula reflected by the VE, and expressing $\vec{u}$ as $\vec{u}\left(\vec{E},\vec{B}\right)$ is due to the fact that ${b}_{1}\equiv 0$ can lead to an equation among $\vec{u},\vec{E},\vec{B}$ [5658]

Equation (13)

The relation between ${F}_{i}$ and ${H}_{i}$ is described as follows: substituting the power series expansion back into Gauss law (where these ${b}_{i}$ and $\vec{E}$ are space-time inhomogeneous functions and ${\varepsilon }_{0}$ is permittivity of free space)

Equation (14)

and utilizing ${b}_{i}$'s expression equation (12), we can express ${b}_{0}$ in terms of $\vec{u},\vec{E},\vec{B}$ and hence convert ${F}_{i}$ into ${H}_{i}.$ Note that equation (14) is indeed an infinite-order PDE (with respect to $t$) of ${b}_{0}$ and hence often inevitably trigger approximations.

Semi-Lagrangian approaches [4350, 5254] face similar difficulty which refers to exact integration of contour of $f$ over $\upsilon $-space difficult to be conducted. As a result, even exact functional dependence of $f$ on $\left(\upsilon ,E,B\right)$ can be known from the contour form, it is difficult to obtain exact solution of $f$ in terms of $\left(\upsilon ,r,t\right).$

The VE reflects global conservation law $\int f{d}^{3}\upsilon {d}^{3}r={constant},$ and the non-negative-probability requirement $f\geqslant 0\forall \upsilon ,r,t$ reflects local positivity-preserving law/requirement [4350] which is mandatory for applying probability theory to various application issues. Many authors try to take this local positivity-preserving law/requirement into account by adjusting key technical details used in Lagrangian approaches [4350], such as using more interpolation functions and dividing a time-step into more sub-steps etc Mathematically, these effort choose to update $f$ in a more complicated manner in which the displacement of a $f$-element in $6D$ phase space is a nonlinear function of the time-step ${\rm{\Delta }}t.$ But these effort also cause computation cost increasing significantly.

Historically, in theoretical plasmas physics, investigations are merely around the VE and hence often drop in a logic contradiction: Starting from the VE,$\to $ deriving its equivalent equation ${EE}\{{VE}\},$ (where ${EE}$ is the acronym of 'equivalent equation' and the phrase 'equivalent' means no any mathematics approximation being applied and hence can share common solutions with the VE),$\to $ then truncating this equivalent equation to obtain an equation ${TEE}\{{VE}\},$ (where ${TEE}$ is the acronym of 'truncated equivalent equation' and the phrase 'truncating' means discarding some non-zero terms in the equation and hence cannot share common solutions with the equivalent equation),$\to $ then combining the amputated ${TEE}\{{VE}\}$ and the VE to obtain solutions of the VE. Obviously, such a procedure is contradict with the fact the a ${TEE}\{{VE}\}$ cannot share solutions with the equivalent equation ${EE}\{{VE}\},$ as well as the VE. Actually, people have been aware of that 'the 'truncation approximation' is a delicate operation, which might lead to a numerical failure of the solution for later times' [20]. Therefore, ignoring the non-negative-probability requirement delays a sound kinetic simulation scheme to be found.

There are different expressions of the PDF $f.$ For example, the set $\{\int {\left[\upsilon -\vec{u}\right]}^{i}f{d}^{3}\upsilon {\rm{;}}i\geqslant 0\},$ where $\vec{u}\left(r,t\right)\equiv \frac{\int \upsilon f{d}^{3}\upsilon }{\int f{d}^{3}\upsilon },$ is convenient to reflect the global conservation law, and the set $\{{f}^{\left(i\right)}{{\rm{| }}}_{\upsilon =\vec{u}}\equiv \frac{{d}^{i}f}{d{\upsilon }^{i}}{{\rm{| }}}_{\upsilon =\vec{u}}{\rm{;}}i\gt 0\}$ is convenient to reflect the local positivity-preserving law/requirement. Compared with sole usage of these expressions, combinative usage of them can avoid unnecessary approximations and is more efficient for establishing a universal and sound scheme of the EM kinetic simulation.

