This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Quasilinear theory for inhomogeneous plasma

I. Y. Dodin\corresp [email protected] \affPrinceton Plasma Physics Laboratory, Princeton, NJ 08543, USA \affDepartment of Astrophysical Sciences, Princeton University, Princeton, NJ 08544, USA
Abstract

This paper presents quasilinear theory (QLT) for classical plasma interacting with inhomogeneous turbulence. The particle Hamiltonian is kept general; for example, relativistic, electromagnetic, and gravitational effects are subsumed. A Fokker–Planck equation for the dressed ‘oscillation-center’ distribution is derived from the Klimontovich equation and captures quasilinear diffusion, interaction with the background fields, and ponderomotive effects simultaneously. The local diffusion coefficient is manifestly positive-semidefinite. Waves are allowed to be off-shell (i.e. not constrained by a dispersion relation), and a collision integral of the Balescu–Lenard type emerges in a form that is not restricted to any particular Hamiltonian. This operator conserves particles, momentum, and energy, and it also satisfies the H\smash{H}-theorem, as usual. As a spin-off, a general expression for the spectrum of microscopic fluctuations is derived. For on-shell waves, which satisfy a quasilinear wave-kinetic equation, the theory conserves the momentum and energy of the wave–plasma system. The action of nonresonant waves is also conserved, unlike in the standard version of QLT. Dewar’s oscillation-center QLT of electrostatic turbulence (1973, Phys. Fluids 16, 1102) is proven formally as a particular case and given a concise formulation. Also discussed as examples are relativistic electromagnetic and gravitational interactions, and QLT for gravitational waves is proposed.

1 Introduction

1.1 Background

Electromagnetic waves are present in plasmas naturally, and they are also launched into plasmas using external antennas, for example, for plasma heating and current drive (Stix, 1992; Fisch, 1987; Pinsker, 2001). Nonlinear effects produced by these waves are often modeled within the quasilinear (QL) approximation, meaning that the nonlinearities are retained in the low-frequency (‘average’) dynamics but neglected in the high-frequency dynamics. Two separate paradigms exist within this approach.

In the first paradigm, commonly known as ‘the’ QL theory (QLT), the focus is made on resonant interactions. Nonresonant particles are considered as a background that is homogeneous in spatial (Vedenov et al., 1961; Drummond & Pines, 1962; Kennel & Engelmann, 1966; Rogister & Oberman, 1968, 1969) or generalized coordinates (Kaufman, 1972; Eriksson & Helander, 1994; Catto et al., 2017); then the oscillating fields can be described in terms of global modes. This approach has the advantage of simplicity, but its applications are limited in that real plasmas are never actually homogeneous in any predefined variables (and, furthermore, tend to exhibit nonlinear instabilities in the presence of intense waves). The ‘ponderomotive’ dynamics determined by the gradients of the wave and plasma parameters is lost in this approach; then, spurious effects can emerge and have to be dealt with (Lee et al., 2018).

The second paradigm successfully captures the ponderomotive dynamics by introducing effective Hamiltonians for the particle average motion (Gaponov & Miller, 1958; Motz & Watson, 1967; Cary & Kaufman, 1981; Kaufman, 1987; Dodin, 2014). But as usual in perturbation theory (Lichtenberg & Lieberman, 1992), those Hamiltonians are by default singular for resonant interactions. Thus, such models have limited reach as well, and remarkable subtleties are still found even in basic QL problems. For example, it is still debated (Ochs & Fisch, 2021a; Ochs, 2021) to which extent the QL effects that remove resonant particles while capturing their energy (Fisch & Rax, 1992) also remove charge along with the resonant particles thereby driving plasma rotation (Fetterman & Fisch, 2008). This state of affairs means, arguably, that a clear comprehensive theory of QL wave–plasma interactions remains to be developed — a challenge that must be faced.

The first framework that subsumed both resonant and nonresonant interactions in inhomogeneous plasmas was proposed by Dewar (1973) for electrostatic turbulence in nonmagnetized plasma and is known as ‘oscillation-center’ (OC) QLT. It was later extended by McDonald et al. (1985) to nonrelativistic magnetized plasma. However, both of these models are partly heuristic and limited in several respects. For example, they are bounded by the limitations of the variational approach used therein, and they separate resonant particles from nonresonant particles somewhat arbitrarily (see also (Ye & Kaufman, 1992)). Both models also assume specific particle Hamiltonians and require that waves be governed by a QL wave-kinetic equation (WKE), i.e. be only weakly dissipative, or ‘on-shell’. (Somewhat similar formulations were also proposed, independently and without references to the OC formalism, in (Weibel, 1981; Yasseen, 1983; Yasseen & Vaclavik, 1986).) This means that collisions and microscopic fluctuations are automatically excluded. Attempts to merge QLT and the WKE with theory of plasma collisions were made (Rogister & Oberman, 1968; Schlickeiser & Yoon, 2014; Yoon et al., 2016) but have not yielded a local theory applicable to inhomogeneous plasma. In particular, the existing models rely on global-mode decompositions and treat complex frequencies heuristically. Thus, the challenge stands.

Related problems are also of interest in the context of gravitostatic interactions (Chavanis, 2012; Hamilton, 2020; Magorrian, 2021), where inhomogeneity of the background fields cannot be neglected in principle (Binney & Tremaine, 2008). (To our knowledge, OC QLT analogs have not been considered in this field.) Similar challenges also arise in QLT of dispersive gravitational waves (Garg & Dodin, 2021a, 2020). Hence, one cannot help but wonder whether a specific form of the particle Hamiltonian really matters for developing QLT or it is irrelevant and therefore should not be assumed. Since basic theory of linear waves is independent of Maxwell’s equations (Tracy et al., 2014; Dodin & Fisch, 2012; Dodin et al., 2017), a general QLT might be possible too, and it might be easier to develop than a zoo of problem-specific models.

1.2 Outline

Here, we propose a general QLT that allows for plasma inhomogeneity and is not restricted to any particular Hamiltonian or interaction field. By starting with the Klimontovich equation, we derive a model that captures QL diffusion, interaction with background fields, and ponderomotive effects simultaneously. The local diffusion coefficient in this model is manifestly positive-semidefinite. Waves are allowed to be off-shell, and a collision integral of the Balescu–Lenard type emerges for general Hamiltonian interactions. This operator conserves particles, momentum, and energy, and it also satisfies the H\smash{H}-theorem, as usual. As a spin-off, a general expression for the spectrum of microscopic fluctuations of the interaction field is derived. For on-shell waves governed by the WKE, the theory conserves the momentum and energy of the wave–plasma system. The action of nonresonant waves is also conserved, unlike in the standard version of QLT.111The standard QLT (as in, for example, (Drummond & Pines, 1962)) does not properly conserve energy–momentum either, even though it is formally conservative (see section 7.3.2). Dewar’s OC QLT of electrostatic turbulence (Dewar, 1973) is proven formally as a particular case and given a concise formulation. Also discussed as examples are relativistic electromagnetic and gravitational interactions, and QLT for gravitational waves is proposed. Overall, our formulation interconnects many known results that in the past were derived independently and reproduces them within a unifying framework.

This progress is made by giving up the traditional Fourier–Laplace approach. The author takes the stance that the global-mode language is unnatural for inhomogeneous-plasma problems (i.e. all real-plasma problems). A fundamental theory must be local. Likewise, the variational approach that is used sometimes in QL calculations is not universally advantageous, especially for describing dissipation. Instead of those methods, we use operator analysis and the Weyl symbol calculus, as has also been proven fruitful in other recent studies of ponderomotive effects and turbulence (Ruiz, 2017; Ruiz & Dodin, 2017b; Zhu & Dodin, 2021) and linear-wave theory (Dodin et al., 2019). No logical leaps are made in this paper other than assuming the QL approximation per se and a certain ordering.222We treat the traditional QL approximation as a given mathematical model. We seek to push this model to its limits rather than to examine its validity, which is a separate issue. For discussions on the validity of the QL approximation, see (Besse et al., 2011; Escande et al., 2018; Crews & Shumlak, 2022). In a nutshell, we treat the commonly known QL-diffusion coefficient as a nonlocal operator, and we systematically approximate it using the Weyl symbol calculus. It is the nonlocality of this operator that gives rise to ponderomotive effects and ensures the proper conservation laws. The existing concept of ‘adiabatic diffusion’ (Galeev & Sagdeev, 1985; Stix, 1992) captures some of that, but systematic application of operator analysis yields a more general, more accurate, and more rigorous theory.

The author hopes not that this paper is an entertaining read. However, the paper was intended as self-contained, maximally structured, and easily searchable, so readers interested in specific questions could find and understand answers without having to read the whole paper. The text is organized as follows. In section 2, we present a primer on the Weyl symbol calculus and the associated notation. In section 3, we formulate our general model. In section 4, we introduce the necessary auxiliary theorems. In section 5, we derive a QL model for plasma interacting with prescribed waves. The waves may or may not be on-shell or self-consistent. (Their origin and dynamics are not addressed in section 5.) In section 6, we consider interactions with self-consistent waves. In particular, we separate out microscopic fluctuations, calculate their average distribution, and derive the corresponding collision operator. In section 7, we assume that the remaining macroscopic waves are on-shell, rederive the WKE, and show that our QL model combined with the WKE is conservative. In section 8, we discuss the general properties that our model predicts for plasmas in thermal equilibrium. In section 9, we show how to apply our theory to nonrelativistic electrostatic interactions, relativistic electromagnetic interactions, Newtonian gravity, and relativistic gravity. In section 10, we summarize our results. Auxiliary calculations are presented in appendices AD, and appendix G summarizes our notation. This notation is extensive and may not be particularly intuitive. Thus, readers are encouraged to occasionally scout section 9 for examples even before fully absorbing the preceding sections.

An impatient reader can also skip calculations entirely and consult only the summaries of the individual sections (2.3, 3.4, 4.4, 5.6, 6.9, 7.6, 8.5; they are mostly self-contained) and then proceed to the examples in section 9. However, the main point of this work is not just the final results per se (surely, some readers will find them obvious) but also that they are derived with minimal assumptions and rigorously, which makes them reliable. A reader may also notice that we rederive some known results along the way, for example, basic linear-wave theory and the WKE. This is done for completeness and, more importantly, with the goal to present all pieces of the story within a unified notation.

2 A math primer

Here, we summarize the machinery to be used in the next sections. This machinery is not new, but a brief overview is in order at least to introduce the necessary notation. A more comprehensive summary, with proofs, can be found in (Dodin et al., 2019, supplementary material). For extended discussions, see (Tracy et al., 2014; Ruiz, 2017; McDonald, 1988; Littlejohn, 1986).

2.1 Weyl symbol calculus on spacetime

2.1.1 Basic notation

We denote the time variable as tt, space coordinates as 𝒙(x1,x2,,xn){\boldsymbol{x}}\equiv(x^{1},x^{2},\ldots,x^{n}), spacetime coordinates as 𝘅(𝗑0,𝗑1,,𝗑n){\boldsymbol{\mathsf{x}}}\equiv(\mathsf{x}^{0},\mathsf{x}^{1},\ldots,\mathsf{x}^{n}), where 𝗑0t\mathsf{x}^{0}\doteq t and 𝗑ixi\mathsf{x}^{i}\doteq x^{i}. The symbol \doteq denotes definitions, and Latin indices from the middle of the alphabet (i,j,i,j,\ldots) range from 1 to nn unless specified otherwise. We assume the spacetime-coordinate domain to be 𝗇\smash{\mathbb{R}^{\mathsf{n}}}.333This excludes periodic boundary conditions, albeit not entirely (section 3.1). Other than that, the spacetime metric can still be non-Euclidean, as illustrated by an application to relativistic gravity in section 9.4. See also the footnotes on pages 5 and 25. Functions on 𝘅{\boldsymbol{\mathsf{x}}} form a Hilbert space 𝗑\smash{\mathscr{H}_{\mathsf{x}}} with an inner product that we define as

ξ|ψd𝘅ξ(𝘅)ψ(𝘅).\braket{\xi}{\psi}\doteq\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\xi^{*}({\boldsymbol{\mathsf{x}}})\psi({\boldsymbol{\mathsf{x}}}). (1)

The symbol denotes complex conjugate,

d𝘅d𝗑0d𝗑1d𝗑n=dtdx1dxn\mathrm{d}{\boldsymbol{\mathsf{x}}}\doteq\mathrm{d}\mathsf{x}^{0}\,\mathrm{d}\mathsf{x}^{1}\ldots\mathrm{d}\mathsf{x}^{n}=\mathrm{d}t\,\mathrm{d}x^{1}\ldots\mathrm{d}x^{n} (2)

(a similar convention is assumed also for other multi-dimensional variables used below), and integrals in this paper are taken over (,)(-\infty,\infty) unless specified otherwise. Operators on 𝗑\mathscr{H}_{\mathsf{x}} will be denoted with carets, and we will use indexes H{}_{\text{H}} and A{}_{\text{A}} to denote their Hermitian and anti-Hermitian parts. For a given operator 𝖠^\smash{\widehat{\mathsf{A}}}, one has 𝖠^=𝖠^H+i𝖠^A\smash{\widehat{\mathsf{A}}=\widehat{\mathsf{A}}_{\text{H}}+\mathrm{i}\widehat{\mathsf{A}}_{\text{A}}},

𝖠^H=𝖠^H12(𝖠^+𝖠^),𝖠^A=𝖠^A12i(𝖠^𝖠^),\widehat{\mathsf{A}}_{\text{H}}=\widehat{\mathsf{A}}_{\text{H}}^{\dagger}\doteq\frac{1}{2}\,(\widehat{\mathsf{A}}+\widehat{\mathsf{A}}^{\dagger}),\qquad\widehat{\mathsf{A}}_{\text{A}}=\widehat{\mathsf{A}}_{\text{A}}^{\dagger}\doteq\frac{1}{2\mathrm{i}}\,(\widehat{\mathsf{A}}-\widehat{\mathsf{A}}^{\dagger}), (3)

where denotes the Hermitian adjoint with respect to the inner product (1). The case of a more general inner product is detailed in (Dodin et al., 2019, supplementary material).

2.1.2 Vector fields

For multi-component fields 𝝍(ψ1,ψ2,,ψM)\smash{{\boldsymbol{\psi}}\equiv(\psi^{1},\psi^{2},\ldots,\psi^{M})^{\intercal}} (our denotes the matrix transpose), or ‘row vectors’ (actually, tuples), we define the dual ‘column vectors’ 𝝍(ψ1,ψ2,,ψM)\smash{{\boldsymbol{\psi}}^{\dagger}\equiv(\psi_{1}^{*},\psi_{2}^{*},\ldots,\psi_{M}^{*})} via 𝝍𝗴𝝍\smash{{\boldsymbol{\psi}}^{\dagger}\doteq{\boldsymbol{\mathsf{g}}}{\boldsymbol{\psi}}^{*}}. The matrix 𝗴\smash{{\boldsymbol{\mathsf{g}}}} is assumed to be real, diagonal, invertible, and constant; other than that, it can be chosen as suits a specific problem. (For example, a unit matrix may suffice.) This induces the standard rule of index manipulation

ψi=𝗀ijψj,ψi=𝗀ijψj,i,j=1,2,,M,\psi_{i}=\mathsf{g}_{ij}\psi^{j},\qquad\psi^{i}=\mathsf{g}^{ij}\psi_{j},\qquad i,j=1,2,\ldots,M, (4)

where 𝗀ij\smash{\mathsf{g}_{ij}} are elements of 𝗴\smash{{\boldsymbol{\mathsf{g}}}} and 𝗀ij\smash{\mathsf{g}^{ij}} are elements of 𝗴1\smash{{\boldsymbol{\mathsf{g}}}^{-1}}. Summation over repeating indices is assumed. The rules of matrix multiplication apply to row and column vectors as usual. Then, for 𝝍(ψ1,ψ2,,ψM)\smash{{\boldsymbol{\psi}}\equiv(\psi^{1},\psi^{2},\ldots,\psi^{M})^{\intercal}} and 𝝃(ξ1,ξ2,,ξM)\smash{{\boldsymbol{\xi}}\equiv(\xi^{1},\xi^{2},\ldots,\xi^{M})^{\intercal}}, the quantity 𝝍𝝃\smash{{\boldsymbol{\psi}}{\boldsymbol{\xi}}} is a matrix with elements ψiξj\smash{\psi^{i}\xi^{j}}, 𝝍𝝃\smash{{\boldsymbol{\psi}}{\boldsymbol{\xi}}^{\dagger}} is a matrix with elements ψiξj\smash{\psi^{i}\xi_{j}^{*}} and 𝝃𝝍\smash{{\boldsymbol{\xi}}^{\dagger}{\boldsymbol{\psi}}} is its scalar trace:444A common notation is 𝝃𝝍=𝝃𝝍\smash{{\boldsymbol{\xi}}^{\dagger}{\boldsymbol{\psi}}={\boldsymbol{\xi}}\cdot{\boldsymbol{\psi}}}, but we reserve the dot-product notation for a scalar product of different quantities (section 2.1.3).

𝝃𝝍=tr(𝝍𝝃)=ξiψi=𝗀ijξiψj,i,j=1,2,,M.{\boldsymbol{\xi}}^{\dagger}{\boldsymbol{\psi}}=\operatorname{tr}({\boldsymbol{\psi}}{\boldsymbol{\xi}}^{\dagger})=\xi_{i}^{*}\psi^{i}=\mathsf{g}_{ij}\xi^{i*}\psi^{j},\qquad i,j=1,2,\ldots,M. (5)

(Similarly, for 𝝌(χ1,χ2,,χM)\smash{{\boldsymbol{\chi}}\equiv(\chi_{1},\chi_{2},\ldots,\chi_{M})} and 𝜼(η1,η2,,ηM)\smash{{\boldsymbol{\eta}}\equiv(\eta_{1},\eta_{2},\ldots,\eta_{M})}, 𝜼𝝌\smash{{\boldsymbol{\eta}}{\boldsymbol{\chi}}} is a matrix with elements ηiχj\smash{\eta_{i}\chi_{j}}.) We use the (5) to define a Hilbert space 𝗑M\smash{\mathscr{H}_{\mathsf{x}}^{M}} of M\smash{M}-dimensional vector fields on 𝘅{\boldsymbol{\mathsf{x}}}, specifically, by adopting the inner product

𝝃|𝝍d𝘅𝝃(𝘅)𝝍(𝘅).\braket{{\boldsymbol{\xi}}}{{\boldsymbol{\psi}}}\doteq\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,{\boldsymbol{\xi}}^{\dagger}({\boldsymbol{\mathsf{x}}}){\boldsymbol{\psi}}({\boldsymbol{\mathsf{x}}}). (6)

Below, the distinction between 𝗑M\mathscr{H}_{\mathsf{x}}^{M} and 𝗑\mathscr{H}_{\mathsf{x}} will be assumed but not emphasized. Also note that (5) yields

𝝃(𝝍𝝍)𝝃=(𝝃𝝍)(𝝍𝝃)=|𝝃𝝍|20{\boldsymbol{\xi}}^{\dagger}({\boldsymbol{\psi}}{\boldsymbol{\psi}}^{\dagger}){\boldsymbol{\xi}}=({\boldsymbol{\xi}}^{\dagger}{\boldsymbol{\psi}})({\boldsymbol{\psi}}^{\dagger}{\boldsymbol{\xi}})=|{\boldsymbol{\xi}}^{\dagger}{\boldsymbol{\psi}}|^{2}\geq 0 (7)

for any 𝝃\smash{{\boldsymbol{\xi}}} and 𝝍\smash{{\boldsymbol{\psi}}}. Thus, dyadic matrices of the form 𝝍𝝍\smash{{\boldsymbol{\psi}}{\boldsymbol{\psi}}^{\dagger}} are positive-semidefinite, even though 𝝍𝝍\smash{{\boldsymbol{\psi}}^{\dagger}{\boldsymbol{\psi}}} may be negative (when 𝗴\smash{{\boldsymbol{\mathsf{g}}}} is not positive-definite).