2.3. A compact self-consistent Eulerian expression of ${\boldsymbol{f}}$

Among different expressions of $f,$ choosing the most suitable is important to obtain exact solutions of $f.$ An unsuitable expression will inevitably lead to unnecessary mathematical approximations supplemented. As far as $f$ is concerned, previously mentioned two Eulerian expressions: mesh expression and power series expressions are far from suitable ones. For example, to make the $\{{b}_{i}\}$-set certain needs to first make the ${b}_{0}$ certain from equation (14)), an infinite-order PDE (with respect to $t$) of ${b}_{0},$ and hence often triggers approximations. In contrast, respecting some mathematical features of $f$ and targeted introducing more suitable Eulerian expression is feasible. For example, because equation (14), or Gauss law, must be respected, the $\{{b}_{i}\}$-set expression can be further improved into a more suitable expression in which $f$ is expressed in term of $f$ itself [59]:

Equation (15)

where ${\rm{\Delta }}n\equiv n-{b}_{0}={\varepsilon }_{0}{\rm{\nabla }}\cdot \vec{E}-{b}_{0}$ is always $\geqslant 0,$ non-negative function $\left[{H}_{1}+{H}_{0}\right],$ which acts as a conditional probability density function (C-PDF), describes the allocation of particles contained in the ${\rm{\Delta }}n$-profile into different values of $\upsilon -\vec{u}\ne 0.$ It is straightforward, from the definition formula equation (15), to derive that the non-negative function $\left[{H}_{1}+{H}_{0}\right]$ satisfies $\int \left[{H}_{1}+{H}_{0}\right]{d}^{3}\upsilon =1,\int \left[\upsilon -\vec{u}\right]\cdot \left[{H}_{1}+{H}_{0}\right]{d}^{3}\upsilon =0,\left[{H}_{1}+{H}_{0}\right]\left(\vec{u},r,t\right)\equiv 0\equiv \left[{H}_{1}+{H}_{0}\right]\left({\rm{| }}\upsilon {\rm{| }}=c,r,t\right).$ Clearly, such a non-negative function can be constructed from functions in two different classes, one class is denoted as ${H}_{1}$ and the other ${H}_{0}$ [56]:

Equation (16)

Equation (17)

and ${H}_{1}$ is $\geqslant 0\forall \upsilon ,r,t$ while ${H}_{0}$ is not. Moreover, it is easy to find, by applying ${\partial }_{t}$-operator and ${{\rm{\nabla }}}_{r}$-operator on the definition formulas equations (16), (17), that, $4$ functions: ${\partial }_{t}{H}_{1},$ ${{\rm{\nabla }}}_{r}{H}_{1},$ ${\partial }_{t}{H}_{0}$ and ${{\rm{\nabla }}}_{r}{H}_{0},$ (denoting them as $F$), are alike to ${H}_{0}$ or satisfy $\int F{d}^{3}\upsilon =0=\int \left[\upsilon -\vec{u}\right]* F{d}^{3}\upsilon .$ Note that any solution of equation (17) multiplying a $\upsilon $-independent function is still a solution of equation (17), but this important property does not hold for solutions of equation (16).

If an initial profile $f{{\rm{| }}}_{t=0}$ can correspond to a $\{{b}_{i\ne 0}{{\rm{| }}}_{t=0}\}$-set convergent with respect to $i,$ it will yield at least $5$ initial profiles: $n{{\rm{| }}}_{t=0},$ $\vec{u}{{\rm{| }}}_{t=0},$ ${b}_{0}{{\rm{| }}}_{t=0},$ ${H}_{1}{{\rm{| }}}_{t=0}$ and ${H}_{0}{{\rm{| }}}_{t=0}\equiv 0.$ Because $W\left(\vec{u},m\right)\equiv {\int }_{{\rm{| }}\upsilon {{\rm{| }}}^{2}\lt {c}^{2}}{\left[\upsilon -\vec{u}\right]}^{m}{d}^{3}\upsilon $ is a function of $\vec{u}$ and $m,$ ${\int }_{{\rm{| }}\upsilon {{\rm{| }}}^{2}\lt {c}^{2}}{H}_{1}{d}^{3}\upsilon $ is therefore a combination of these functions of $\vec{u}$ and $m:$

Equation (18)

where ${\tilde{b}}_{m\geqslant 2}\equiv \frac{{b}_{m\geqslant 2}{\rm{| }}\left(r,t\right)}{{\rm{\Delta }}n\left(r,t\right)},$ and its value equaling to $1$ implies an equation of combination coefficients ${\tilde{b}}_{m\geqslant 2}.$ Likewise, ${\int }_{{\rm{| }}\upsilon {{\rm{| }}}^{2}\lt {c}^{2}}\left[\upsilon -\vec{u}\right]* {H}_{1}{d}^{3}\upsilon =0,$ ${H}_{1}\left(\vec{u},r,t\right)\equiv 0\equiv {H}_{1}\left({\rm{| }}\upsilon {\rm{| }}=c,r,t\right)$ and equation (18) also imply equation of combination coefficients ${\tilde{b}}_{m\geqslant 2}.$ If expressing these combination coefficients in terms of power series of $\tilde{\vec{u}}\equiv \vec{u}/c:$ $\{{\tilde{\vec{u}}}^{0},{\tilde{\vec{u}}}^{1},...,{\tilde{\vec{u}}}^{n},...\},$ equations (14), (16) will imply that the ${H}_{1}{{\rm{| }}}_{t=0}$-profile naturally defines a matrix $M$ linking the set $\{...,{\tilde{b}}_{i\geqslant 2},...\}$ with the set $\{{\tilde{\vec{u}}}^{0},{\tilde{\vec{u}}}^{1},...,{\tilde{\vec{u}}}^{n},...\}$

Equation (19)

Now that ${\sum }_{i\geqslant 2}{\tilde{b}}_{i}\left(\vec{u}\left(r,t=0\right)\right)* {\left[\upsilon -\vec{u}\left(r,t=0\right)\right]}^{i}$ can satisfy equations (14), (16), (17) at $t=0,$ it is easy to prove its counterpart, which is to replace $\vec{u}\left(r,0\right)$ with $\vec{u}\left(r,{\rm{t}}\right),$ to still satisfy equations (14), (16), (17) at $t={\rm{\Delta }}t.$ We term it as translation operation on the ${H}_{1}{{\rm{| }}}_{t=0}$-profile and denote it as $\hat{{TL}}\left({\rm{\Delta }}t\right){H}_{1}{{\rm{| }}}_{t=0}$ in which $\vec{u}\left(r,0\right)$ in ${H}_{1}{{\rm{| }}}_{t=0}$ is replaced by $\vec{u}\left(r,{\rm{\Delta }}t\right),$ but the matrix linking $\{...,{\tilde{b}}_{i\geqslant 2}{{\rm{| }}}_{t={\rm{\Delta }}t},...\}$ and $\{...,{\tilde{\vec{u}}}^{m}\left(r,{\rm{\Delta }}t\right),...\}$ is as same as that linking $\{...,{\tilde{b}}_{i\geqslant 2}{{\rm{| }}}_{t=0},...\}$ and $\{...,{\tilde{\vec{u}}}^{m}\left(r,0\right),...\}.$ Such an operation can be expressed via a formula

Equation (20)

and the function $\lambda \left({\rm{\Delta }}t\right)\equiv 1$ if $g$ represents ${H}_{1}$ (because $\int {H}_{1}{{\rm{| }}}_{t=0}{d}^{3}\upsilon \equiv 1$ and $\int \hat{{TL}}\left({\rm{\Delta }}t\right){H}_{1}{{\rm{| }}}_{t=0}{d}^{3}\upsilon \equiv 1$ do not allow $\lambda \left({\rm{\Delta }}t\right)\ne 1$ ), but $\lambda \left({\rm{\Delta }}t\right)$ can be not constant-valued if $g$ represents ${H}_{0}$ (because $\int {H}_{0}{{\rm{| }}}_{t=0}{d}^{3}\upsilon \equiv 0$ and $\int \hat{{TL}}\left({\rm{\Delta }}t\right){H}_{0}{{\rm{| }}}_{t=0}{d}^{3}\upsilon \equiv 0$ allow $\lambda \left({\rm{\Delta }}{\rm{t}}\right)\ne 1$ ).