For general matrices, the indices can be raised and lowered using 𝗴\smash{{\boldsymbol{\mathsf{g}}}} and 𝗴1\smash{{\boldsymbol{\mathsf{g}}}^{-1}} as usual. The Hermitian adjoint 𝗔\smash{{\boldsymbol{\mathsf{A}}}^{\dagger}} for a given matrix 𝗔\smash{{\boldsymbol{\mathsf{A}}}} is defined such that (𝗔𝝃)𝝍=𝝃(𝗔𝝍)\smash{({\boldsymbol{\mathsf{A}}}^{\dagger}{\boldsymbol{\xi}})^{\dagger}{\boldsymbol{\psi}}={\boldsymbol{\xi}}^{\dagger}({\boldsymbol{\mathsf{A}}}{\boldsymbol{\psi}})} for any 𝝍\smash{{\boldsymbol{\psi}}} and 𝝃\smash{{\boldsymbol{\xi}}}, which means

(𝗔)j=i(𝗔)ij𝖠i,ji,j=1,2,,M.({\boldsymbol{\mathsf{A}}}^{\dagger})_{j}{}^{i}=({\boldsymbol{\mathsf{A}}})^{i*}{}_{j}\equiv\mathsf{A}^{i*}{}_{j},\qquad i,j=1,2,\ldots,M. (8)

The Hermitian and anti-Hermitian parts are defined as

𝗔H=𝗔H12(𝗔+𝗔),𝗔A=𝗔A12i(𝗔𝗔),{\boldsymbol{\mathsf{A}}}_{\text{H}}={\boldsymbol{\mathsf{A}}}_{\text{H}}^{\dagger}\doteq\frac{1}{2}\,({\boldsymbol{\mathsf{A}}}+{\boldsymbol{\mathsf{A}}}^{\dagger}),\qquad{\boldsymbol{\mathsf{A}}}_{\text{A}}={\boldsymbol{\mathsf{A}}}_{\text{A}}^{\dagger}\doteq\frac{1}{2\mathrm{i}}\,({\boldsymbol{\mathsf{A}}}-{\boldsymbol{\mathsf{A}}}^{\dagger}), (9)

so 𝗔=𝗔H+i𝗔A\smash{{\boldsymbol{\mathsf{A}}}={\boldsymbol{\mathsf{A}}}_{\text{H}}+\mathrm{i}{\boldsymbol{\mathsf{A}}}_{\text{A}}}. For one-dimensional matrices (scalars), one has 𝗔=𝖠{\boldsymbol{\mathsf{A}}}=\mathsf{A},

𝖠H=𝖠H=re𝖠,𝖠A=𝖠A=im𝖠,\mathsf{A}_{\text{H}}=\mathsf{A}_{\text{H}}^{*}=\operatorname{re}\mathsf{A},\qquad\mathsf{A}_{\text{A}}=\mathsf{A}_{\text{A}}^{*}=\operatorname{im}\mathsf{A}, (10)

where re\smash{\operatorname{re}} and im\smash{\operatorname{im}} denote the real part and the imaginary part, respectively. We also define matrix operators 𝗔^\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}} as matrices of the corresponding operators 𝖠^ij\smash{\widehat{\mathsf{A}}^{i}{}_{j}}. Because 𝗴\smash{{\boldsymbol{\mathsf{g}}}} is constant, index manipulation applies as usual. Also as usual, one has

𝗔^H=𝗔^H12(𝗔^+𝗔^),𝗔^A=𝗔^A12i(𝗔^𝗔^),\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}_{\text{H}}=\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}_{\text{H}}^{\dagger}\doteq\frac{1}{2}\,(\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}+\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}^{\dagger}),\qquad\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}_{\text{A}}=\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}_{\text{A}}^{\dagger}\doteq\frac{1}{2\mathrm{i}}\,(\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}-\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}^{\dagger}), (11)

and 𝗔^=𝗔^H+i𝗔^A\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}=\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}_{\text{H}}+\mathrm{i}\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}_{\text{A}}}, where is the Hermitian adjoint with respect to the inner product (6).

2.1.3 Bra–ket notation

Let us define the following operators that are Hermitian under the inner product (1):

𝗑^0t^t,𝗑^ix^ixi,𝗄^0ω^it,k^iii,\widehat{\mathsf{x}}^{0}\equiv\widehat{t}\doteq t,\quad\widehat{\mathsf{x}}^{i}\equiv\widehat{x}^{i}\doteq x^{i},\quad\widehat{\mathsf{k}}_{0}\equiv-\widehat{\omega}\doteq-\mathrm{i}\partial_{t},\quad\widehat{k}_{i}\doteq-\mathrm{i}\partial_{i}, (12)

where 0t/x0\partial_{0}\equiv\partial_{t}\doteq\partial/\partial x^{0} and i/xi\partial_{i}\doteq\partial/\partial x^{i}. Accordingly,

𝘅^(𝗑^0,𝗑^1,,𝗑^n)=(t^,𝒙^),𝗸^(𝗄^0,𝗄^1,,𝗄^n)=(ω^,𝒌^)\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}}\equiv(\widehat{\mathsf{x}}^{0},\widehat{\mathsf{x}}^{1},\ldots,\widehat{\mathsf{x}}^{n})=(\widehat{t},\widehat{\boldsymbol{x}}),\qquad\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}}\equiv(\widehat{\mathsf{k}}_{0},\widehat{\mathsf{k}}_{1},\ldots,\widehat{\mathsf{k}}_{n})=(-\widehat{\omega},\widehat{\boldsymbol{k}}) (13)

are understood as the spacetime-position operator and the corresponding wavevector operator, which will also be expressed as follows:

𝘅^=𝘅,𝗸^=i𝘅.\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}}={\boldsymbol{\mathsf{x}}},\qquad\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}}=-\mathrm{i}\partial_{\boldsymbol{\mathsf{x}}}. (14)

Also note the commutation property, where δij\smash{\delta_{i}^{j}} is the Kronecker symbol:555Spaces with periodic boundary conditions require a different approach (Rigas et al., 2011), so they are not considered here (yet see section 3.1). That said, for a system that is large enough, the boundary conditions are unimportant; then the toolbox presented here is applicable as is.

[𝗑^i,𝗄^j]=iδji,i,j=0,1,,n.[\widehat{\mathsf{x}}^{i},\widehat{\mathsf{k}}_{j}]=\mathrm{i}\delta^{i}_{j},\qquad i,j=0,1,\ldots,n. (15)

The eigenvectors of the operators (14) will be denoted as ‘kets’ |𝘅\smash{\ket{{\boldsymbol{\mathsf{x}}}}} and |𝗸\smash{\ket{{\boldsymbol{\mathsf{k}}}}}:666More precisely, |𝘅\smash{\ket{{\boldsymbol{\mathsf{x}}}}} is the ket |𝔢(𝘅^;𝘅)\smash{\ket{\mathfrak{e}(\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}};{\boldsymbol{\mathsf{x}}})}} that is an eigenvector of each 𝗑^i\smash{\widehat{\mathsf{x}}^{i}} with the corresponding eigenvalues being 𝗑i\smash{\mathsf{x}^{i}}. A similar comment applies to |𝗸\smash{\ket{{\boldsymbol{\mathsf{k}}}}}.

𝘅^|𝘅=𝘅|𝘅,𝗸^|𝗸=𝗸|𝗸,\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}}\ket{{\boldsymbol{\mathsf{x}}}}={\boldsymbol{\mathsf{x}}}\ket{{\boldsymbol{\mathsf{x}}}},\qquad\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}}\ket{{\boldsymbol{\mathsf{k}}}}={\boldsymbol{\mathsf{k}}}\ket{{\boldsymbol{\mathsf{k}}}}, (16)

and we assume the usual normalization:

𝘅1|𝘅2=δ(𝘅1𝘅2),𝗸1|𝗸2=δ(𝗸1𝗸2),\braket{{\boldsymbol{\mathsf{x}}}_{1}}{{\boldsymbol{\mathsf{x}}}_{2}}=\delta({\boldsymbol{\mathsf{x}}}_{1}-{\boldsymbol{\mathsf{x}}}_{2}),\qquad\braket{{\boldsymbol{\mathsf{k}}}_{1}}{{\boldsymbol{\mathsf{k}}}_{2}}=\delta({\boldsymbol{\mathsf{k}}}_{1}-{\boldsymbol{\mathsf{k}}}_{2}), (17)

where δ\delta is the Dirac delta function. Both sets {|𝘅,𝘅𝗇}\{\ket{{\boldsymbol{\mathsf{x}}}},{\boldsymbol{\mathsf{x}}}\in\mathbb{R}^{\mathsf{n}}\} and {|𝗸,𝗸𝗇}\{\ket{{\boldsymbol{\mathsf{k}}}},{\boldsymbol{\mathsf{k}}}\in\mathbb{R}^{\mathsf{n}}\}, where 𝗇n+1{\mathsf{n}}\doteq n+1, form a complete basis on 𝗑\mathscr{H}_{\mathsf{x}}, and the eigenvalues of these operators form an extended real phase space (𝘅,𝗸)({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}), where

𝘅(t,𝒙),𝗸(ω,𝒌).{\boldsymbol{\mathsf{x}}}\equiv(t,{\boldsymbol{x}}),\qquad{\boldsymbol{\mathsf{k}}}\equiv(-\omega,{\boldsymbol{k}}). (18)

The notation 𝗸𝘀ωτ+𝒌𝒔\smash{{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{s}}}\doteq-\omega\tau+{\boldsymbol{k}}\cdot{\boldsymbol{s}}} will be assumed for any 𝘀(τ,𝒔){\boldsymbol{\mathsf{s}}}\equiv(\tau,{\boldsymbol{s}}), and 𝒌𝒔kisi\smash{{\boldsymbol{k}}\cdot{\boldsymbol{s}}\doteq k_{i}s^{i}}. In particular, for any ψ\psi and constant 𝘀{\boldsymbol{\mathsf{s}}}, one has

exp(i𝗸^𝘀)ψ(𝘅)=exp(𝘀𝘅)ψ(𝘅)=ψ(𝘅+𝘀),\exp(\mathrm{i}\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}}\cdot{\boldsymbol{\mathsf{s}}})\psi({\boldsymbol{\mathsf{x}}})=\exp({\boldsymbol{\mathsf{s}}}\cdot\partial_{\boldsymbol{\mathsf{x}}})\psi({\boldsymbol{\mathsf{x}}})=\psi({\boldsymbol{\mathsf{x}}}+{\boldsymbol{\mathsf{s}}}), (19)

as seen from comparing the Taylor expansions of the latter two expressions. (A generalization of this formula is discussed in section 4.1.) Also,

𝘅|𝗸=𝗸|𝘅=(2\upi)𝗇/2exp(i𝗸𝘅),\braket{{\boldsymbol{\mathsf{x}}}}{{\boldsymbol{\mathsf{k}}}}=\braket{{\boldsymbol{\mathsf{k}}}}{{\boldsymbol{\mathsf{x}}}}^{*}=(2\upi)^{-{\mathsf{n}}/2}\exp(\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{x}}}), (20)

and

d𝘅|𝘅𝘅|=𝟣^,d𝗸|𝗸𝗸|=𝟣^.\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\ket{{\boldsymbol{\mathsf{x}}}}\bra{{\boldsymbol{\mathsf{x}}}}=\widehat{\mathsf{1}},\qquad\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\ket{{\boldsymbol{\mathsf{k}}}}\bra{{\boldsymbol{\mathsf{k}}}}=\widehat{\mathsf{1}}. (21)

Here, ‘bra’ 𝘅|\bra{{\boldsymbol{\mathsf{x}}}} is the one-form dual to |𝘅\ket{{\boldsymbol{\mathsf{x}}}}, 𝗸|\bra{{\boldsymbol{\mathsf{k}}}} is the one-form dual to |𝗸\ket{{\boldsymbol{\mathsf{k}}}}, and 𝟣^\widehat{\mathsf{1}} is the unit operator. Any field ψ\psi on 𝘅{\boldsymbol{\mathsf{x}}} can be viewed as the 𝘅{\boldsymbol{\mathsf{x}}} representation (‘coordinate representation’) of |ψ\ket{\psi}, i.e. the projection of an abstract ket vector |ψ𝗑\ket{\psi}\in\mathscr{H}_{\mathsf{x}} on |𝘅\smash{\ket{{\boldsymbol{\mathsf{x}}}}}:

ψ(𝘅)=𝘅|ψ.\psi({\boldsymbol{\mathsf{x}}})=\braket{{\boldsymbol{\mathsf{x}}}}{\psi}. (22)

Similarly, 𝗸|ψ\braket{{\boldsymbol{\mathsf{k}}}}{\psi} is the 𝗸{\boldsymbol{\mathsf{k}}} representation (‘spectral representation’) of |ψ\ket{\psi}, or the Fourier image of ψ\psi:

ψ̊(𝗸)𝗸|ψ=1(2\upi)𝗇/2d𝘅ei𝗸𝘅ψ(𝘅).\mathring{\psi}({\boldsymbol{\mathsf{k}}})\doteq\braket{{\boldsymbol{\mathsf{k}}}}{\psi}=\frac{1}{(2\upi)^{{\mathsf{n}}/2}}\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\mathrm{e}^{-\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{x}}}}\psi({\boldsymbol{\mathsf{x}}}). (23)

2.1.4 Wigner–Weyl transform

For a given operator 𝖠^\widehat{\mathsf{A}} and a given field ψ\psi, 𝖠^ψ\widehat{\mathsf{A}}\psi can be expressed in the integral form

𝖠^ψ(𝘅)=d𝘅𝘅|𝖠^|𝘅ψ(𝘅),\widehat{\mathsf{A}}\psi({\boldsymbol{\mathsf{x}}})=\int\mathrm{d}{\boldsymbol{\mathsf{x}}}^{\prime}\braket{{\boldsymbol{\mathsf{x}}}}{\widehat{\mathsf{A}}}{{\boldsymbol{\mathsf{x}}}^{\prime}}\psi({\boldsymbol{\mathsf{x}}}^{\prime}), (24)

where 𝘅|𝖠^|𝘅\smash{\braket{{\boldsymbol{\mathsf{x}}}}{\widehat{\mathsf{A}}}{{\boldsymbol{\mathsf{x}}}^{\prime}}} is a function of (𝘅,𝘅)({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{x}}}^{\prime}). This is called the 𝘅{\boldsymbol{\mathsf{x}}} representation (‘coordinate representation’) of 𝖠^\smash{\widehat{\mathsf{A}}}. Equivalently, 𝖠^\widehat{\mathsf{A}} can be given a phase-space, or Weyl, representation, i.e. expressed through a function of the phase-space coordinates, 𝖠(𝘅,𝗸)\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}):777Analytic continuation to complex arguments is possible, but by default, 𝘅\smash{{\boldsymbol{\mathsf{x}}}} and 𝗸\smash{{\boldsymbol{\mathsf{k}}}} are real.