The matrix linking the $\{{\tilde{b}}_{i\geqslant 2}\}$-set and the $\{{\tilde{\vec{u}}}^{m}\}$-set reflects geometric/toplogical nature/shape of the ${H}_{1}$-profile, and likewise the ${H}_{0}$-profile also has a matrix reflecting its geometric/toplogical nature/shape. If this matrix is unchanged during evolution, the related profile is viewed as keeping geometric/toplogical nature/shape. Thus, the phrase 'deformation' refers to this matrix varying with respect to $t$ and hence the geometric/toplogical nature/shape is not fixed relative to an observer in the $u$-frame. Namely, the ${H}_{0}$-profile can allow deformation, (relative to an observer in the $u$-frame), varying with respect to $t.$

2.4. Interpreting the initial-value problem of V-M system in Eulerian-like approach as collective mode in ${\bf{6}}{\boldsymbol{D}}$ phase space

Because of $\int {H}_{0}{d}^{3}\upsilon =0,$ ${H}_{0}$ is therefore not a function positivity-preserving, and hence has zeros, beside $\upsilon =\vec{u}$ and ${\rm{| }}\upsilon {{\rm{| }}}^{2}={c}^{2},$ in $\upsilon $-space. Thus, a typical form of ${H}_{0},$ or that of functions satisfying equation (17), is

Equation (21)

each root ${\rm{| }}{Zer}{o}_{m}/c{\rm{| }}\lt 1$ is mono-variable functional of $\vec{u}$ and each number ${Re}{p}_{m}$ is repetition index of the root ${Zer}{o}_{m}$ (and at least a root satisfies ${Zer}{o}_{n}\left(\vec{u}\right)\equiv \vec{u}$). Such a product ${\prod }_{m}$ represents a class of $\upsilon $-space geometric/toplogic nature/shape. Any $\upsilon $-independent function multiplying this typical form still satisfies equation (17). From

Equation (22)

Equation (23)

we can find

Equation (24)

Equation (25)

and (in similar procedure)

Equation (26)

Equation (27)

Because of

Equation (28)

Equation (29)

Equations (24)–(29) will yield

Equation (30)

Equation (31)

If we use the symbol '$\{{H}_{0}\}$' to represent the set of functions satisfying equation (17), equations (30), (31) imply that if ${H}_{0}^{{typical}}\in \{{H}_{0}\}$ exists, so does ${\partial }_{\vec{u}}{H}_{0}^{{typical}}\in \{{H}_{0}\}.$

The VE is indeed an expression of the local growth rate ${\partial }_{t}\left[{H}_{1}+{H}_{0}\right]$ [59]:

Equation (32)

where the effect of the operator $\hat{{GR}}$ on a function $F$ obey following formula [59]

Equation (33)

Equation (34)

Note that a fraction of the local growth rate

Equation (35)

is responsible for $\hat{{TL}}\left(t\right)F{{\rm{| }}}_{t=0},$ and another fraction

Equation (36)

is responsible for the deformation of $F$ relative to an observer in the $u$-frame.

Because ${H}_{0}$ must satisfy equation (17), its general form reads $\lambda \left(t\right){H}_{0}^{{typical}}.$ Substituting this general form into the VE, we have

Equation (37)

or

Equation (38)

where

Equation (39)

Equation (40)

Equation (41)

Equation (42)

and the functional $K$ defined by equation (34) has been contained in $\hat{{GR}}{H}_{1},$ as well as in $\hat{{DEGR}}{H}_{1}.$

Because of equations (16), (17), (32)–(36), it is easy to find that

Equation (43)

Equation (44)

, or $\hat{{DEGR}}\left[{H}_{1}+{H}_{0}\right]\in \{{H}_{0}\}.$ Therefore, the VE reflects a relation among these members of the set $\{{H}_{0}\},$ and has been converted into equation (38), a formal $1$st-order ordinary differential equation (ODE) of ${H}_{0}^{{typical}}$ with respect to variable $\vec{u}.$ Its formal strict solution

Equation (45)

reflects a relation between ${H}_{0}^{{typical}},$ a member of $\{{H}_{0}\},$ and another member $\hat{{DEGR}}{H}_{1}.$

Asymptotic behavior of ${H}_{0}$ at $t=0$ are $0={H}_{0}{{\rm{| }}}_{t=0}=\{\lambda * {H}_{0}^{{typical}}\}{{\rm{| }}}_{t=0}$ and $0\ne {\partial }_{t}\{\lambda * {H}_{0}^{{typical}}\}{{\rm{| }}}_{t=0},$ and lead to

Equation (46)

Equation (47)

which can easily lead to

Equation (48)

where $\nu $ is a constant corresponding to the minimum of all time scales (each represents the time for $\left[{H}_{1}+{H}_{0}\right]$ at a given coordinate $\left(\upsilon ,r\right)$ decreasing to $0:$