𝖠^=1(2\upi)𝗇d𝘅d𝗸d𝘀|𝘅+𝘀/2𝖠(𝘅,𝗸)ei𝗸𝘀𝘅𝘀/2|oper𝗑𝖠.\widehat{\mathsf{A}}=\frac{1}{(2\upi)^{{\mathsf{n}}}}\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\mathrm{d}{\boldsymbol{\mathsf{s}}}\,\ket{{\boldsymbol{\mathsf{x}}}+{\boldsymbol{\mathsf{s}}}/2}\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\,\mathrm{e}^{\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{s}}}}\bra{{\boldsymbol{\mathsf{x}}}-{\boldsymbol{\mathsf{s}}}/2}\equiv\text{oper}_{\mathsf{x}}\mathsf{A}. (25)

The function 𝖠(𝘅,𝗸)\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}), called the Weyl symbol (or just ‘symbol’) of 𝖠^\widehat{\mathsf{A}}, is given by

𝖠(𝘅,𝗸)d𝘀𝘅+𝘀/2|𝖠^|𝘅𝘀/2ei𝗸𝘀symb𝗑𝖠^.\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\doteq\int\mathrm{d}{\boldsymbol{\mathsf{s}}}\braket{{\boldsymbol{\mathsf{x}}}+{\boldsymbol{\mathsf{s}}}/2}{\widehat{\mathsf{A}}}{{\boldsymbol{\mathsf{x}}}-{\boldsymbol{\mathsf{s}}}/2}\mathrm{e}^{-\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{s}}}}\equiv\text{symb}_{\mathsf{x}}\widehat{\mathsf{A}}. (26)

The 𝘅{\boldsymbol{\mathsf{x}}} and phase-space representations are connected by the Fourier transform:

𝘅|𝖠^|𝘅=1(2\upi)𝗇d𝗸ei𝗸(𝘅𝘅)𝖠(𝘅+𝘅2,𝗸).\displaystyle\braket{{\boldsymbol{\mathsf{x}}}}{\widehat{\mathsf{A}}}{{\boldsymbol{\mathsf{x}}}^{\prime}}=\frac{1}{(2\upi)^{\mathsf{n}}}\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\mathrm{e}^{\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot({\boldsymbol{\mathsf{x}}}-{\boldsymbol{\mathsf{x}}}^{\prime})}\,\mathsf{A}\left(\frac{{\boldsymbol{\mathsf{x}}}+{\boldsymbol{\mathsf{x}}}^{\prime}}{2},{\boldsymbol{\mathsf{k}}}\right). (27)

This also leads to the following notable properties of Weyl symbols:

𝘅|𝖠^|𝘅=1(2\upi)𝗇d𝗸𝖠(𝘅,𝗸),𝗸|𝖠^|𝗸=1(2\upi)𝗇d𝘅𝖠(𝘅,𝗸).\displaystyle\braket{{\boldsymbol{\mathsf{x}}}}{\widehat{\mathsf{A}}}{{\boldsymbol{\mathsf{x}}}}=\frac{1}{(2\upi)^{{\mathsf{n}}}}\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}),\qquad\braket{{\boldsymbol{\mathsf{k}}}}{\widehat{\mathsf{A}}}{{\boldsymbol{\mathsf{k}}}}=\frac{1}{(2\upi)^{{\mathsf{n}}}}\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}). (28)

An operator unambiguously determines its symbol, and vice versa. We denote this isomorphism as 𝖠^𝖠\smash{\widehat{\mathsf{A}}\leftrightarrow\mathsf{A}}. The mapping 𝖠^𝖠\smash{\widehat{\mathsf{A}}\mapsto\mathsf{A}} is called the Wigner transform, and 𝖠𝖠^\smash{\mathsf{A}\mapsto\widehat{\mathsf{A}}} is called the Weyl transform. For uniformity, we call them the direct and inverse Wigner–Weyl transform. The isomorphism \smash{\leftrightarrow} is natural in that it has the following properties:

𝟣^1,𝘅^𝘅,𝗸^𝗸,h(𝘅^)h(𝘅),h(𝗸^)h(𝗸),𝖠^𝖠,\displaystyle\widehat{\mathsf{1}}\leftrightarrow 1,\quad\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}}\leftrightarrow{\boldsymbol{\mathsf{x}}},\quad\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}}\leftrightarrow{\boldsymbol{\mathsf{k}}},\quad h(\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}})\leftrightarrow h({\boldsymbol{\mathsf{x}}}),\quad h(\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}})\leftrightarrow h({\boldsymbol{\mathsf{k}}}),\quad\widehat{\mathsf{A}}^{\dagger}\leftrightarrow\mathsf{A}^{*}, (29)

where hh is any function and 𝖠^\smash{\widehat{\mathsf{A}}} is any operator. The product of two operators maps to the so-called Moyal product, or star product, of their symbols (Moyal, 1949):

𝖠^𝖡^𝖠(𝘅,𝗸)𝖡(𝘅,𝗸)𝖠(𝘅,𝗸)ei^𝗑/2𝖡(𝘅,𝗸),\displaystyle\widehat{\mathsf{A}}\widehat{\mathsf{B}}\,\leftrightarrow\,\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\star\mathsf{B}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\doteq\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\mathrm{e}^{\mathrm{i}\widehat{\mathcal{L}}_{\mathsf{x}}/2}\mathsf{B}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}), (30)

which is associative:

𝖠^𝖡^𝖢^(𝖠𝖡)𝖢=𝖠(𝖡𝖢)𝖠𝖡𝖢.\displaystyle\widehat{\mathsf{A}}\widehat{\mathsf{B}}\widehat{\mathsf{C}}\,\leftrightarrow\,(\mathsf{A}\star\mathsf{B})\star\mathsf{C}=\mathsf{A}\star(\mathsf{B}\star\mathsf{C})\equiv\mathsf{A}\star\mathsf{B}\star\mathsf{C}. (31)

Here, ^𝗑𝘅𝗸𝗸𝘅\smash{\widehat{\mathcal{L}}_{\mathsf{x}}\doteq\overset{{\scriptscriptstyle\leftarrow}}{\partial}_{\boldsymbol{\mathsf{x}}}\cdot\overset{{\scriptscriptstyle\rightarrow}}{\partial}_{\boldsymbol{\mathsf{k}}}-\overset{{\scriptscriptstyle\leftarrow}}{\partial}_{\boldsymbol{\mathsf{k}}}\cdot\overset{{\scriptscriptstyle\rightarrow}}{\partial}_{\boldsymbol{\mathsf{x}}}}, and the arrows indicate the directions in which the derivatives act. For example, 𝖠^𝗑𝖡\smash{\mathsf{A}\widehat{\mathcal{L}}_{\mathsf{x}}\mathsf{B}} is just the canonical Poisson bracket on (𝘅,𝗸)\smash{({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})}:

𝖠^𝗑𝖡={𝖠,𝖡}𝗑𝖠t𝖡ω+𝖠ω𝖡t+𝖠xi𝖡ki𝖠ki𝖡xi.\mathsf{A}\widehat{\mathcal{L}}_{\mathsf{x}}\mathsf{B}=\{\mathsf{A},\mathsf{B}\}_{\mathsf{x}}\doteq-\frac{\partial\mathsf{A}}{\partial t}\frac{\partial\mathsf{B}}{\partial\omega}+\frac{\partial\mathsf{A}}{\partial\omega}\frac{\partial\mathsf{B}}{\partial t}+\frac{\partial\mathsf{A}}{\partial x^{i}}\frac{\partial\mathsf{B}}{\partial k_{i}}-\frac{\partial\mathsf{A}}{\partial k_{i}}\frac{\partial\mathsf{B}}{\partial x^{i}}. (32)

These formulas readily yield

h(𝘅^)𝗄^α𝗄αh(𝘅)+i2αh(𝘅),𝗄^αh(𝘅^)𝗄αh(𝘅)i2αh(𝘅),\displaystyle h(\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}})\widehat{\mathsf{k}}_{\alpha}\leftrightarrow\mathsf{k}_{\alpha}h({\boldsymbol{\mathsf{x}}})+\frac{\mathrm{i}}{2}\,\partial_{\alpha}h({\boldsymbol{\mathsf{x}}}),\qquad\widehat{\mathsf{k}}_{\alpha}h(\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}})\leftrightarrow\mathsf{k}_{\alpha}h({\boldsymbol{\mathsf{x}}})-\frac{\mathrm{i}}{2}\,\partial_{\alpha}h({\boldsymbol{\mathsf{x}}}), (33)

also h(𝗸^)ei𝗞𝘅^h(𝗸)ei𝗞𝘅=h(𝗸+𝗞/2)ei𝗞𝘅h(\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}})\mathrm{e}^{\mathrm{i}{\boldsymbol{\mathsf{K}}}\cdot\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}}}\leftrightarrow h({\boldsymbol{\mathsf{k}}})\star\mathrm{e}^{\mathrm{i}{\boldsymbol{\mathsf{K}}}\cdot{\boldsymbol{\mathsf{x}}}}=h({\boldsymbol{\mathsf{k}}}+{\boldsymbol{\mathsf{K}}}/2)\mathrm{e}^{\mathrm{i}{\boldsymbol{\mathsf{K}}}\cdot{\boldsymbol{\mathsf{x}}}}, etc. Another notable formula to be used below, which flows from (28) and (31), is

𝘅|𝖠^𝖡^𝖢^|𝘅=1(2\upi)𝗇d𝗸(𝖠𝖡𝖢)(𝘅,𝗸).\displaystyle\braket{{\boldsymbol{\mathsf{x}}}}{\widehat{\mathsf{A}}\widehat{\mathsf{B}}\widehat{\mathsf{C}}}{{\boldsymbol{\mathsf{x}}}}=\frac{1}{(2\upi)^{{\mathsf{n}}}}\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,(\mathsf{A}\star\mathsf{B}\star\mathsf{C})({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}). (34)

The Moyal product is particularly handy when 𝘅𝗸ϵ1\partial_{{\boldsymbol{\mathsf{x}}}}\partial_{{\boldsymbol{\mathsf{k}}}}\sim\epsilon\ll 1. Such ϵ\epsilon is often called the geometrical-optics parameter. Since ^𝗑=𝒪(ϵ)\widehat{\mathcal{L}}_{\mathsf{x}}=\mathcal{O}(\epsilon), one can express the Moyal product as an asymptotic series in powers of ϵ\epsilon:

=𝟣^+i^𝗑/2^𝗑2/8+\star=\widehat{\mathsf{1}}+\mathrm{i}\widehat{\mathcal{L}}_{\mathsf{x}}/2-\widehat{\mathcal{L}}_{\mathsf{x}}^{2}/8+\ldots (35)

2.1.5 Weyl expansion of operators

Operators can be approximated by approximating their symbols (Dodin et al., 2019; McDonald, 1988). If 𝖠^\smash{\widehat{\mathsf{A}}} is approximately local in 𝘅{\boldsymbol{\mathsf{x}}} (i.e. if 𝖠^ψ(𝘅)\smash{\widehat{\mathsf{A}}\psi({\boldsymbol{\mathsf{x}}})} is determined by values ψ(𝘅+𝘀)\psi({\boldsymbol{\mathsf{x}}}+{\boldsymbol{\mathsf{s}}}) only with small enough 𝘀{\boldsymbol{\mathsf{s}}}), its symbol can be adequately represented by the first few terms of the Taylor expansion in 𝗸{\boldsymbol{\mathsf{k}}}:

𝖠(𝘅,𝗸)=𝖠(𝘅,𝟬)+𝚯0(𝘅)𝗸+,𝚯0(𝘅)(𝗸𝖠(𝘅,𝗸))𝗸=𝟬.\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})=\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{0}}})+{\boldsymbol{\Uptheta}}_{0}({\boldsymbol{\mathsf{x}}})\cdot{\boldsymbol{\mathsf{k}}}+\ldots,\qquad{\boldsymbol{\Uptheta}}_{0}({\boldsymbol{\mathsf{x}}})\doteq(\partial_{\boldsymbol{\mathsf{k}}}\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}))_{{\boldsymbol{\mathsf{k}}}={\boldsymbol{\mathsf{0}}}}. (36)

Application of oper𝗑\smash{\text{oper}_{\mathsf{x}}} to this formula leads to

𝖠^𝖠(𝘅^,𝟬)+12(𝚯^0𝗸^+𝗸^𝚯^0)+,\widehat{\mathsf{A}}\approx\mathsf{A}(\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}},{\boldsymbol{\mathsf{0}}})+\frac{1}{2}\,(\widehat{\boldsymbol{\Uptheta}}_{0}\cdot\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}}+\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}}\cdot\widehat{\boldsymbol{\Uptheta}}_{0})+\ldots, (37)

where 𝚯^0𝚯0(𝘅^)\widehat{\boldsymbol{\Uptheta}}_{0}\doteq{\boldsymbol{\Uptheta}}_{0}(\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}}). One can also rewrite (37) using the commutation property

[𝗸^,𝚯^0]=i(𝘅𝚯0)(𝘅^).[\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}},\widehat{\boldsymbol{\Uptheta}}_{0}]=-\mathrm{i}(\partial_{{\boldsymbol{\mathsf{x}}}}\cdot{\boldsymbol{\Uptheta}}_{0})(\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}}). (38)

In the 𝘅{\boldsymbol{\mathsf{x}}} representation, this leads to

𝖠^=𝖠(𝘅,𝟬)i𝚯0(𝘅)𝘅i2(𝘅𝚯0(𝘅))+\widehat{\mathsf{A}}=\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{0}}})-\mathrm{i}{\boldsymbol{\Uptheta}}_{0}({\boldsymbol{\mathsf{x}}})\cdot\partial_{{\boldsymbol{\mathsf{x}}}}-\frac{\mathrm{i}}{2}\,(\partial_{{\boldsymbol{\mathsf{x}}}}\cdot{\boldsymbol{\Uptheta}}_{0}({\boldsymbol{\mathsf{x}}}))+\ldots (39)

The effect of a nonlocal operator on eikonal (monochromatic or quasimonochromatic) fields can be approximated similarly. Suppose ψ=eiθψ˘\psi=\mathrm{e}^{\mathrm{i}\theta}{\breve{\psi}}, where the dependence of 𝗸¯𝘅θ\overline{{\boldsymbol{\mathsf{k}}}}\doteq\partial_{{\boldsymbol{\mathsf{x}}}}\theta and ψ˘{\breve{\psi}} on 𝘅{\boldsymbol{\mathsf{x}}} is slower than that of θ\theta by factor ϵ1\epsilon\ll 1. Then, 𝖠^ψ=eiθ𝖠^ψ˘\widehat{\mathsf{A}}\psi=\mathrm{e}^{\mathrm{i}\theta}\widehat{\mathsf{A}}^{\prime}{\breve{\psi}}, where 𝖠^eiθ(𝗑^)𝖠^eiθ(𝗑^)\widehat{\mathsf{A}}^{\prime}\doteq\mathrm{e}^{-\mathrm{i}\theta(\widehat{\mathsf{x}})}\widehat{\mathsf{A}}\mathrm{e}^{\mathrm{i}\theta(\widehat{\mathsf{x}})}, and the symbol of 𝖠^\widehat{\mathsf{A}}^{\prime} can be approximated as follows:

𝖠(𝘅,𝗸)=𝖠(𝘅,𝗸¯(𝘅)+𝗸)+𝒪(ϵ2).\mathsf{A}^{\prime}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})=\mathsf{A}({\boldsymbol{\mathsf{x}}},\overline{{\boldsymbol{\mathsf{k}}}}({\boldsymbol{\mathsf{x}}})+{\boldsymbol{\mathsf{k}}})+\mathcal{O}(\epsilon^{2}). (40)

By expanding this in 𝗸{\boldsymbol{\mathsf{k}}} and applying oper𝗑\smash{\text{oper}_{\mathsf{x}}}, one obtains

𝖠^=𝖠(𝘅,𝗸¯(𝘅))i𝚯(𝘅)𝘅i2(𝘅𝚯(𝘅))+𝒪(ϵ2),\displaystyle\widehat{\mathsf{A}}^{\prime}=\mathsf{A}({\boldsymbol{\mathsf{x}}},\overline{{\boldsymbol{\mathsf{k}}}}({\boldsymbol{\mathsf{x}}}))-\mathrm{i}{\boldsymbol{\Uptheta}}({\boldsymbol{\mathsf{x}}})\cdot\partial_{{\boldsymbol{\mathsf{x}}}}-\frac{\mathrm{i}}{2}\,(\partial_{{\boldsymbol{\mathsf{x}}}}\cdot{\boldsymbol{\Uptheta}}({\boldsymbol{\mathsf{x}}}))+\mathcal{O}(\epsilon^{2}), (41)

where 𝚯(𝘅)(𝗸𝖠(𝘅,𝗸))𝗸=𝗸¯(𝘅){\boldsymbol{\Uptheta}}({\boldsymbol{\mathsf{x}}})\doteq(\partial_{\boldsymbol{\mathsf{k}}}\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}))_{{\boldsymbol{\mathsf{k}}}=\overline{{\boldsymbol{\mathsf{k}}}}({\boldsymbol{\mathsf{x}}})}. Neglecting the 𝒪(ϵ2)\mathcal{O}(\epsilon^{2}) corrections in this formula leads to what is commonly known as the geometrical-optics approximation (Dodin et al., 2019).

2.1.6 Wigner functions

Any ket |ψ\ket{\psi} generates a dyadic |ψψ|\ket{\psi}\bra{\psi}. In quantum mechanics, such dyadics are known as density operators (of pure states). For our purposes, though, it is more convenient to define the density operator in a slightly different form, namely, as

𝖶^ψ(2\upi)𝗇|ψψ|.\widehat{\mathsf{W}}_{\psi}\doteq(2\upi)^{-{\mathsf{n}}}\ket{\psi}\bra{\psi}. (42)

The symbol of this operator, 𝖶ψ=symb𝗑𝖶^ψ\smash{\mathsf{W}_{\psi}=\text{symb}_{\mathsf{x}}\widehat{\mathsf{W}}_{\psi}}, is a real function called the Wigner function. It is given by

𝖶ψ(𝘅,𝗸)\displaystyle\mathsf{W}_{\psi}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}) =1(2\upi)𝗇d𝘀𝘅+𝘀/2|ψψ|𝘅𝘀/2ei𝗸𝘀\displaystyle=\frac{1}{(2\upi)^{\mathsf{n}}}\int\mathrm{d}{\boldsymbol{\mathsf{s}}}\braket{{\boldsymbol{\mathsf{x}}}+{\boldsymbol{\mathsf{s}}}/2}{\psi}\braket{\psi}{{\boldsymbol{\mathsf{x}}}-{\boldsymbol{\mathsf{s}}}/2}\mathrm{e}^{-\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{s}}}}
=1(2\upi)𝗇d𝘀ψ(𝘅+𝘀/2)ψ(𝘅𝘀/2)ei𝗸𝘀,\displaystyle=\frac{1}{(2\upi)^{\mathsf{n}}}\int\mathrm{d}{\boldsymbol{\mathsf{s}}}\,\psi({\boldsymbol{\mathsf{x}}}+{\boldsymbol{\mathsf{s}}}/2)\psi^{*}({\boldsymbol{\mathsf{x}}}-{\boldsymbol{\mathsf{s}}}/2)\,\mathrm{e}^{-\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{s}}}}, (43)

which is manifestly real and can be understood as the (inverse) Fourier image of

𝖢ψ(𝘅,𝘀)ψ(𝘅+𝘀/2)ψ(𝘅𝘀/2)=d𝗸𝖶ψ(𝘅,𝗸)ei𝗸𝘀.\displaystyle\mathsf{C}_{\psi}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{s}}})\doteq\psi({\boldsymbol{\mathsf{x}}}+{\boldsymbol{\mathsf{s}}}/2)\psi^{*}({\boldsymbol{\mathsf{x}}}-{\boldsymbol{\mathsf{s}}}/2)=\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\mathsf{W}_{\psi}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\,\mathrm{e}^{\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{s}}}}. (44)

Any function bilinear in ψ\smash{\psi} and ψ\smash{\psi^{*}} can be expressed through 𝖶ψ\smash{\mathsf{W}_{\psi}}. Specifically, for any operators 𝖫^\smash{\widehat{\mathsf{L}}} and 𝖱^\smash{\widehat{\mathsf{R}}}, one has