Equation (49)

Such a choice of $\nu $ is for warranting ${H}_{1}+{H}_{0}={H}_{1}+\lambda \left(t\right){H}_{0}^{{typical}}$ non-negative always. Thus, equations (45), (48), (49) reveal that the C-PDF ${H}_{1}+{H}_{0}$ corresponds to a time-dependent global oscillation of ${H}_{0}$ around ${H}_{0}\equiv 0\forall \upsilon ,r,t,$ or that of ${H}_{1}+{H}_{0}$ around $\hat{{TL}}{H}_{1}.$ Because these fixed points, which correspond to ${H}_{0}=0,$ exist, the collective mode exhibits as a 'standing-wave' in $6D$ phase space.

Considering the phrase 'Eulerian' is often for describing $f$ by its values at a mesh of the $6D$ phase space, we term above procedure as 'Eulerian-like' because it is based on the description of $f$ by specified functions defined over the $6D$ phase space. These specified functions reflect geometric/toplogial nature of initial $f$-profile. The well-known drawback of the Eulerian approach, i.e., too huge mount of mesh points and corresponding discrete values, is avoided by high-efficiently expressing them through collective mode, represented by the ${H}_{0}$-profile, in the $6D$ phase space. Namely, the Eulerian approach is in a one-by-one manner, to update a mesh ${f}_{{Ir},I\upsilon }^{n}$ to another mesh ${f}_{{Ir},{\rm{I}}\upsilon }^{n+1}$ through a mesh ${\partial }_{t}{f}_{{Ir},I\upsilon }^{n}.$ Each mesh has its own geometric/tpolgical nature. In contrast, our approach is to utilize respective geometric/topologic nature of three meshes and high-efficiently express the difference ${f}_{{Ir},I\upsilon }^{n+1}-{f}_{{Ir},I\upsilon }^{n}$ as a collective mode ${H}_{0}^{n+1}-{H}_{0}^{n},$ which is determined by ${H}_{1}^{n}.$ One-by-one mesh points calculation on individual $f$-values is replaced by calculating amplitude of the collective mode. Moreover, compared with various ansatz of $f$ [40] used in simulating the V-M system, $f=n* \left[{H}_{1}+{H}_{0}\right]$ is a strict expression. On the other hand, the Lagrangian approach is to express the initial-value problem of the V-M system or the N-M system as those of many initial mono-variable functions (of $t$), or Lagrangian particles, identified by different points on a $6D$ $\left({r}_{s},{\upsilon }_{s}\right)$-profile. By now, the calculation in the Lagrangian approach is rarely to take into account the geometrical/toplogical nature of this $6D$ initial-values-profile, and to choose a rough treatment in which all initial-value problems of these Lagrangian particles are uniformly/equally treated as having a same time-convergence. This might be an unavoidable difficulty for the Lagrangian approach.

2.5. Example of application of this simulation based on the 5-function expression of ${\boldsymbol{f}}$

Here, we present some examples of application of this new simulation method. It simulates an electron bunch (e-bunch) travelling within a space inhomogeneous DC-fields configuration. The DC-fields configuration is described as follows: ${\vec{E}}_{{DC}}={E}_{0}\mathop{{e}_{x}}\limits^{\leftarrow}$ and ${\vec{B}}_{{DC}}=\vec{B}\left(x\right)\mathop{{e}_{z}}\limits^{\leftarrow}={B}_{0}\frac{x}{w}\mathop{{e}_{z}}\limits^{\leftarrow}$ if ${\rm{| }}x{\rm{| }}\lt w$ and ${\vec{B}}_{{DC}}={B}_{0}{sign}(x)\mathop{{e}_{z}}\limits^{\leftarrow}$ elsewhere. Namely, ${\rm{| }}{\vec{B}}_{{DC}}$ drops from ${B}_{0}$ at $x=w$ to $-{B}_{0}$ at $x=-w.$ The initial PDF of the e-bunch is $f{{\rm{| }}}_{t=0}={\exp }\left(-\frac{{\left(x-{x}_{0}\right)}^{2}+{\left(y-{y}_{0}\right)}^{2}}{{R}_{T}^{2}}-\frac{{z}^{2}}{{R}_{L}^{2}}\right){\exp }\left(-\frac{1}{\sqrt{1-{\rm{| }}\upsilon /c{{\rm{| }}}^{2}}}+\beta \cdot \frac{{\upsilon }_{z}/c}{\sqrt{1-{\rm{| }}\upsilon /c{{\rm{| }}}^{2}}}\right),$ where ${R}_{T}=30\mu m$ and ${R}_{L}=10\mu m$ are transverse and longitudinal radius respectively and $\beta c=0.05c$ is the global moving velocity (along the $z$-direction) of the e-bunch, ${x}_{0}=5\mu m$ and ${y}_{0}=5\mu m$ are initial transverse coordinates of the symmetric axis of the e-bunch relative to the $z$-axis. Note that this $f{{\rm{| }}}_{t=0}$-profile corresponds to initial current profile having a vortex ${\rm{\nabla }}\times \vec{j}{{\rm{| }}}_{t=0}\ne 0.$ As described in last subsection, we can derive, from this initial PDF, $5$ derivative initial profiles and calculate their evolutions. Some examples are presented in figures 4, 5. Velocity subspace boundary conditions are ${H}_{0}{{\rm{| }}}_{{v}^{2}={c}^{2}}=0={H}_{1}{{\rm{| }}}_{{v}^{2}={c}^{2}},$ ${H}_{0}{{\rm{| }}}_{v=u}=0={H}_{1}{{\rm{| }}}_{v=u}.$ ${H}_{0}$ and ${H}_{1}$ are finite-valued and ${H}_{1}$ are always non-negatively-valued. These examples well illustrate that phase space details of the evolving $f$-profile can be well reflected by those of the evolving ${H}_{0}$-profile.

Figure 4.

Figure 4. Snapshots of $H$ at $\left({\upsilon }_{y},{\upsilon }_{z}\right)=\left(0,0\right),$ where $\lambda \equiv c/\nu $ and $\nu $ is the frequency of the oscillation in the off-center density $n-{n}_{0}.$

Standard image High-resolution image
Figure 5.

Figure 5. Snapshots of $f\left(t,r,{\rm{\upsilon }}\ne u\right)\equiv \left[n-{n}_{0}\right]H$ at $\left({\upsilon }_{y},{\upsilon }_{z}\right)=\left(0,0\right),$ where $\lambda \equiv c/\nu $ and $\nu $ is the frequency of the oscillation in the off-center density $n-{n}_{0}.$

Standard image High-resolution image

2.6. Comparison with others

For comparison with other kinetic simulation schemes, we present, in figure 6, results of applying this method to laser wakefield, a topic attracting many PIC simulations [60, 61]. Electronic density wake of a laser with a triangular-function-type vector potential initial envelope $a\left(t,r\right)={a}_{i}{\cos }\left(0.5\pi t/{\tau }_{i}\right){\cos }\left(0.5\pi r/{\sigma }_{i}\right),$ which is displayed in figure 6(a), has a shape similar to that displayed in figure 1 of [60], For a laser with a Gaussian-function-type initial envelope $a={a}_{0}{\exp }\left(-{z}^{2}/2{L}^{2}\right){\exp }\left(-{x}^{2}/{r}_{0}^{2}\right),$ electronic density wake displayed in figure 6(c) has a shape similar to that displayed in figure 19 of [61].

Figure 6.

Figure 6. Snapshots of electronic density profile and longitudinal electric field profile of the wakefield of a laser propagating in a plasma layer from left-to-right. (a,b) are at $t=70\lambda /c$ and for laser initial envelope and plasma parameters in figure 1 of [60]: ${\tau }_{i}=6.6{fs},{a}_{i}=1.7,{\sigma }_{i}=5\lambda ,\lambda =1\mu m,{n}_{{av}}={n}_{e}=3.5\times 1{0}^{19}c{m}^{-3}.$ (c,d) are at ${\omega }_{p}t=25$ and for initial laser initial envelope and plasma parameters in figure 19 of [61]: ${a}_{0}=5,L=4.2\mu m,{r}_{0}=9\mu m,\lambda =0.8\mu m,{n}_{{av}}={n}_{0}=7\times 1{0}^{18}c{m}^{-3}$ and same numerical parameters: ${dx}={dy}=0.7\mu m$ and ${dz}=0.03\mu m.$