(𝖫^ψ(𝘅))(𝖱^ψ(𝘅))\displaystyle(\widehat{\mathsf{L}}\psi({\boldsymbol{\mathsf{x}}}))(\widehat{\mathsf{R}}\psi({\boldsymbol{\mathsf{x}}}))^{*} =𝘅|𝖫^|ψψ|𝖱^|𝘅\displaystyle=\braket{{\boldsymbol{\mathsf{x}}}}{\widehat{\mathsf{L}}}{\psi}\braket{\psi}{\widehat{\mathsf{R}}^{\dagger}}{{\boldsymbol{\mathsf{x}}}}
=(2\upi)𝗇𝘅|𝖫^𝖶^ψ𝖱^|𝘅\displaystyle=(2\upi)^{\mathsf{n}}\braket{{\boldsymbol{\mathsf{x}}}}{\widehat{\mathsf{L}}\widehat{\mathsf{W}}_{\psi}\widehat{\mathsf{R}}^{\dagger}}{{\boldsymbol{\mathsf{x}}}}
=d𝗸𝖫(𝘅,𝗸)𝖶ψ(𝘅,𝗸)𝖱(𝘅,𝗸),\displaystyle=\textstyle\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\mathsf{L}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\star\mathsf{W}_{\psi}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\star\mathsf{R}^{*}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}), (45)

where 𝖫\mathsf{L} and 𝖱\mathsf{R} are the corresponding symbols and (28) was used along with (31). As a corollary, and as also seen from (28), one has

|ψ(𝘅)|2=d𝗸𝖶ψ(𝘅,𝗸),|ψ̊(𝗸)|2=d𝘅𝖶ψ(𝘅,𝗸).|\psi({\boldsymbol{\mathsf{x}}})|^{2}=\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\mathsf{W}_{\psi}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}),\qquad|\mathring{\psi}({\boldsymbol{\mathsf{k}}})|^{2}=\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\mathsf{W}_{\psi}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}). (46)

As a reminder, ψ(𝘅)=𝘅|ψ\psi({\boldsymbol{\mathsf{x}}})=\braket{{\boldsymbol{\mathsf{x}}}}{\psi} and ψ̊(𝗸)𝗸|ψ\mathring{\psi}({\boldsymbol{\mathsf{k}}})\doteq\braket{{\boldsymbol{\mathsf{k}}}}{\psi} is the Fourier image of ψ\psi (23), so |ψ(𝘅)|2\smash{|\psi({\boldsymbol{\mathsf{x}}})|^{2}} and |ψ̊(𝗸)|2\smash{|\mathring{\psi}({\boldsymbol{\mathsf{k}}})|^{2}} can be loosely understood as the densities of quanta (associated with the field ψ\smash{\psi}) in the 𝘅\smash{{\boldsymbol{\mathsf{x}}}}-space and the 𝗸\smash{{\boldsymbol{\mathsf{k}}}}-space, respectively. Because of (46), 𝖶ψ\smash{\mathsf{W}_{\psi}} is commonly attributed as a quasiprobability distribution of wave quanta in phase space. (The prefix ‘quasi’ is added because 𝖶ψ\smash{\mathsf{W}_{\psi}} can be negative.) In case of real fields, which satisfy 𝘅|ψ=ψ|𝘅\braket{{\boldsymbol{\mathsf{x}}}}{\psi}=\braket{\psi}{{\boldsymbol{\mathsf{x}}}}, one also has

𝖶ψ(𝘅,𝗸)=𝖶ψ(𝘅,𝗸).\mathsf{W}_{\psi}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})=\mathsf{W}_{\psi}({\boldsymbol{\mathsf{x}}},-{\boldsymbol{\mathsf{k}}}). (47)

Of particular importance are Wigner functions averaged over a sufficiently large phase-space volume Δ𝘅Δ𝗸1\Delta{\boldsymbol{\mathsf{x}}}\,\Delta{\boldsymbol{\mathsf{k}}}\gtrsim 1. The average Wigner function 𝖶¯ψ\smash{\overline{\mathsf{W}}_{\psi}} is a local property of the field (as opposed to, say, the field’s global Fourier spectrum) and satisfies (appendix A)

𝖶¯ψ0.\overline{\mathsf{W}}_{\psi}\geq 0. (48)

2.1.7 Generalization to vector fields

In case of vector (tuple) fields 𝝍=(ψ1,ψ2,,ψM){\boldsymbol{\psi}}=(\psi^{1},\psi^{2},\ldots,\psi^{M})^{\intercal}, kets are column vectors, |𝝍=(|ψ1,|ψ2,,|ψM)\ket{{\boldsymbol{\psi}}}=(\ket{\psi^{1}},\ket{\psi^{2}},\ldots,\ket{\psi^{M}}), and bras are row vectors, 𝝍|=(ψ1|,ψ2|,,ψM|)\bra{{\boldsymbol{\psi}}}=(\bra{\psi^{1}},\bra{\psi^{2}},\ldots,\bra{\psi^{M}}). The operators acting on such kets and bras are matrices of operators. The Weyl symbol of a matrix operator is defined as the matrix of the corresponding symbols. As a result, the symbol of a Hermitian adjoint of a given operator is the Hermitian adjoint of the symbol of that operator:

𝗔^𝗔,\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{A}}}}}}^{\dagger}\leftrightarrow\smash{{\boldsymbol{\mathsf{A}}}}^{\dagger}, (49)

and as a corollary, the symbol of a Hermitian matrix operator is a Hermitian matrix.

In particular, the density operator of a given vector field 𝝍{\boldsymbol{\psi}} is a matrix operator

𝗪^𝝍(2\upi)𝗇|𝝍𝝍|.\widehat{\boldsymbol{{\boldsymbol{\mathsf{W}}}}}_{{\boldsymbol{\psi}}}\doteq(2\upi)^{-{\mathsf{n}}}\ket{{\boldsymbol{\psi}}}\bra{{\boldsymbol{\psi}}}. (50)

The symbol of this operator, 𝗪𝝍=symb𝗑𝗪^𝝍\smash{{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}=\text{symb}_{\mathsf{x}}\widehat{\boldsymbol{{\boldsymbol{\mathsf{W}}}}}_{{\boldsymbol{\psi}}}}, is a Hermitian matrix function888By construction, 𝗪^𝝍\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{W}}}}}_{{\boldsymbol{\psi}}}} is a matrix with mixed indices, (𝗪^𝝍)ij\smash{(\widehat{\boldsymbol{{\boldsymbol{\mathsf{W}}}}}_{{\boldsymbol{\psi}}})^{i}{}_{j}}. In sections 5.1 and 5.2, we also operate with a Wigner matrix that has two upper indices. Because the field of interest is real there, the dagger in (51) is assumed omitted in that case.

𝗪𝝍(𝘅,𝗸)=1(2\upi)𝗇d𝘀𝝍(𝘅+𝘀/2)𝝍(𝘅𝘀/2)ei𝗸𝘀,{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})=\frac{1}{(2\upi)^{\mathsf{n}}}\int\mathrm{d}{\boldsymbol{\mathsf{s}}}\,{\boldsymbol{\psi}}({\boldsymbol{\mathsf{x}}}+{\boldsymbol{\mathsf{s}}}/2){\boldsymbol{\psi}}^{\dagger}({\boldsymbol{\mathsf{x}}}-{\boldsymbol{\mathsf{s}}}/2)\,\mathrm{e}^{-\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{s}}}}, (51)

called the Wigner matrix. (It is also called the ‘Wigner tensor’ when 𝝍\smash{{\boldsymbol{\psi}}} is a true vector rather than a tuple.) It can be understood as the (inverse) Fourier image of

𝗖𝝍(𝘅,𝘀)𝝍(𝘅+𝘀/2)𝝍(𝘅𝘀/2)=d𝗸𝗪𝝍(𝘅,𝗸)ei𝗸𝘀.\displaystyle{\boldsymbol{\mathsf{C}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{s}}})\doteq{\boldsymbol{\psi}}({\boldsymbol{\mathsf{x}}}+{\boldsymbol{\mathsf{s}}}/2){\boldsymbol{\psi}}^{\dagger}({\boldsymbol{\mathsf{x}}}-{\boldsymbol{\mathsf{s}}}/2)=\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\,\mathrm{e}^{\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{s}}}}. (52)

The analog of (45) is (appendix B.1)

(𝗟^𝝍(𝘅))(𝗥^𝝍(𝘅))=d𝗸𝗟(𝘅,𝗸)𝗪𝝍(𝘅,𝗸)𝗥(𝘅,𝗸).\displaystyle(\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}{\boldsymbol{\psi}}({\boldsymbol{\mathsf{x}}}))(\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}{\boldsymbol{\psi}}({\boldsymbol{\mathsf{x}}}))^{\dagger}=\textstyle\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,{\boldsymbol{\mathsf{L}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\star{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\star{\boldsymbol{\mathsf{R}}}^{\dagger}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}). (53a)
The Wigner matrix averaged over a sufficiently large phase-space volume Δ𝘅Δ𝗸1\Delta{\boldsymbol{\mathsf{x}}}\,\Delta{\boldsymbol{\mathsf{k}}}\gtrsim 1 is a local property of the field, and it is positive-semidefinite (appendix A).

For real fields, one also has

𝗪𝝍(𝘅,𝗸)=𝗪𝝍(𝘅,𝗸)=𝗪𝝍(𝘅,𝗸),{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})={\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}^{\intercal}({\boldsymbol{\mathsf{x}}},-{\boldsymbol{\mathsf{k}}})={\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}^{*}({\boldsymbol{\mathsf{x}}},-{\boldsymbol{\mathsf{k}}}), (53b)

and (53) yields the following corollary at ϵ0\smash{\epsilon\to 0}, when \smash{\star} becomes the usual product (appendix B.1):

(𝗟^𝝍)𝗥^𝝍=d𝗸tr(𝗪𝝍(𝗟𝗥)H).\displaystyle(\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}{\boldsymbol{\psi}})^{\dagger}\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}{\boldsymbol{\psi}}=\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\operatorname{tr}\big{(}{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{L}}}^{\dagger}{\boldsymbol{\mathsf{R}}})_{\text{H}}\big{)}. (53c)

The generalizations of the other formulas from the previous sections are obvious.

2.2 Weyl symbol calculus on phase space

2.2.1 Notation

Consider a Hamiltonian system with coordinates 𝒙(x1,x2,,xn){\boldsymbol{x}}\equiv(x^{1},x^{2},\ldots,x^{n}) and canonical momenta 𝒑(p1,p2,,pn){\boldsymbol{p}}\equiv(p_{1},p_{2},\ldots,p_{n}). Together, these variables comprise the phase-space coordinates 𝒛(𝒙,𝒑){\boldsymbol{z}}\equiv({\boldsymbol{x}},{\boldsymbol{p}}), i.e.

𝒛(z1,,z2n)=(x1,,xn,p1,,pn).{\boldsymbol{z}}\equiv(z^{1},\ldots,z^{2n})=(x^{1},\ldots,x^{n},p_{1},\ldots,p_{n}). (54)

Components of 𝒛{\boldsymbol{z}} will be denoted with Greek indices ranging from 1 to 2n2n.999However, the index σ\smash{\sigma} is reserved as a tag for individual particles and waves.

Hamilton’s equations for zαz^{\alpha} can be written as z˙α={zα,H}\dot{z}^{\alpha}=\{z^{\alpha},H\}, or equivalently, as

z˙α=JαββH.\dot{z}^{\alpha}=J^{\alpha\beta}\,\partial_{\beta}H. (55)

Here, H=H(t,𝒛)H=H(t,{\boldsymbol{z}}) is a Hamiltonian, β/zβ\partial_{\beta}\doteq\partial/\partial z^{\beta},

{A,B}Jαβ(αA)(βB)\{A,B\}\doteq J^{\alpha\beta}\,(\partial_{\alpha}A)(\partial_{\beta}B) (56)

is the Poisson bracket on 𝒛\smash{{\boldsymbol{z}}}, JαβJ^{\alpha\beta} is the canonical Poisson structure:

𝑱=𝑱=(𝟎n𝟏n𝟏n𝟎n),{\boldsymbol{J}}=-{\boldsymbol{J}}^{\intercal}=\left(\begin{array}[]{cc}{\boldsymbol{0}}_{n}&{\boldsymbol{1}}_{n}\\ -{\boldsymbol{1}}_{n}&{\boldsymbol{0}}_{n}\\ \end{array}\right), (57)

𝟎n{\boldsymbol{0}}_{n} is an nn-dimensional zero matrix, and 𝟏n{\boldsymbol{1}}_{n} is an nn-dimensional unit matrix. The corresponding equation for the probability distribution f(t,𝒛)f(t,{\boldsymbol{z}}) is

tf={H,f}.\partial_{t}f=\{H,f\}. (58)

Solutions of (58) and other functions of the extended-phase-space coordinates 𝑿(t,𝒛){\boldsymbol{X}}\equiv(t,{\boldsymbol{z}}) can be considered as vectors in the Hilbert space X\mathscr{H}_{X} with the usual inner product101010Note that the inner product (59) is different from (1). Still, we use the same notation assuming it will be clear from the context which inner product is used in each given case.

ξ|ψd𝑿ξ(𝑿)ψ(𝑿).\braket{\xi}{\psi}\doteq\int\mathrm{d}{\boldsymbol{X}}\,\xi^{*}({\boldsymbol{X}})\psi({\boldsymbol{X}}). (59)

Assuming the notation Ndim𝑿=2n+1N\doteq\dim{\boldsymbol{X}}=2n+1, one has

d𝑿dX1dX2dXN=dtdx1dxndp1,,dpn.\mathrm{d}{\boldsymbol{X}}\doteq\mathrm{d}X^{1}\,\mathrm{d}X^{2}\ldots\mathrm{d}X^{N}=\mathrm{d}t\,\mathrm{d}x^{1}\ldots\mathrm{d}x^{n}\,\mathrm{d}p_{1},\ldots,\mathrm{d}p_{n}. (60)

Let us introduce the position operator on 𝒛{\boldsymbol{z}},

𝒛^(x1,,xn𝒙^,p1,,pn𝒑^),\widehat{\boldsymbol{z}}\doteq(\underbrace{x^{1},\ldots,x^{n}}_{\widehat{\boldsymbol{x}}},\underbrace{p_{1},\ldots,p_{n}}_{\widehat{\boldsymbol{p}}}), (61)

and the momentum operator on 𝒛{\boldsymbol{z}},

𝒒^(i1,,in𝒌^,i1,,in𝒓^),\widehat{\boldsymbol{q}}\equiv(\underbrace{-\mathrm{i}\partial_{1},\ldots,-\mathrm{i}\partial_{n}}_{\widehat{\boldsymbol{k}}},\underbrace{-\mathrm{i}\partial^{1},\ldots,-\mathrm{i}\partial^{n}}_{\widehat{\boldsymbol{r}}}), (62)

where i/xi\partial_{i}\doteq\partial/\partial x^{i} but i/pi\partial^{i}\doteq\partial/\partial p_{i}; that is, 𝒛^=(𝒙^,𝒑^)\widehat{\boldsymbol{z}}=(\widehat{\boldsymbol{x}},\widehat{\boldsymbol{p}}), 𝒒^=(𝒌^,𝒓^)\widehat{\boldsymbol{q}}=(\widehat{\boldsymbol{k}},\widehat{\boldsymbol{r}}), and

z^αzα,q^αiα.\widehat{z}^{\alpha}\doteq z^{\alpha},\qquad\widehat{q}_{\alpha}\doteq-\mathrm{i}\partial_{\alpha}. (63)

Then, much like in section 2.1, one can also introduce the position and momentum operators on the extended phase space 𝑿{\boldsymbol{X}}:

𝑿^=(t^,𝒛^)=(t^,𝒙^,𝒑^),𝑲^=(ω^,𝒒^)=(ω^,𝒌^,𝒓^).\widehat{\boldsymbol{X}}=(\widehat{t},\widehat{\boldsymbol{z}})=(\widehat{t},\widehat{\boldsymbol{x}},\widehat{\boldsymbol{p}}),\qquad\widehat{\boldsymbol{K}}=(-\widehat{\omega},\widehat{\boldsymbol{q}})=(-\widehat{\omega},\widehat{\boldsymbol{k}},\widehat{\boldsymbol{r}}). (64)

Assuming the convention that Latin indices from the beginning of the alphabet (a,b,c,)(a,b,c,\ldots) range from 0 to 2n2n, and a/Xa\partial_{a}\doteq\partial/\partial X^{a}, one can compactly express this as

X^a=Xa,K^a=ia.\widehat{X}^{a}=X^{a},\qquad\widehat{K}_{a}=-\mathrm{i}\partial_{a}. (65)

The eigenvectors of these operators will be denoted |𝑿\ket{{\boldsymbol{X}}} and |𝑲\ket{{\boldsymbol{K}}}:

𝑿^|𝑿=𝑿|𝑿,𝑲^|𝑲=𝑲|𝑲,\widehat{\boldsymbol{X}}\ket{{\boldsymbol{X}}}={\boldsymbol{X}}\ket{{\boldsymbol{X}}},\qquad\widehat{\boldsymbol{K}}\ket{{\boldsymbol{K}}}={\boldsymbol{K}}\ket{{\boldsymbol{K}}}, (66)

and we assume the usual normalization:

𝑿1|𝑿2=δ(𝑿1𝑿2),𝑲1|𝑲2=δ(𝑲1𝑲2).\braket{{\boldsymbol{X}}_{1}}{{\boldsymbol{X}}_{2}}=\delta({\boldsymbol{X}}_{1}-{\boldsymbol{X}}_{2}),\qquad\braket{{\boldsymbol{K}}_{1}}{{\boldsymbol{K}}_{2}}=\delta({\boldsymbol{K}}_{1}-{\boldsymbol{K}}_{2}). (67)

Both sets {|𝑿,𝑿N}\{\ket{{\boldsymbol{X}}},{\boldsymbol{X}}\in\mathbb{R}^{N}\} and {|𝑲,𝑲N}\{\ket{{\boldsymbol{K}}},{\boldsymbol{K}}\in\mathbb{R}^{N}\} form a complete basis on X\mathscr{H}_{X}, and the eigenvalues of these operators form a real extended phase space (𝑿,𝑲)({\boldsymbol{X}},{\boldsymbol{K}}), where

𝑿(t,𝒛),𝑲(ω,𝒒).{\boldsymbol{X}}\equiv(t,{\boldsymbol{z}}),\qquad{\boldsymbol{K}}\equiv(-\omega,{\boldsymbol{q}}). (68)

Particularly note the following formula, which will be used below:

Jαβq^αqβ=(𝒌^𝒓^)(𝟎n𝟏n𝟏n𝟎n)(𝒌𝒓)=𝒌^𝒓𝒓^𝒌.J^{\alpha\beta}\widehat{q}_{\alpha}q_{\beta}=\left(\begin{array}[]{cc}\widehat{\boldsymbol{k}}&\widehat{\boldsymbol{r}}\end{array}\right)\left(\begin{array}[]{cc}{\boldsymbol{0}}_{n}&{\boldsymbol{1}}_{n}\\ -{\boldsymbol{1}}_{n}&{\boldsymbol{0}}_{n}\\ \end{array}\right)\left(\begin{array}[]{c}{\boldsymbol{k}}\\ {\boldsymbol{r}}\\ \end{array}\right)=\widehat{\boldsymbol{k}}\cdot{\boldsymbol{r}}-\widehat{\boldsymbol{r}}\cdot{\boldsymbol{k}}. (69)

2.2.2 Wigner–Weyl transform

One can construct the Weyl symbol calculus on the extended phase space 𝑿{\boldsymbol{X}} just like it is done on spacetime 𝘅{\boldsymbol{\mathsf{x}}} in section 2.1, with an obvious modification of the notation. The Wigner–Weyl transform is defined as

A(𝑿,𝑲)=d𝑺𝑿+𝑺/2|A^|𝑿𝑺/2ei𝑲𝑺symbXA^,\displaystyle\displaystyle A({\boldsymbol{X}},{\boldsymbol{K}})=\int\mathrm{d}{\boldsymbol{S}}\braket{{\boldsymbol{X}}+{\boldsymbol{S}}/2}{\widehat{A}}{{\boldsymbol{X}}-{\boldsymbol{S}}/2}\mathrm{e}^{-\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}\equiv\text{symb}_{X}\widehat{A}, (70)
A^=1(2\upi)Nd𝑿d𝑲d𝑺|𝑿+𝑺/2A(𝑿,𝑲)𝑿𝑺/2|ei𝑲𝑺operXA.\displaystyle\displaystyle\widehat{A}=\frac{1}{(2\upi)^{N}}\int\mathrm{d}{\boldsymbol{X}}\,\mathrm{d}{\boldsymbol{K}}\,\mathrm{d}{\boldsymbol{S}}\ket{{\boldsymbol{X}}+{\boldsymbol{S}}/2}A({\boldsymbol{X}},{\boldsymbol{K}})\bra{{\boldsymbol{X}}-{\boldsymbol{S}}/2}\mathrm{e}^{\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}\equiv\text{oper}_{X}A. (71)

(Notice the change in the font and in the index compared to (26) and (25).) The corresponding Moyal product is denoted 🟊\bigstar (as opposed to \star introduced earlier):

A🟊B=A(𝑿,𝑲)ei^X/2B(𝑿,𝑲),A\bigstar B=A({\boldsymbol{X}},{\boldsymbol{K}})\,\mathrm{e}^{\mathrm{i}\widehat{\mathcal{L}}_{X}/2}B({\boldsymbol{X}},{\boldsymbol{K}}), (72)

where ^X𝑿𝑲𝑲𝑿\widehat{\mathcal{L}}_{X}\doteq\overset{{\scriptscriptstyle\leftarrow}}{\partial}_{\boldsymbol{X}}\cdot\overset{{\scriptscriptstyle\rightarrow}}{\partial}_{\boldsymbol{K}}-\overset{{\scriptscriptstyle\leftarrow}}{\partial}_{\boldsymbol{K}}\cdot\overset{{\scriptscriptstyle\rightarrow}}{\partial}_{\boldsymbol{X}} can be expressed as follows:

A^XBAtBω+AωBt+AxiBkiAkiBxi+ApiBriAriBpi.A\widehat{\mathcal{L}}_{X}B\doteq-\frac{\partial A}{\partial t}\frac{\partial B}{\partial\omega}+\frac{\partial A}{\partial\omega}\frac{\partial B}{\partial t}+\frac{\partial A}{\partial x^{i}}\frac{\partial B}{\partial k_{i}}-\frac{\partial A}{\partial k_{i}}\frac{\partial B}{\partial x^{i}}+\frac{\partial A}{\partial p^{i}}\frac{\partial B}{\partial r_{i}}-\frac{\partial A}{\partial r_{i}}\frac{\partial B}{\partial p^{i}}. (73)

If an operator A^\smash{\widehat{A}} is local in 𝒑\smash{{\boldsymbol{p}}}, its 𝑿\smash{{\boldsymbol{X}}} representation and 𝘅\smash{{\boldsymbol{\mathsf{x}}}} representation satisfy

t,𝒙,𝒑|𝑨^|t,𝒙,𝒑=t,𝒙|𝑨^|t,𝒙δ(𝒑𝒑),\displaystyle\braket{t,{\boldsymbol{x}},{\boldsymbol{p}}}{\widehat{\boldsymbol{A}}}{t^{\prime},{\boldsymbol{x}}^{\prime},{\boldsymbol{p}}^{\prime}}=\braket{t,{\boldsymbol{x}}}{\widehat{\boldsymbol{A}}}{t^{\prime},{\boldsymbol{x}}^{\prime}}\delta({\boldsymbol{p}}-{\boldsymbol{p}}^{\prime}), (74)

and therefore the Weyl symbol of A^\smash{\widehat{A}} is the same irrespective of whether the operator is considered on X\smash{\mathscr{H}_{X}} or on 𝗑\smash{\mathscr{H}_{\mathsf{x}}}. In this case, we will use a unifying notation symbA^\smash{\text{symb}\,\widehat{A}} instead of symbXA^\smash{\text{symb}_{X}\widehat{A}} and symb𝗑A^\smash{\text{symb}_{\mathsf{x}}\widehat{A}}.

2.2.3 Wigner functions and Wigner matrices

The density operator of a given scalar field ψ\psi is given by

W^ψ(2\upi)N|ψψ|.\widehat{W}_{\psi}\doteq(2\upi)^{-N}\ket{\psi}\bra{\psi}. (75)

The symbol of this operator, Wψ=symbXW^ψ\smash{W_{\psi}=\text{symb}_{X}\widehat{W}_{\psi}}, is a real function called the Wigner function. It is given by

Wψ(𝑿,𝑲)=1(2\upi)Nd𝑺ψ(𝑿+𝑺/2)ψ(𝑿𝑺/2)ei𝑲𝑺,W_{\psi}({\boldsymbol{X}},{\boldsymbol{K}})=\frac{1}{(2\upi)^{N}}\int\mathrm{d}{\boldsymbol{S}}\,\psi({\boldsymbol{X}}+{\boldsymbol{S}}/2)\psi^{*}({\boldsymbol{X}}-{\boldsymbol{S}}/2)\,\mathrm{e}^{-\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}, (76)

which can be understood as the (inverse) Fourier image of

Cψ(𝑿,𝑺)ψ(𝑿+𝑺/2)ψ(𝑿𝑺/2)=d𝑲Wψ(𝑿,𝑲)ei𝑲𝑺.\displaystyle C_{\psi}({\boldsymbol{X}},{\boldsymbol{S}})\doteq\psi({\boldsymbol{X}}+{\boldsymbol{S}}/2)\psi^{*}({\boldsymbol{X}}-{\boldsymbol{S}}/2)=\int\mathrm{d}{\boldsymbol{K}}\,W_{\psi}({\boldsymbol{X}},{\boldsymbol{K}})\,\mathrm{e}^{\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}. (77)

In particular, one has

d𝒓symbXW^ψ=symb𝗑𝖶^ψ(𝒑),\displaystyle\int\mathrm{d}{\boldsymbol{r}}\,\text{symb}_{X}\widehat{W}_{\psi}=\text{symb}_{\mathsf{x}}\widehat{\mathsf{W}}_{\psi}({\boldsymbol{p}}), (78)

where the right-hand side is 𝖶ψ\smash{\mathsf{W}_{\psi}} given by (43), with 𝒑{\boldsymbol{p}} treated as a parameter. Also, for real fields,

Wψ(𝑿,𝑲)=Wψ(𝑿,𝑲).W_{\psi}({\boldsymbol{X}},{\boldsymbol{K}})=W_{\psi}({\boldsymbol{X}},-{\boldsymbol{K}}). (79)

The density operator of a given vector field 𝝍=(ψ1,ψ2,,ψM){\boldsymbol{\psi}}=(\psi^{1},\psi^{2},\ldots,\psi^{M}) is a matrix operator

𝑾^𝝍(2\upi)N|𝝍𝝍|.\widehat{\boldsymbol{W}}_{{\boldsymbol{\psi}}}\doteq(2\upi)^{-N}\ket{{\boldsymbol{\psi}}}\bra{{\boldsymbol{\psi}}}. (80)

The symbol of this operator, or the Wigner matrix, is a Hermitian matrix function

𝑾𝝍(𝑿,𝑲)=1(2\upi)Nd𝑺𝝍(𝑿+𝑺/2)𝝍(𝑿𝑺/2)ei𝑲𝑺,{\boldsymbol{W}}_{{\boldsymbol{\psi}}}({\boldsymbol{X}},{\boldsymbol{K}})=\frac{1}{(2\upi)^{N}}\int\mathrm{d}{\boldsymbol{S}}\,{\boldsymbol{\psi}}({\boldsymbol{X}}+{\boldsymbol{S}}/2){\boldsymbol{\psi}}^{\dagger}({\boldsymbol{X}}-{\boldsymbol{S}}/2)\,\mathrm{e}^{-\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}, (81)

which can be understood as the (inverse) Fourier image of

𝑪𝝍(𝑿,𝑺)𝝍(𝑿+𝑺/2)𝝍(𝑿𝑺/2)=d𝑲𝑾𝝍(𝑿,𝑲)ei𝑲𝑺.\displaystyle{\boldsymbol{C}}_{{\boldsymbol{\psi}}}({\boldsymbol{X}},{\boldsymbol{S}})\doteq{\boldsymbol{\psi}}({\boldsymbol{X}}+{\boldsymbol{S}}/2){\boldsymbol{\psi}}^{\dagger}({\boldsymbol{X}}-{\boldsymbol{S}}/2)=\int\mathrm{d}{\boldsymbol{K}}\,{\boldsymbol{W}}_{{\boldsymbol{\psi}}}({\boldsymbol{X}},{\boldsymbol{K}})\,\mathrm{e}^{\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}. (82)

In particular, one has

d𝒓symbX𝑾^𝝍=symb𝗑𝗪^𝝍(𝒑),\displaystyle\int\mathrm{d}{\boldsymbol{r}}\,\text{symb}_{X}\widehat{\boldsymbol{W}}_{{\boldsymbol{\psi}}}=\text{symb}_{\mathsf{x}}\widehat{\boldsymbol{{\boldsymbol{\mathsf{W}}}}}_{{\boldsymbol{\psi}}}({\boldsymbol{p}}), (83)

where the right-hand side is 𝗪𝝍\smash{{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}} given by (51), with 𝒑{\boldsymbol{p}} treated as a parameter. Also, for real fields,

𝑾𝝍(𝑿,𝑲)=𝑾𝝍(𝑿,𝑲)=𝑾𝝍(𝑿,𝑲).{\boldsymbol{W}}_{{\boldsymbol{\psi}}}({\boldsymbol{X}},{\boldsymbol{K}})={\boldsymbol{W}}_{{\boldsymbol{\psi}}}^{\intercal}({\boldsymbol{X}},-{\boldsymbol{K}})={\boldsymbol{W}}_{{\boldsymbol{\psi}}}^{*}({\boldsymbol{X}},-{\boldsymbol{K}}). (84)

Like those on (𝘅,𝗸)\smash{({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})}, the Wigner matrices (Wigner functions) on (𝑿,𝑲)\smash{({\boldsymbol{X}},{\boldsymbol{K}})} become positive-semidefinite (non-negative), and characterize local properties of the corresponding fields, when averaged over a sufficiently large phase-space volume Δ𝑿Δ𝑲1\Delta{\boldsymbol{X}}\,\Delta{\boldsymbol{K}}\gtrsim 1.

2.3 Summary of section 2

In summary, we have introduced a generic n\smash{n}-dimensional physical space 𝒙\smash{{\boldsymbol{x}}}, the dual n\smash{n}-dimensional wavevector space 𝒌\smash{{\boldsymbol{k}}}, the corresponding 𝗇\smash{{\mathsf{n}}}-dimensional (𝗇=n+1\smash{{\mathsf{n}}=n+1}) spacetime 𝘅(t,𝒙)\smash{{\boldsymbol{\mathsf{x}}}\equiv(t,{\boldsymbol{x}})}, and the dual 𝗇\smash{{\mathsf{n}}}-dimensional wavevector space 𝗸(ω,𝒌)\smash{{\boldsymbol{\mathsf{k}}}\equiv(-\omega,{\boldsymbol{k}})}. We have also introduced an n\smash{n}-dimensional momentum space 𝒑\smash{{\boldsymbol{p}}}, the corresponding 2n\smash{2n}-dimensional phase space 𝒛(𝒙,𝒑)\smash{{\boldsymbol{z}}\equiv({\boldsymbol{x}},{\boldsymbol{p}})}, the N\smash{N}-dimensional (N=2n+1\smash{N=2n+1}) extended space 𝑿(t,𝒛)(t,𝒙,𝒑)\smash{{\boldsymbol{X}}\equiv(t,{\boldsymbol{z}})\equiv(t,{\boldsymbol{x}},{\boldsymbol{p}})}, and the dual N\smash{N}-dimensional wavevector space 𝑲(ω,𝒒)(ω,𝒌,𝒓)\smash{{\boldsymbol{K}}\equiv(-\omega,{\boldsymbol{q}})\equiv(-\omega,{\boldsymbol{k}},{\boldsymbol{r}})}, where 𝒓\smash{{\boldsymbol{r}}} is the n\smash{n}-dimensional wavevector space dual to 𝒑\smash{{\boldsymbol{p}}}. We have also introduced the 2N\smash{2N}-dimensional phase space (𝑿,𝑲)\smash{({\boldsymbol{X}},{\boldsymbol{K}})}. Each of the said variables has a corresponding operator associated with it, which is denoted with a caret. For example, 𝒙^\smash{\widehat{\boldsymbol{x}}} is the operator of position in the 𝒙\smash{{\boldsymbol{x}}} space, and 𝒌^=i𝒙\smash{\widehat{\boldsymbol{k}}=-\mathrm{i}\partial_{{\boldsymbol{x}}}} is the corresponding wavevector operator.

Functions on 𝘅\smash{{\boldsymbol{\mathsf{x}}}} form a Hilbert space 𝗑\smash{\mathscr{H}_{\mathsf{x}}}, and the corresponding bra-ket notation is introduced as usual. Any operator 𝖠^\smash{\widehat{\mathsf{A}}} on 𝗑\smash{\mathscr{H}_{\mathsf{x}}} can be represented by its Weyl symbol 𝖠(𝘅,𝗸)\smash{\mathsf{A}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})}. The correspondence between operators and their symbols, 𝖠^𝖠\smash{\widehat{\mathsf{A}}\leftrightarrow\mathsf{A}}, is determined by the Wigner–Weyl transform and is natural in the sense that (29) is satisfied. In particular, 𝖠^𝖡^𝖠𝖡\smash{\widehat{\mathsf{A}}\widehat{\mathsf{B}}\leftrightarrow\mathsf{A}\star\mathsf{B}}, where \smash{\star} is the Moyal product on (𝘅,𝗸)\smash{({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})}. When the geometrical-optics parameter is negligible (ϵ0\smash{\epsilon\to 0}), one has 𝖠^=𝖠(𝘅^,𝗸^)\smash{\widehat{\mathsf{A}}=\mathsf{A}(\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}},\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}})} and the Moyal product becomes the usual product. Similarly, functions on 𝑿\smash{{\boldsymbol{X}}} form a Hilbert space X\smash{\mathscr{H}_{X}}, the corresponding bra-ket notation is also introduced as usual, any operator A^\smash{\widehat{A}} on X\smash{\mathscr{H}_{X}} can be represented by its Weyl symbol A(𝑿,𝑲)\smash{A({\boldsymbol{X}},{\boldsymbol{K}})}, and A^B^A🟊B\smash{\widehat{A}\widehat{B}\leftrightarrow A\bigstar B}. An operator that is local in 𝒑\smash{{\boldsymbol{p}}} has the same symbol irrespective of whether it is considered on 𝗑\smash{\mathscr{H}_{\mathsf{x}}} or on X\smash{\mathscr{H}_{X}}.

Any given field ψ\smash{\psi} generates the corresponding density operator (2\upi)𝗇|ψψ|\smash{(2\upi)^{-{\mathsf{n}}}\ket{\psi}\bra{\psi}} and its symbol called the Wigner function (Wigner matrix, if the field is a vector). If the density operator is considered on 𝗑\smash{\mathscr{H}_{\mathsf{x}}}, it is denoted 𝖶^ψ\smash{\widehat{\mathsf{W}}_{\psi}} and the corresponding Wigner function is denoted 𝖶ψ(𝘅,𝗸)\smash{\mathsf{W}_{\psi}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})}. If the density operator is considered on X\smash{\mathscr{H}_{X}}, it is denoted W^ψ\smash{\widehat{W}_{\psi}} and the corresponding Wigner function is denoted Wψ(𝑿,𝑲)\smash{W_{\psi}({\boldsymbol{X}},{\boldsymbol{K}})}. The two Wigner functions are related via d𝒓Wψ(t,𝒙,𝒑,ω,𝒌,𝒓)=𝖶ψ(t,𝒙,ω,𝒌;𝒑)\smash{\int\mathrm{d}{\boldsymbol{r}}\,W_{\psi}(t,{\boldsymbol{x}},{\boldsymbol{p}},\omega,{\boldsymbol{k}},{\boldsymbol{r}})=\mathsf{W}_{\psi}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}})}, where 𝒑\smash{{\boldsymbol{p}}} enters 𝖶ψ\smash{\mathsf{W}_{\psi}} as a parameter, if at all. If averaged over a sufficiently large phase-space volume, the Wigner functions (matrices) are non-negative (positive-semidefinite) and characterize local properties of the corresponding fields.

3 Model

Here, we introduce the general assumptions and the key ingredients of our theory.