Standard image High-resolution image

Because of incomplete information disclosure in these refs, such as missing detailed forms of the interpolation functions used in PIC simulations, it is only possible to make a qualitative comparison. Our results presented in figure 6 also display 'bubble structure' displayed in figure 19 of [61]. But the depth and longitudinal position of density valley differ from those in figure 19 of [61]. For example, in the upper-left panel, the region $5\lt z/{lambda}\lt 10,x/{lambda}0$ is a density valley, it is spaced by a density peak from another density valley at the region $15\lt z/{lambda}\lt 25,x/{lambda}0.$ Likewise, in the down-left panel, a large density valley is at the region $25\lt z/{lambda},$ another small valley is at the region $15\lt z/{lambda}\lt 16.$

But in-depth understanding the unavoidable difficulty in particle simulation and seeking for possible solution for it are necessary. A possible solution might arise from a subtle, dual role of an electron. Each electron can be both the source of a scalar potential and that of a vector potential. A notable fact is that scalar potentials from different source electrons will not offset mutually but vector potentials will. This fact implies that the microscopic vector field exerted by an electron can be divided into two parts: ${Vec}={Ve}{c}_{c}+{Ve}{c}_{o},$ where the subscript 'c' represents 'contributed' (to macroscopic vector field) and 'o' represents 'offset'. For all electrons, there are at least a scalar field and two vector fields defined over the initial phase space $\left({r}_{s},{\upsilon }_{s}\right)$ giving a description on the system. One vector field is from those ${Ve}{c}_{o}$ and the other from those ${Ve}{c}_{c}.$ For example, accelerating rate of each electron $a\left(t{\rm{;}}{r}_{s},{\upsilon }_{s}\right)$ is a vector, and hence there should be

Equation (50)

Equation (51)

These unique properties of the ${Ve}{c}_{o}$-field can be utilized for solving difficulties mentioned before.

Note that the MEs do not couple with ${Ve}{c}_{o}.$ The uniform convergence requirement forces us to re-consider the role of those RNEs. It is appropriate to view them as a definition of respective ${a}_{o},$ or all those RNEs yield the expression of an ${a}_{o}$-field, which does not couple with the MEs. In other words, the particle simulation can be in a more efficient scheme if noticing the division ${Vec}={Ve}{c}_{c}+{Ve}{c}_{o}$ and the fact that the MEs do not couple with ${Ve}{{\rm{c}}}_{o}.$ Ignoring them will make the particle simulation entangled by exactly solving ${a}_{o}$ from those RNEs. Once the ${a}_{o}$ field cannot be exactly solved, the whole particle simulation is ruined. Actually, it is unnecessary to bind the particle simulation with this nearly impossible task (because of the uniform convergence requirement). The condition for $N$ expression formula of ${a}_{o}$ able to be simultaneously solved, or the fact that those ${Ve}{c}_{c}$ (such as ${\upsilon }_{c}\equiv u,$ ${E}_{{Lo}},$ ${E}_{{Tr}}$ and $B$) allow so diverse, even opposite, values of ${a}_{o},$ lead to an equation of these ${Ve}{c}_{c},$ and this equation couples with the MEs to form a natural closure of these ${Ve}{c}_{c}$ (because the MEs do not couple with ${Ve}{c}_{o}$) [56].

3. Summary

Two difficulties previously mentioned limit the validity and efficiency of the EM kinetic simulation in both Eulerian approach and Lagrangian one. This forces us to try more efficient methods on kinetic theory. Starting from a compact expression of $f$ and utilizing geometrical/toplogical nature of initial conditions and conservation laws behind the V-M system, we can take a more efficient form of the initial-value problem of the V-M system. Difference among these ${BOGRs}$ in the EM kinetic simulation in both Eulerian approach and Lagrangian one are contained in the geometrical/toplogical nature of the $f{{\rm{| }}}_{t=0}$-profile and are taken into account. The initial-value problem of V-M system is therefore condensed into that of the magnitude of the ${H}_{0}$-profile. In short, this new method enables us to consistently treat these ${BOGRs}$ while treating them through the EM kinetic simulation in both Eulerian approach and Lagrangian one is always perplexed by two difficulties.

Acknowledgments

This work is supported by NSFC.Grant.No.12074398.

Data availability statement

All data that support the findings of this study are included within the article (and any supplementary files).

Please wait… references are loading.
10.1088/2516-1067/ac743e