3.1 Basic assumptions

3.1.1 Ordering

Let us consider particles governed by a Hamiltonian H=H¯+H~H=\overline{H}+\widetilde{H} such that

H~=𝒪(ε)H¯=𝒪(1).\displaystyle\widetilde{H}=\mathcal{O}(\varepsilon)\ll\overline{H}=\mathcal{O}(1). (85)

In other words, H~\widetilde{H} serves as a small perturbation to the leading-order Hamiltonian H¯\overline{H}. The system will be described in canonical variables 𝒛(𝒙,𝒑)2n\smash{{\boldsymbol{z}}\equiv({\boldsymbol{x}},{\boldsymbol{p}})\in\mathbb{R}^{2n}}. Let us also assume that the system is close to being homogeneous in 𝒙{\boldsymbol{x}}. This includes two conditions. First, we require that the external fields are weak (yet see section 3.1.2), meaning

𝒙H¯κxH¯=𝒪(ϵ),𝒑H¯κpH¯=𝒪(1),\displaystyle\partial_{{\boldsymbol{x}}}\overline{H}\sim\kappa_{x}\overline{H}=\mathcal{O}(\epsilon),\qquad\partial_{{\boldsymbol{p}}}\overline{H}\sim\kappa_{p}\overline{H}=\mathcal{O}(1), (86)

where ϵ1\epsilon\ll 1 is a small parameter, κx\kappa_{x} and κp\kappa_{p} are the characteristic inverse scales in the 𝒙{\boldsymbol{x}} and 𝒑{\boldsymbol{p}} spaces, respectively, and the bar denotes local averaging.111111An exception will be made for eikonal waves, specifically, for quantities evaluated on the local wavevector 𝗸¯(ω¯,𝒌¯)\overline{{\boldsymbol{\mathsf{k}}}}\equiv(-\overline{\omega},\overline{{\boldsymbol{k}}}). Hence, the particle momenta 𝒑{\boldsymbol{p}} are close to being local invariants. Second, the statistical properties of H~\smash{\widetilde{H}} are also assumed to vary in 𝒙{\boldsymbol{x}} slowly. These properties can be characterized using the density operator of the perturbation Hamiltonian,

W^(2\upi)N|H~H~|,\displaystyle\widehat{W}\doteq(2\upi)^{-N}\ket{\widetilde{H}}\bra{\widetilde{H}}, (87)

and its symbol, the (real) Wigner function, as in (43):

W(𝑿,𝑲)=1(2\upi)Nd𝑺H~(𝑿+𝑺/2)H~(𝑿𝑺/2)ei𝑲𝑺.\displaystyle W({\boldsymbol{X}},{\boldsymbol{K}})=\frac{1}{(2\upi)^{N}}\int\mathrm{d}{\boldsymbol{S}}\,\widetilde{H}({\boldsymbol{X}}+{\boldsymbol{S}}/2)\widetilde{H}({\boldsymbol{X}}-{\boldsymbol{S}}/2)\,\mathrm{e}^{-\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}. (88)

Specifically, we will use the average Wigner function, W¯\overline{W}, which represents the Fourier spectrum of the symmetrized autocorrelation function of H~\widetilde{H}:

C¯(𝑿,𝑺)H~(𝑿+𝑺/2)H~(𝑿𝑺/2)¯=d𝑲W¯(𝑿,𝑲)ei𝑲𝑺.\displaystyle\overline{C}({\boldsymbol{X}},{\boldsymbol{S}})\doteq\overline{\widetilde{H}({\boldsymbol{X}}+{\boldsymbol{S}}/2)\widetilde{H}({\boldsymbol{X}}-{\boldsymbol{S}}/2)}=\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}({\boldsymbol{X}},{\boldsymbol{K}})\,\mathrm{e}^{\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}. (89)

The averaging is performed over sufficiently large volume of 𝒙{\boldsymbol{x}} to eliminate rapid oscillations and also over phase-space volumes Δ𝑿Δ𝑲1\Delta{\boldsymbol{X}}\,\Delta{\boldsymbol{K}}\gtrsim 1, which guarantees W¯\overline{W} to be non-negative and local (section 2.2.3). The function W¯\overline{W} can be understood as a measure of the phase-space density of wave quanta when the latter is well defined (section 7).

We will assume121212As a reminder, the notation A=𝒪(ϵ)\smash{A=\mathcal{O}(\epsilon)} does not rule out the possibility that A/ϵ\smash{A/\epsilon} is small. Also note that the terms ‘\sim’ and ‘of order’ in this paper mean the same as ‘𝒪\mathcal{O}’.

tW¯=𝒪(ϵ),𝒙W¯=𝒪(ϵ),𝒑W¯=𝒪(1).\displaystyle\partial_{t}\overline{W}=\mathcal{O}(\epsilon),\qquad\partial_{\boldsymbol{x}}\overline{W}=\mathcal{O}(\epsilon),\qquad\partial_{\boldsymbol{p}}\overline{W}=\mathcal{O}(1). (90)

That said, we will also allow (albeit not require) for oscillations to be constrained by a dispersion relation. In this case, W¯δ(ωω¯(t,𝒙))\overline{W}\,\propto\,\delta(\omega-\overline{\omega}(t,{\boldsymbol{x}})), so (90) per se is not satisfied; then we assume a similar ordering for dωW¯\int\mathrm{d}\omega\,\overline{W} instead. Also note that in application to the standard QLT of homogeneous turbulence (Stix, 1992, chapter 16), ϵ\epsilon is understood as the geometrical-optics parameter characterizing the smallness of the linear-instability growth rates. (We discuss the ordering further in the end of section 3.3.)

3.1.2 Quasilinear approximation

The particle-motion equations can be written as

z˙α={zα,H¯+H~}=vα+uα,\displaystyle\dot{z}^{\alpha}=\{z^{\alpha},\overline{H}+\widetilde{H}\}=v^{\alpha}+u^{\alpha}, (91)

where vαv^{\alpha} and uαu^{\alpha} are understood as the unperturbed phase-space velocity and the perturbation to the phase-space velocity, respectively:

vαJαββH¯,uαJαββH~.\displaystyle v^{\alpha}\doteq J^{\alpha\beta}\partial_{\beta}\overline{H},\qquad u^{\alpha}\doteq J^{\alpha\beta}\partial_{\beta}\widetilde{H}. (92)

The notation viv^{i} (with i=1,2,ni=1,2,\ldots n) will also be used for the spatial part of the phase-space velocity vαv^{\alpha}, i.e. for the true velocity per se. Likewise, 𝒗{\boldsymbol{v}} will be used to denote either the phase-space velocity vector or the spatial velocity vector depending on the context. Also note that a slightly different definition of 𝒗{\boldsymbol{v}} will be used starting from section 5.6.

The corresponding Klimontovich equation for the particle distribution f(t,𝒛)f(t,{\boldsymbol{z}}) is

tf={H¯+H~,f}.\displaystyle\partial_{t}f=\{\overline{H}+\widetilde{H},f\}. (93)

(If collisions are not of interest, (93) can as well be understood as the Vlasov equation. Also, a small collision term can be included ad hoc; see the comment in the end of section 3.3.) Let us search for ff in the form

f=f¯+f~,f~¯=0.\displaystyle f=\overline{f}+\widetilde{f},\qquad\overline{\widetilde{f}}=0. (94)

The equations for f¯\smash{\overline{f}} and f~\smash{\widetilde{f}} are obtained as the average and oscillating parts of (93), and we neglect the nonlinearity in the equation for f~\smash{\widetilde{f}}, following the standard QL approximation (Stix, 1992, chapter 16). Then, one obtains

tf¯={H¯,f¯}+{H~,f~}¯,\displaystyle\partial_{t}\overline{f}=\{\overline{H},\overline{f}\}+\overline{\{\widetilde{H},\widetilde{f}\}}, (95)
tf~={H¯,f~}+{H~,f¯}.\displaystyle\partial_{t}\widetilde{f}=\{\overline{H},\widetilde{f}\}+\{\widetilde{H},\overline{f}\}. (96)

A comment is due here regarding plasmas in strong fields and magnetized plasmas in particular. Our formulation can be applied to such plasmas in canonical angle–action variables (ϕ,𝗝)\smash{({\boldsymbol{\phi}},{\boldsymbol{\mathsf{J}}})}. For fast angle variables, the ordering (86) is not satisfied and the Weyl symbol calculus is inapplicable as is (see the footnote on p. 5). Such systems can be accommodated by representing the distribution function as a Fourier series in ϕ\smash{{\boldsymbol{\phi}}} and treating the individual-harmonic amplitudes separately as slow functions of the remaining coordinates. Then, our averaging procedure subsumes averaging over ϕ\smash{{\boldsymbol{\phi}}}, so the averaged quantities are ϕ\smash{{\boldsymbol{\phi}}}-independent and (86) is reinstated. In particular, magnetized plasmas can be described using guiding-center variables. Although not canonical by default (Littlejohn, 1983), they can always be cast in a canonical form, at least in principle (Littlejohn, 1979). Examples of canonical guiding-center variables are reviewed in (Cary & Brizard, 2009). To make the connection with the homogeneous-plasma theory, one can also order the canonical pairs of guiding-center variables such that they would describe the gyromotion, the parallel motion, and the drifts separately (Wong, 2000). This readily leads to results similar to those in (Catto et al., 2017). Further discussions on this topic are left to future papers.

3.2 Equation for f~{\widetilde{f}}

Let us consider solutions of (96) as a subclass of solutions of the more general equation

τf~=L^f~+,(𝑿){H~,f¯}.\displaystyle\partial_{\tau}\widetilde{f}=\widehat{L}\widetilde{f}+\mathscr{F},\qquad\mathscr{F}({\boldsymbol{X}})\doteq\{\widetilde{H},\overline{f}\}. (97)

Here, we have introduced an auxiliary second ‘time’ τ\smash{\tau}, the operator

L^t+{H¯,}=t+Jαβ(αH¯)β=tvλλ=Vaa\displaystyle\widehat{L}\doteq-\partial_{t}+\{\overline{H},\sqbullet\}=-\partial_{t}+J^{\alpha\beta}(\partial_{\alpha}\overline{H})\partial_{\beta}=-\partial_{t}-v^{\lambda}\partial_{\lambda}=-V^{a}\partial_{a} (98)

(here and further, \sqbullet denotes a placeholder), and 𝑽(𝑿)(1,𝒗(t,𝒛)){\boldsymbol{V}}({\boldsymbol{X}})\equiv(1,{\boldsymbol{v}}(t,{\boldsymbol{z}})) is the unperturbed velocity in the 𝑿{\boldsymbol{X}} space. Note that

aVa=λvλ=0\displaystyle\partial_{a}V^{a}=\partial_{\lambda}v^{\lambda}=0 (99)

due to the incompressibility of the phase flow. Hence, [a,Va]=0\smash{[\partial_{a},V^{a}]=0}, so L^\smash{\widehat{L}} is anti-Hermitian.

Let us search for a solution of (97) in the form131313Using the auxiliary variable τ\smash{\tau} allows us to express the propagator as a regular exponential, rather than ordered exponential, even for tt-dependent H¯\smash{\overline{H}}, because L^\smash{\widehat{L}} is independent of τ\smash{\tau}.

f~(τ,𝑿)=eL^τξ(τ,𝑿).\displaystyle\widetilde{f}(\tau,{\boldsymbol{X}})=\mathrm{e}^{\widehat{L}\tau}\xi(\tau,{\boldsymbol{X}}). (100)

Then, τf~=L^f~+eL^ττξ\partial_{\tau}\widetilde{f}=\widehat{L}\widetilde{f}+\mathrm{e}^{\widehat{L}\tau}\partial_{\tau}\xi, so τξ=eL^τ(𝑿)\partial_{\tau}\xi=\mathrm{e}^{-\widehat{L}\tau}\mathscr{F}({\boldsymbol{X}}) and therefore

ξ(τ,𝑿)=eL^τ0ξ0(𝑿)+τ0τdτeL^τ(𝑿),\displaystyle\xi(\tau,{\boldsymbol{X}})=\mathrm{e}^{-\widehat{L}\tau_{0}}\xi_{0}({\boldsymbol{X}})+\int_{\tau_{0}}^{\tau}\mathrm{d}\tau^{\prime}\,\mathrm{e}^{-\widehat{L}\tau^{\prime}}\mathscr{F}({\boldsymbol{X}}), (101)

where ξ0(𝑿)f~(τ0,𝑿)\smash{\xi_{0}({\boldsymbol{X}})\doteq\widetilde{f}(\tau_{0},{\boldsymbol{X}})}. Hence, one obtains

f~(τ,𝑿)=eL^(ττ0)ξ0(𝑿)+τ0τdτeL^(ττ)(𝑿),\displaystyle\widetilde{f}(\tau,{\boldsymbol{X}})=\mathrm{e}^{\widehat{L}(\tau-\tau_{0})}\xi_{0}({\boldsymbol{X}})+\int_{\tau_{0}}^{\tau}\mathrm{d}\tau^{\prime}\,\mathrm{e}^{-\widehat{L}(\tau^{\prime}-\tau)}\mathscr{F}({\boldsymbol{X}}), (102)

or equivalently, using τ′′ττ\tau^{\prime\prime}\doteq\tau-\tau^{\prime},

f~(τ,𝑿)=g0(τ,𝑿)+0ττ0dτ′′T^τ′′(𝑿).\displaystyle\widetilde{f}(\tau,{\boldsymbol{X}})=g_{0}(\tau,{\boldsymbol{X}})+\int_{0}^{\tau-\tau_{0}}\mathrm{d}\tau^{\prime\prime}\,\widehat{T}_{\tau^{\prime\prime}}\mathscr{F}({\boldsymbol{X}}). (103)

Here, g0\smash{g_{0}} is a solution of τg0=L^g0\partial_{\tau}g_{0}=\widehat{L}g_{0}, specifically,

g0(τ,𝑿)T^ττ0ξ0(𝑿),g0(τ0,𝑿)=f~(τ0,𝑿),\displaystyle g_{0}(\tau,{\boldsymbol{X}})\doteq\widehat{T}_{\tau-\tau_{0}}\xi_{0}({\boldsymbol{X}}),\qquad g_{0}(\tau_{0},{\boldsymbol{X}})=\widetilde{f}(\tau_{0},{\boldsymbol{X}}), (104)

and we have also introduced

T^τeL^τ=eτVaa.\displaystyle\widehat{T}_{\tau}\doteq\mathrm{e}^{\widehat{L}\tau}=\mathrm{e}^{-\tau V^{a}\partial_{a}}. (105)

Because L^\smash{\widehat{L}} is anti-Hermitian, the operator T^τ\smash{\widehat{T}_{\tau}} is unitary, and comparison with (19) shows that it can be recognized as a shift operator. For further details, see section 4.1.

Using T^τ\smash{\widehat{T}_{\tau}}, one can express (103) as

f~=g0+𝒢^,𝒢^0ττ0dτT^τ,\displaystyle\widetilde{f}=g_{0}+\widehat{\mathscr{G}}\mathscr{F},\qquad\textstyle\widehat{\mathscr{G}}\doteq\int_{0}^{\tau-\tau_{0}}\mathrm{d}\tau^{\prime}\,\widehat{T}_{\tau^{\prime}}, (106)

where 𝒢^\widehat{\mathscr{G}} is the Green’s operator understood as the right inverse of the operator τL^\smash{\partial_{\tau}-\widehat{L}}, or on the space of τ\smash{\tau}-independent functions, t{H¯,}\smash{\partial_{t}-\{\overline{H},\sqbullet\}}. Let us rewrite this operator as 𝒢^=𝒢^<+𝒢^>\smash{\widehat{\mathscr{G}}=\widehat{\mathscr{G}}_{<}+\widehat{\mathscr{G}}_{>}}, where

𝒢^<=0ττ0dτeντT^τ,𝒢^>=0ττ0dτ(1eντ)T^τ,\displaystyle\widehat{\mathscr{G}}_{<}=\int_{0}^{\tau-\tau_{0}}\mathrm{d}\tau^{\prime}\,\mathrm{e}^{-\nu\tau^{\prime}}\widehat{T}_{\tau^{\prime}},\qquad\widehat{\mathscr{G}}_{>}=\int_{0}^{\tau-\tau_{0}}\mathrm{d}\tau^{\prime}\,(1-\mathrm{e}^{-\nu\tau^{\prime}})\widehat{T}_{\tau^{\prime}}, (107)

and ν\nu is a positive constant. Note that 𝒢^<\widehat{\mathscr{G}}_{<} is well defined at τ0\tau_{0}\to-\infty, meaning that 𝒢^<\widehat{\mathscr{G}}_{<}\mathscr{F} is well defined for any physical (bounded) field \mathscr{F}.141414Unlike classic plasma-wave theory, this approach does not involve spectral decomposition, so there is no need to consider fields that are exponential in time on the whole interval (,)\smash{(-\infty,\infty)}. Thus, so is g0+𝒢^>\smash{g_{0}+\widehat{\mathscr{G}}_{>}\mathscr{F}}. Let us take τ0\tau_{0}\to-\infty and then take ν0+\nu\to 0+. (Here, 0+0+ denotes that ν\nu must remain positive, i.e. the upper limit is taken.) Then, (106) can be expressed as

f~=g+G^,glimν0+limτ0(g0+𝒢^>).\displaystyle\widetilde{f}=g+\widehat{G}\mathscr{F},\qquad g\doteq\lim_{\nu\to 0+}\lim_{\tau_{0}\to-\infty}(g_{0}+\widehat{\mathscr{G}}_{>}\mathscr{F}). (108)

Here, we introduced an ‘effective’ Green’s operator G^limν0+limτ0𝒢^<\smash{\widehat{G}\doteq\lim_{\nu\to 0+}\lim_{\tau_{0}\to-\infty}\widehat{\mathscr{G}}_{<}}, i.e.

G^limν0+0dτeντT^τ.\displaystyle\widehat{G}\doteq\lim_{\nu\to 0+}\int_{0}^{\infty}\mathrm{d}\tau\,\mathrm{e}^{-\nu\tau}\widehat{T}_{\tau}. (109)

This operator will be discussed in section 4.2, and gg will be discussed in section 4.3. Meanwhile, note that because τ\smash{\tau} is just an auxiliary variable, we will be interested in solutions independent of τ\tau. In particular, this means that f~(τ0,𝑿)=f~(𝑿)\smash{\widetilde{f}(\tau_{0},{\boldsymbol{X}})=\widetilde{f}({\boldsymbol{X}})}, so ξ0(𝑿)=f~(𝑿)\smash{\xi_{0}({\boldsymbol{X}})=\widetilde{f}({\boldsymbol{X}})}, so (104) leads to

g0(τ,𝑿)=T^ττ0f~(𝑿).\displaystyle g_{0}(\tau,{\boldsymbol{X}})=\widehat{T}_{\tau-\tau_{0}}\widetilde{f}({\boldsymbol{X}}). (110)

3.3 Equation for f¯{\overline{f}}

Using (106), one can rewrite (96) for f¯\overline{f} as follows:

tf¯={H¯,f¯}+{H~,g}¯+{H~,G^{H~,f¯}}¯.\displaystyle\partial_{t}\overline{f}=\{\overline{H},\overline{f}\}+\overline{\{\widetilde{H},g\}}+\overline{\{\widetilde{H},\widehat{G}\{\widetilde{H},\overline{f}\}\}}. (111)

Notice that

{H~,g}={g,H~}=α(JαβgβH~)=α(uαg)\{\widetilde{H},g\}=-\{g,\widetilde{H}\}=-\partial_{\alpha}(J^{\alpha\beta}\,g\partial_{\beta}\widetilde{H})=-\partial_{\alpha}(u^{\alpha}g) (112)

and also

{H~,G^{H~,f¯}}\displaystyle\{\widetilde{H},\widehat{G}\{\widetilde{H},\overline{f}\}\} =β(Jαβ(αH~)G^{H~,f¯})\displaystyle=\partial_{\beta}(J^{\alpha\beta}(\partial_{\alpha}\widetilde{H})\widehat{G}\{\widetilde{H},\overline{f}\})
=β(Jαβ(αH~)G^(Jμν(μH~)(νf¯)))\displaystyle=\partial_{\beta}(J^{\alpha\beta}(\partial_{\alpha}\widetilde{H})\widehat{G}(J^{\mu\nu}(\partial_{\mu}\widetilde{H})(\partial_{\nu}\overline{f})))
=β(uβG^(uννf¯)).\displaystyle=\partial_{\beta}(u^{\beta}\widehat{G}(u^{\nu}\partial_{\nu}\overline{f})). (113)

The field uαu^{\alpha} enters here as a multiplication factor and can be considered as an operator:

u^αψ(𝑿)uα(𝑿)ψ(𝑿).\displaystyle\widehat{u}^{\alpha}\psi({\boldsymbol{X}})\doteq u^{\alpha}({\boldsymbol{X}})\psi({\boldsymbol{X}}). (114)

Then, (113) can be compactly represented as

{H~,G^{H~,f¯}}=α(u^αG^u^ββf¯).\displaystyle\{\widetilde{H},\widehat{G}\{\widetilde{H},\overline{f}\}\}=\partial_{\alpha}(\widehat{u}^{\alpha}\widehat{G}\widehat{u}^{\beta}\partial_{\beta}\overline{f}). (115)

We will also use the notation

dtt+vγγ=t{H¯,}.\displaystyle\mathrm{d}_{t}\doteq\partial_{t}+v^{\gamma}\partial_{\gamma}=\partial_{t}-\{\overline{H},\sqbullet\}. (116)

This leads to the following equation for f¯\overline{f}:

dtf¯=α(D^αββf¯)+Γ,\displaystyle\mathrm{d}_{t}\overline{f}=\partial_{\alpha}(\widehat{D}^{\alpha\beta}\partial_{\beta}\overline{f})+\Gamma, (117)

where we introduced the following average quantities:

D^αβu^αG^u^β¯,Γα(uαg¯),\displaystyle\widehat{D}^{\alpha\beta}\doteq\overline{\widehat{u}^{\alpha}\widehat{G}\widehat{u}^{\beta}},\qquad\Gamma\doteq-\partial_{\alpha}(\overline{u^{\alpha}g}), (118)

Our goal is to derive explicit approximate expressions for the quantities (118) and to rewrite (117) in a more tractable form using the assumptions introduced in section 3.1. We will use151515Starting with section 5.6, we will assume dtf¯ϵε2f¯\smash{\mathrm{d}_{t}\overline{f}\sim\epsilon\varepsilon^{2}\overline{f}} instead.

tf¯{H¯,f¯}=𝒪(ϵ),dtf¯=𝒪(ε2),\displaystyle\partial_{t}\overline{f}\sim\{\overline{H},\overline{f}\}=\mathcal{O}(\epsilon),\qquad\mathrm{d}_{t}\overline{f}=\mathcal{O}(\varepsilon^{2}), (119)

and we will keep terms of order ϵ\epsilon, ε2\varepsilon^{2}, and ϵε2\smash{\epsilon\varepsilon^{2}} in the equation for f¯\smash{\overline{f}}, while terms of order ε4\smash{\varepsilon^{4}}, ϵ2ε2\smash{\epsilon^{2}\varepsilon^{2}}, and higher will be neglected. This implies the ordering

ε2ϵε1.\displaystyle\varepsilon^{2}\ll\epsilon\ll\varepsilon\ll 1. (120)

As a reminder, ε\smash{\varepsilon} is a linear measure of the characteristic amplitude of oscillations, and ϵ\smash{\epsilon} is the geometrical-optics parameter, which is proportional to the inverse scale of the plasma inhomogeneity in spacetime. As usual then, linear dissipation is assumed to be of order ϵ\smash{\epsilon}. This model implies the assumption that collisionless dissipation is much stronger than collisional dissipation, which is to emerge as an effect quadratic in f~\smash{\widetilde{f}} (section 6). Furthermore, the inverse plasma parameter161616By the plasma parameter we mean the number of particles within the Debye sphere. will be assumed to be of order ϵ\smash{\epsilon}, so the collision operator for f¯\smash{\overline{f}} (section 6.8) will be of order ϵε2\smash{\epsilon\varepsilon^{2}}. Within the assumed accuracy, this operator must be retained, while the dynamics of f~\smash{\widetilde{f}} is considered linear and therefore collisionless. Alternatively, one can switch from the Klimontovich description to the Vlasov–Boltzmann description and introduce an ad hoc order-ϵ\smash{\epsilon} collision operator directly in (93). This will alter the Green’s operator, but the conceptual formulation would remain the same, so it will not be considered separately in detail.

3.4 Summary of section 3

Our QL model is defined as usual, except: (i) we allow for a general particle Hamiltonian H\smash{H}; (ii) we use the Klimontovich equation rather than the Vlasov equation to retain collisions; (iii) we use local averaging (denoted with overbar) and allow for weak inhomogeneity of all averaged quantities; (iv) we retain the initial conditions g\smash{g} for the oscillating part of the distribution function (defined as in (108) but yet to be calculated explicitly). Then, the average part of the distribution function satisfies

tf¯{H¯,f¯}=α(D^αββf¯)+Γ,\displaystyle\partial_{t}\overline{f}-\{\overline{H},\overline{f}\}=\partial_{\alpha}(\widehat{D}^{\alpha\beta}\partial_{\beta}\overline{f})+\Gamma, (121)

where D^αβu^αG^u^β¯\smash{\widehat{D}^{\alpha\beta}\doteq\overline{\widehat{u}^{\alpha}\widehat{G}\widehat{u}^{\beta}}}, Γα(uαg¯)\smash{\Gamma\doteq-\partial_{\alpha}(\overline{u^{\alpha}g})}, uα\smash{u^{\alpha}} is the wave-driven perturbation of the phase-space velocity (see (92)), u^α\smash{\widehat{u}^{\alpha}} is the same quantity considered as an operator on X\smash{\mathscr{H}_{X}} (see (114)), and G^\smash{\widehat{G}} is the ‘effective’ Green’s operator given by

G^limν0+0dτeνττVaa.\displaystyle\widehat{G}\doteq\lim_{\nu\to 0+}\int_{0}^{\infty}\mathrm{d}\tau\,\mathrm{e}^{-\nu\tau-\tau V^{a}\partial_{a}}. (122)

Also, α/zα\smash{\partial_{\alpha}\equiv\partial/\partial z^{\alpha}}, and {,}\smash{\{\sqbullet,\sqbullet\}} is the Poisson bracket on the particle phase space 𝒛\smash{{\boldsymbol{z}}}. The equation for f¯\smash{\overline{f}} used in the standard QLT is recovered from (121) by neglecting Γ\smash{\Gamma} and the spatial gradients (in particular, the whole Poisson bracket) and also by approximating the operator D^αβ\smash{\widehat{D}^{\alpha\beta}} with a local function of 𝒛\smash{{\boldsymbol{z}}}.

4 Preliminaries

Before we start calculating the functions in (121) explicitly, let us get some preliminaries out of the way. In this section, we discuss the shift operators T^τ\smash{\widehat{T}_{\tau}} (section 4.1), approximate the operator G^\smash{\widehat{G}} (section 4.2), and develop a model for the function gg that encodes the initial conditions for f~\smash{\widetilde{f}} (section 4.3).

4.1 Shift operator

Here, we derive some properties of the shift operator T^τ\smash{\widehat{T}_{\tau}} introduced in section 3.2.

4.1.1 T^τ{\widehat{T}_{\tau}} as a shift

Here, we formally prove (an admittedly obvious fact) that

T^τψ(𝑿)=ψ(𝑿τ(𝑿)),τa(𝑿)0τdtVa(𝒀(t,𝑿)),\displaystyle\widehat{T}_{\tau}\psi({\boldsymbol{X}})=\psi({\boldsymbol{X}}-{\boldsymbol{\ell}}_{\tau}({\boldsymbol{X}})),\qquad\textstyle\ell_{\tau}^{a}({\boldsymbol{X}})\doteq\int_{0}^{\tau}\mathrm{d}t\,V^{a}({\boldsymbol{Y}}(t,{\boldsymbol{X}})), (123)

where the ‘characteristics’ Ya\smash{Y^{a}} solve171717In terms of ttτ\smash{t^{\prime}\doteq t-\tau}, (124) has a more recognizable form dYa/dt=Va(𝒀)\smash{\mathrm{d}Y^{a}/\mathrm{d}t^{\prime}=V^{a}({\boldsymbol{Y}})}, with Ya(t=t)=Xa\smash{Y^{a}(t^{\prime}=t)=X^{a}}.

dYadτ=Va(𝒀),Ya(τ=0)=Xa,\displaystyle\frac{\mathrm{d}Y^{a}}{\mathrm{d}\tau}=-V^{a}({\boldsymbol{Y}}),\qquad Y^{a}(\tau=0)=X^{a}, (124)

and thus τa\ell_{\tau}^{a} can be Taylor-expanded in τ\tau as

τa(𝑿)=τVa12τ2VbbVa+,VaVa(𝑿).\displaystyle\ell_{\tau}^{a}({\boldsymbol{X}})=\tau V^{a}-\frac{1}{2}\,\tau^{2}V^{b}\partial_{b}V^{a}+\ldots,\qquad V^{a}\equiv V^{a}({\boldsymbol{X}}). (125)

As the first step to proving (123), let us Taylor-expand VaV^{a} around a fixed point 𝑿1{\boldsymbol{X}}_{1}:

Va=V1a+(bV1a)δXb+,δXaXaX1a,\displaystyle V^{a}=V^{a}_{1}+(\partial_{b}V^{a}_{1})\,\delta X^{b}+\ldots,\qquad\delta X^{a}\doteq X^{a}-X_{1}^{a}, (126)

where V1aVa(𝑿1)V^{a}_{1}\equiv V^{a}({\boldsymbol{X}}_{1}). If one neglects the first and higher derivatives of VaV^{a}, one obtains

T^τψ(𝑿)eτV1aaψ(𝑿)=ψ(𝑿τ𝑽1).\displaystyle\widehat{T}_{\tau}\psi({\boldsymbol{X}})\approx\mathrm{e}^{-\tau V^{a}_{1}\partial_{a}}\psi({\boldsymbol{X}})=\psi({\boldsymbol{X}}-\tau{\boldsymbol{V}}_{1}). (127)

By taking the limit 𝑿1𝑿{\boldsymbol{X}}_{1}\to{\boldsymbol{X}}, which corresponds to 𝑽1𝑽{\boldsymbol{V}}_{1}\to{\boldsymbol{V}}, one obtains

T^τψ(𝑿)=ψ(𝑿τ𝑽)+𝒪(τ2).\displaystyle\widehat{T}_{\tau}\psi({\boldsymbol{X}})=\psi({\boldsymbol{X}}-\tau{\boldsymbol{V}})+\mathcal{O}(\tau^{2}). (128)

Similarly, if one neglects the second and higher derivatives of VaV^{a}, one obtains181818We use the Zassenhaus formula eA^+B^=eA^eB^e[A^,B^]/2e[B^,[A^,B^]]/3+[A^,[A^,B^]]/6\smash{\mathrm{e}^{\widehat{A}+\widehat{B}}=\mathrm{e}^{\widehat{A}}\,\mathrm{e}^{\widehat{B}}\,\mathrm{e}^{-[\widehat{A},\widehat{B}]/2}\mathrm{e}^{[\widehat{B},[\widehat{A},\widehat{B}]]/3+[\widehat{A},[\widehat{A},\widehat{B}]]/6}\cdots}

T^τψ(𝑿)\displaystyle\widehat{T}_{\tau}\psi({\boldsymbol{X}}) =eτ(V1a+(bV1a)δXb+)aψ(𝑿)\displaystyle=\mathrm{e}^{-\tau(V^{a}_{1}+(\partial_{b}V^{a}_{1})\delta X^{b}+\ldots)\partial_{a}}\psi({\boldsymbol{X}})
eτ(bV1a)δXbaeτV1aae12[τ(bV1a)δXba,τV1cc]ψ(𝑿)\displaystyle\approx\mathrm{e}^{-\tau(\partial_{b}V^{a}_{1})\delta X^{b}\partial_{a}}\,\mathrm{e}^{-\tau V^{a}_{1}\partial_{a}}\,\mathrm{e}^{-\frac{1}{2}[-\tau(\partial_{b}V^{a}_{1})\delta X^{b}\partial_{a},-\tau V^{c}_{1}\partial_{c}]}\psi({\boldsymbol{X}})
eτ(bV1a)δXbaeτV1aae12τ2V1c(bV1a)[c,δXba]ψ(𝑿)\displaystyle\approx\mathrm{e}^{-\tau(\partial_{b}V^{a}_{1})\delta X^{b}\partial_{a}}\,\mathrm{e}^{-\tau V^{a}_{1}\partial_{a}}\,\mathrm{e}^{\frac{1}{2}\tau^{2}V^{c}_{1}(\partial_{b}V^{a}_{1})[\partial_{c},\delta X^{b}\partial_{a}]}\psi({\boldsymbol{X}})
eτ(bV1a)δXbaeτV1aae12τ2V1b(bV1a)aψ(𝑿)\displaystyle\approx\mathrm{e}^{-\tau(\partial_{b}V^{a}_{1})\delta X^{b}\partial_{a}}\,\mathrm{e}^{-\tau V^{a}_{1}\partial_{a}}\,\mathrm{e}^{\frac{1}{2}\tau^{2}V^{b}_{1}(\partial_{b}V^{a}_{1})\partial_{a}}\psi({\boldsymbol{X}})
eτ(bV1a)δXbaeτV1aa+12τ2V1b(bV1a)aψ(𝑿)\displaystyle\approx\mathrm{e}^{-\tau(\partial_{b}V^{a}_{1})\delta X^{b}\partial_{a}}\,\mathrm{e}^{-\tau V^{a}_{1}\partial_{a}+\frac{1}{2}\tau^{2}V^{b}_{1}(\partial_{b}V^{a}_{1})\partial_{a}}\psi({\boldsymbol{X}})
eτ(bV1a)δXbaψ(𝑿τ𝑽1+12τ2V1bb𝑽1).\displaystyle\approx\mathrm{e}^{-\tau(\partial_{b}V^{a}_{1})\delta X^{b}\partial_{a}}\psi({\boldsymbol{X}}-\tau{\boldsymbol{V}}_{1}+\textstyle\frac{1}{2}\,\tau^{2}V^{b}_{1}\partial_{b}{\boldsymbol{V}}_{1}). (129)

In the limit 𝑿1𝑿{\boldsymbol{X}}_{1}\to{\boldsymbol{X}}, when eτ(bV1a)δXba1\smash{\mathrm{e}^{-\tau(\partial_{b}V^{a}_{1})\delta X^{b}\partial_{a}}\to 1} and 𝑽1𝑽{\boldsymbol{V}}_{1}\to{\boldsymbol{V}}, one obtains

T^τψ(𝑿)=ψ(𝑿τ𝑽(𝑿)+12τ2(𝑽𝑿)𝑽)+𝒪(τ3).\displaystyle\widehat{T}_{\tau}\psi({\boldsymbol{X}})=\psi\left({\boldsymbol{X}}-\tau{\boldsymbol{V}}({\boldsymbol{X}})+\frac{1}{2}\,\tau^{2}({\boldsymbol{V}}\cdot\partial_{{\boldsymbol{X}}}){\boldsymbol{V}}\right)+\mathcal{O}(\tau^{3}). (130)

In conjunction with (125), equations (128) and (130) show agreement with the sought result (125) within the assumed accuracy. One can also retain 𝗆\mathsf{m} derivatives of 𝑽{\boldsymbol{V}} and derive the corresponding approximations similarly. Then the error will be 𝒪(τ𝗆+2)\smash{\mathcal{O}(\tau^{\mathsf{m}+2})}.

For an order-one time interval τ\tau, one can split this interval on Nτ1N_{\tau}\gg 1 subintervals of small duration τ/Nτ\tau/N_{\tau} and apply finite-𝗆\mathsf{m} formulas (for example, (128) or (130)) to those. Then the total error scales as 𝒪(Nτ𝗆1)\smash{\mathcal{O}(N_{\tau}^{-\mathsf{m}-1})} and the exact formula (123) is obtained at NτN_{\tau}\to\infty.

4.1.2 Symbol of T^τ{\widehat{T}_{\tau}}

Using the bra-ket notation, (123) can be written as

𝑿|T^τ|ψ=𝑿τ(𝑿)|ψ.\displaystyle\braket{{\boldsymbol{X}}}{\widehat{T}_{\tau}}{\psi}=\braket{{\boldsymbol{X}}-{\boldsymbol{\ell}}_{\tau}({\boldsymbol{X}})}{\psi}. (131)

Thus, 𝑿|T^τ=𝑿τ(𝑿)|\smash{\bra{{\boldsymbol{X}}}\widehat{T}_{\tau}=\bra{{\boldsymbol{X}}-{\boldsymbol{\ell}}_{\tau}({\boldsymbol{X}})}}, so

𝑿1|T^τ|𝑿2=𝑿1τ(𝑿1)|𝑿2=δ(𝑿1𝑿2τ(𝑿1)).\displaystyle\braket{{\boldsymbol{X}}_{1}}{\widehat{T}_{\tau}}{{\boldsymbol{X}}_{2}}=\braket{{\boldsymbol{X}}_{1}-{\boldsymbol{\ell}}_{\tau}({\boldsymbol{X}}_{1})}{{\boldsymbol{X}}_{2}}=\delta({\boldsymbol{X}}_{1}-{\boldsymbol{X}}_{2}-{\boldsymbol{\ell}}_{\tau}({\boldsymbol{X}}_{1})). (132)

Using (70), one obtains the Weyl symbol of T^τ\smash{\widehat{T}_{\tau}} in the form

Tτ(𝑿,𝑲)=d𝑺ei𝑲𝑺δ(𝑺τ(𝑿+𝑺/2)).\displaystyle T_{\tau}({\boldsymbol{X}},{\boldsymbol{K}})=\int\mathrm{d}{\boldsymbol{S}}\,\mathrm{e}^{-\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}\,\delta({\boldsymbol{S}}-{\boldsymbol{\ell}}_{\tau}({\boldsymbol{X}}+{\boldsymbol{S}}/2)). (133)

From (125), one has

τa(𝑿+𝑺/2)\displaystyle\ell^{a}_{\tau}({\boldsymbol{X}}+{\boldsymbol{S}}/2) =τVa(𝑿+𝑺/2)(τ2/2)VbbVa+𝒪(ϵ2)\displaystyle=\tau V^{a}({\boldsymbol{X}}+{\boldsymbol{S}}/2)-(\tau^{2}/2)\,V^{b}\partial_{b}V^{a}+\mathcal{O}(\epsilon^{2})
=τVa+(τ/2)(bVa)Sb(τ2/2)VbbVa+𝒪(ϵ2)\displaystyle=\tau V^{a}+(\tau/2)(\partial_{b}V^{a})S^{b}-(\tau^{2}/2)\,V^{b}\partial_{b}V^{a}+\mathcal{O}(\epsilon^{2})
=MabVbτ+mabSb+𝒪(ϵ2),\displaystyle={M^{a}}_{b}V^{b}\tau+{m^{a}}_{b}S^{b}+\mathcal{O}(\epsilon^{2}), (134)

where we introduced a matrix 𝑴𝟏𝒎\smash{{\boldsymbol{M}}\doteq{\boldsymbol{1}}-{\boldsymbol{m}}}, or explicitly,

Mabδbamab,mab(τ/2)(bVa).\displaystyle{M^{a}}_{b}\doteq\delta^{a}_{b}-{m^{a}}_{b},\qquad{m^{a}}_{b}\doteq(\tau/2)(\partial_{b}V^{a}). (135)

Let us express the term 𝒪(ϵ2)\smash{\mathcal{O}(\epsilon^{2})} in (134) as Mabμb\smash{-{M^{a}}_{b}\mu^{b}}. Then,

δ(𝑺τ(𝑿+𝑺/2))\displaystyle\delta({\boldsymbol{S}}-{\boldsymbol{\ell}}_{\tau}({\boldsymbol{X}}+{\boldsymbol{S}}/2)) =δ(𝑺𝑴𝑽τ𝒎𝑺+𝑴𝝁)\displaystyle=\delta({\boldsymbol{S}}-{\boldsymbol{M}}{\boldsymbol{V}}\tau-{\boldsymbol{m}}{\boldsymbol{S}}+{\boldsymbol{M}}{\boldsymbol{\mu}})
=δ(𝑴(𝑺𝑽τ+𝝁))\displaystyle=\delta({\boldsymbol{M}}({\boldsymbol{S}}-{\boldsymbol{V}}\tau+{\boldsymbol{\mu}}))
=δ(𝑺𝑽τ+𝝁)/|det𝑴|.\displaystyle=\delta({\boldsymbol{S}}-{\boldsymbol{V}}\tau+{\boldsymbol{\mu}})/|\det{\boldsymbol{M}}|. (136)

Because 𝒎=𝒪(ϵ)\smash{{\boldsymbol{m}}=\mathcal{O}(\epsilon)}, the well-known formula yields det𝑴=1+tr𝒎+𝒪(ϵ2)\smash{\det{\boldsymbol{M}}=1+\operatorname{tr}{\boldsymbol{m}}+\mathcal{O}(\epsilon^{2})}. But tr𝒎=0\operatorname{tr}{\boldsymbol{m}}=0 by (99), so

δ(𝑺τ(𝑿+𝑺/2))=δ(𝑺𝑽τ+𝝁)+𝒪(ϵ2).\displaystyle\delta({\boldsymbol{S}}-{\boldsymbol{\ell}}_{\tau}({\boldsymbol{X}}+{\boldsymbol{S}}/2))=\delta({\boldsymbol{S}}-{\boldsymbol{V}}\tau+{\boldsymbol{\mu}})+\mathcal{O}(\epsilon^{2}). (137)

The last term 𝒪(ϵ2)\smash{\mathcal{O}(\epsilon^{2})} is insignificant and can be neglected right away, so (133) leads to

Tτ(𝑿,𝑲)exp(iτΩ(𝑿,𝑲)+i𝑲𝝁),\displaystyle T_{\tau}({\boldsymbol{X}},{\boldsymbol{K}})\approx\exp(\mathrm{i}\tau\Omega({\boldsymbol{X}},{\boldsymbol{K}})+\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{\mu}}), (138)

where we have introduced the following notation:

Ω(𝑿,𝑲)𝑲𝑽(𝑿)=ωqαvα=ω𝒌𝒗+𝒪(ϵ).\displaystyle\Omega({\boldsymbol{X}},{\boldsymbol{K}})\doteq-{\boldsymbol{K}}\cdot{\boldsymbol{V}}({\boldsymbol{X}})=\omega-q_{\alpha}v^{\alpha}=\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}+\mathcal{O}(\epsilon). (139)

By definition, 𝝁\smash{{\boldsymbol{\mu}}} is a polynomial of τ\smash{\tau} with coefficients that are of order ϵ2\smash{\epsilon^{2}} and therefore small. But because τ\smash{\tau} can be large, and because 𝝁\smash{{\boldsymbol{\mu}}} is under the exponent, this makes Tτ\smash{T_{\tau}} potentially sensitive to this term, so we retain it (for now).

4.2 Effective Green’s operator

The effective Green’s operator (109) can be understood as the right inverse of the operator (cf. section 3.2)

L^efflimν0+(t{H¯,}+ν),\displaystyle\widehat{L}_{\text{eff}}\doteq\lim_{\nu\to 0+}(\partial_{t}-\{\overline{H},\sqbullet\}+\nu), (140)

so we denote it also as G^=L^eff1\widehat{G}=\widehat{L}^{-1}_{\text{eff}} (which is admittedly abuse of notation). Because L^eff\smash{\widehat{L}_{\text{eff}}} has real 𝑿{\boldsymbol{X}} representation by definition, the 𝑿{\boldsymbol{X}} representation of G^\smash{\widehat{G}} is real too. In particular, 𝑿+𝑺/2|G^|𝑿𝑺/2\smash{\braket{{\boldsymbol{X}}+{\boldsymbol{S}}/2}{\widehat{G}}{{\boldsymbol{X}}-{\boldsymbol{S}}/2}} is real, hence

G(𝑿,𝑲)=G(𝑿,𝑲)\displaystyle G({\boldsymbol{X}},-{\boldsymbol{K}})=G^{*}({\boldsymbol{X}},{\boldsymbol{K}}) (141)

by definition of the Weyl symbol (70). As a corollary, the derivative of G(𝑿,𝑲)G({\boldsymbol{X}},{\boldsymbol{K}}) with respect to the aath component of the whole second argument, denoted G|aG^{|a}, satisfies

(G|a(𝑿,𝑲))=G|a(𝑿,𝑲).\displaystyle(G^{|a}({\boldsymbol{X}},{\boldsymbol{K}}))^{*}=-G^{|a}({\boldsymbol{X}},-{\boldsymbol{K}}). (142)

Also note that G^\widehat{G} can be expressed as

G^=limν0+i(ω^x˙ik^ip˙ir^i+iν)1\displaystyle\widehat{G}=\lim_{\nu\to 0+}\mathrm{i}(\widehat{\omega}-\dot{x}^{i}\widehat{k}_{i}-\dot{p}_{i}\widehat{r}^{i}+\mathrm{i}\nu)^{-1} (143)

(the notation ‘limν0+A(ω+iν)\smash{\lim_{\nu\to 0+}A(\omega+\mathrm{i}\nu)}’ will also be shortened as ‘A(ω+i0)\smash{A(\omega+\mathrm{i}0)}’), whence

Gri=𝒪(p˙i)=𝒪(ϵ).\displaystyle\frac{\partial G}{\partial r^{i}}=\mathcal{O}(\dot{p}_{i})=\mathcal{O}(\epsilon). (144)

Due to (138), the leading-order approximation of the symbol of the operator (109) is G(𝑿,𝑲)=G0(Ω(𝑿,𝑲))G({\boldsymbol{X}},{\boldsymbol{K}})=G_{0}(\Omega({\boldsymbol{X}},{\boldsymbol{K}})), where

G0(Ω)limν0+0dτeντ+iΩτ=\upiδ(Ω)+ipv1Ω\displaystyle G_{0}(\Omega)\doteq\lim_{\nu\to 0+}\int_{0}^{\infty}\mathrm{d}\tau\,\mathrm{e}^{-\nu\tau+\mathrm{i}\Omega\tau}=\upi\,\delta(\Omega)+\mathrm{i}\,\operatorname{pv}\frac{1}{\Omega} (145)

and the (standard) notation pv(1/Ω)\operatorname{pv}(1/\Omega) is defined as follows:

pv1Ωlimν0+Ων2+Ω2.\displaystyle\operatorname{pv}\frac{1}{\Omega}\doteq\lim_{\nu\to 0+}\frac{\Omega}{\nu^{2}+\Omega^{2}}. (146)

This means, in particular, that for any AA, one has

𝒥[A,G0]\displaystyle\mathcal{J}[A,G_{0}] d𝑲A(𝑿,𝑲)G0(Ω(𝑿,𝑲))\displaystyle\doteq\int\mathrm{d}{\boldsymbol{K}}\,A({\boldsymbol{X}},{\boldsymbol{K}})G_{0}(\Omega({\boldsymbol{X}},{\boldsymbol{K}}))
=\upid𝑲A(𝑿,𝑲)δ(Ω(𝑿,𝑲))+id𝑲A(𝑿,𝑲)Ω(𝑿,𝑲),\displaystyle=\upi\int\mathrm{d}{\boldsymbol{K}}\,A({\boldsymbol{X}},{\boldsymbol{K}})\delta(\Omega({\boldsymbol{X}},{\boldsymbol{K}}))+\mathrm{i}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{A({\boldsymbol{X}},{\boldsymbol{K}})}{\Omega({\boldsymbol{X}},{\boldsymbol{K}})}, (147)

where \fint is a principal-value integral. Also usefully, G¯0=G0\smash{\overline{G}_{0}=G_{0}} and

a𝒥[A,G0]\displaystyle\partial_{a}\mathcal{J}[A,G_{0}] =d𝑲A(𝑿,𝑲)G0(Ω(𝑿,𝑲))aΩ(𝑿,𝑲)\displaystyle=\int\mathrm{d}{\boldsymbol{K}}\,A({\boldsymbol{X}},{\boldsymbol{K}})G_{0}^{\prime}(\Omega({\boldsymbol{X}},{\boldsymbol{K}}))\,\partial_{a}\Omega({\boldsymbol{X}},{\boldsymbol{K}})
=(aVb(𝑿))d𝑲KbA(𝑿,𝑲)G0(Ω(𝑿,𝑲))\displaystyle=-(\partial_{a}V^{b}({\boldsymbol{X}}))\int\mathrm{d}{\boldsymbol{K}}\,K_{b}A({\boldsymbol{X}},{\boldsymbol{K}})G_{0}^{\prime}(\Omega({\boldsymbol{X}},{\boldsymbol{K}}))
=(aVb(𝑿))ðΩd𝑲KbA(𝑿,𝑲)G0(Ω(𝑿,𝑲)),\displaystyle=-(\partial_{a}V^{b}({\boldsymbol{X}}))\,\frac{\eth}{\partial\Omega}\int\mathrm{d}{\boldsymbol{K}}\,K_{b}A({\boldsymbol{X}},{\boldsymbol{K}})G_{0}(\Omega({\boldsymbol{X}},{\boldsymbol{K}})), (148)

where the notation ð/λðλ\eth/\partial\lambda\equiv\eth_{\lambda} is defined, for any λ\smash{\lambda} and QQ, as follows:

ðλQ(λ)(ϑQ(λ+ϑ))ϑ=0.\displaystyle\frac{\eth}{\partial\lambda}\int Q(\lambda)\doteq\left(\frac{\partial}{\partial\vartheta}\,\int Q(\lambda+\vartheta)\right)_{\vartheta=0}. (149)

Now let us reinstate the term 𝝁{\boldsymbol{\mu}} in (138). It is readily seen (appendix B.2) that although 𝝁{\boldsymbol{\mu}} may significantly affect Tτ\smash{T_{\tau}} per se, its effect on 𝒥[A,G]\mathcal{J}[A,G] is small, namely,

𝒥[A,G]𝒥[A,G0]=𝒪(ϵ2).\displaystyle\mathcal{J}[A,G]-\mathcal{J}[A,G_{0}]=\mathcal{O}(\epsilon^{2}). (150)

Below, we apply this formulation to A=𝒪(ε2)\smash{A=\mathcal{O}(\varepsilon^{2})}, in which case (150) becomes 𝒥[A,G]𝒥[A,G0]=𝒪(ϵ2ε2)\smash{\mathcal{J}[A,G]-\mathcal{J}[A,G_{0}]=\mathcal{O}(\epsilon^{2}\varepsilon^{2})}. Such corrections are negligible within our model, so from now on we adopt

G(𝑿,𝑲)G¯(𝑿,𝑲)G0(Ω(𝑿,𝑲)).\displaystyle G({\boldsymbol{X}},{\boldsymbol{K}})\approx\overline{G}({\boldsymbol{X}},{\boldsymbol{K}})\approx G_{0}(\Omega({\boldsymbol{X}},{\boldsymbol{K}})). (151)

4.3 Initial conditions

Consider the function gg from (108). Using (110), the latter can be written as follows:

g=limν0+limτ0(T^ττ0f~(𝑿)+0ττ0dτ(1eντ)T^τ(𝑿)).\displaystyle g=\lim_{\nu\to 0+}\lim_{\tau_{0}\to-\infty}\left(\widehat{T}_{\tau-\tau_{0}}\widetilde{f}({\boldsymbol{X}})+\int_{0}^{\tau-\tau_{0}}\mathrm{d}\tau^{\prime}\,(1-\mathrm{e}^{-\nu\tau^{\prime}})\widehat{T}_{\tau^{\prime}}\mathscr{F}({\boldsymbol{X}})\right). (152)

Because (1eντ)(1-\mathrm{e}^{-\nu\tau}) is smooth and T^τ\smash{\widehat{T}_{\tau}\mathscr{F}} is rapidly oscillating, the second term in the external parenthesis is an oscillatory function of τ0\tau_{0} with the average negligible at ν0\smash{\nu\to 0}. But the whole expression in these parenthesis is independent of τ0\tau_{0} at large τ0\tau_{0} (section 3.2). Thus, it can be replaced with its own average over τ0\tau_{0}, denoted τ0\smash{\langle\ldots\rangle_{\tau_{0}}}. Because there is no ν\smash{\nu}-dependence left in this case, one can also omit limν0+\smash{\lim_{\nu\to 0+}}. That gives

g=limτ0T^ττ0f~(𝑿)τ0.\displaystyle g=\lim_{\tau_{0}\to-\infty}\langle\widehat{T}_{\tau-\tau_{0}}\widetilde{f}({\boldsymbol{X}})\rangle_{\tau_{0}}. (153)

Using

f(𝑿)=σδ(𝒛𝒛σ(t)),𝔣(𝑿)σδ(𝒛𝒛¯σ(t)),\displaystyle f({\boldsymbol{X}})=\sum_{\sigma}\delta({\boldsymbol{z}}-{\boldsymbol{z}}_{\sigma}(t)),\qquad\mathfrak{f}({\boldsymbol{X}})\doteq\sum_{\sigma}\delta({\boldsymbol{z}}-\overline{{\boldsymbol{z}}}_{\sigma}(t)), (154)

where the sum is taken over individual particles, one can write191919Taylor-expanding delta functions is admittedly a questionable procedure, but here it is understood as a shorthand for Taylor-expanding integrals of f\smash{f}.

f(𝑿)𝔣(𝑿)σ𝒛~σ(𝑿)𝒛δ(𝒛𝒛¯σ(𝑿)),\displaystyle\textstyle f({\boldsymbol{X}})\approx\mathfrak{f}({\boldsymbol{X}})-\sum_{\sigma}\widetilde{{\boldsymbol{z}}}_{\sigma}({\boldsymbol{X}})\partial_{{\boldsymbol{z}}}\delta({\boldsymbol{z}}-\overline{{\boldsymbol{z}}}_{\sigma}({\boldsymbol{X}})), (155)

where 𝒛σ𝒛σ𝒛¯σ\smash{{\boldsymbol{z}}_{\sigma}\doteq{\boldsymbol{z}}_{\sigma}-\overline{{\boldsymbol{z}}}_{\sigma}} are the H~\smash{\widetilde{H}}-driven small deviations from the particle unperturbed trajectories 𝒛¯σ\smash{\overline{{\boldsymbol{z}}}_{\sigma}}. Then, f¯=𝔣¯(𝑿)\smash{\overline{f}=\overline{\mathfrak{f}}({\boldsymbol{X}})}, and the linearized perturbation f~ff¯\smash{\widetilde{f}\doteq f-\overline{f}} is given by

f~(𝑿)=𝔣(𝑿)𝔣¯(𝑿) ~f ~  σ𝒛~σ(𝑿)𝒛δ(𝒛𝒛¯σ(𝑿))f¯~.\displaystyle\textstyle\widetilde{f}({\boldsymbol{X}})=\underbrace{\,\mathfrak{f}({\boldsymbol{X}})-\overline{\mathfrak{f}}({\boldsymbol{X}})}_{\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}}\underbrace{\textstyle-\sum_{\sigma}\widetilde{{\boldsymbol{z}}}_{\sigma}({\boldsymbol{X}})\partial_{{\boldsymbol{z}}}\delta({\boldsymbol{z}}-\overline{{\boldsymbol{z}}}_{\sigma}({\boldsymbol{X}}))}_{\underline{\widetilde{f}}}. (156)

By definition, the unperturbed trajectories 𝒛¯σ\smash{\overline{{\boldsymbol{z}}}_{\sigma}} satisfy L^δ(𝒛𝒛¯σ(𝑿))=0\smash{\widehat{L}\delta({\boldsymbol{z}}-\overline{{\boldsymbol{z}}}_{\sigma}({\boldsymbol{X}}))=0}, where L^\smash{\widehat{L}} as in (98); thus,

T^ττ0 ~f ~  =eL^(ττ0) ~f ~  = ~f ~  .\displaystyle\widehat{T}_{\tau-\tau_{0}}\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}=\mathrm{e}^{\widehat{L}(\tau-\tau_{0})}\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}=\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}. (157)

Also, T^ττ0f¯~τ0=0\smash{\langle\widehat{T}_{\tau-\tau_{0}}\underline{\widetilde{f}}\rangle_{\tau_{0}}=0}, because 𝒛~σ\smash{\widetilde{{\boldsymbol{z}}}_{\sigma}} are oscillatory functions of 𝑿{\boldsymbol{X}} that is slowly evolved by T^ττ0\smash{\widehat{T}_{\tau-\tau_{0}}}. Hence, g\smash{g} is the microscopic part of the unperturbed distribution function:

g= ~f ~  =𝔣(𝑿)𝔣¯(𝑿).\displaystyle g=\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{f}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}=\mathfrak{f}({\boldsymbol{X}})-\overline{\mathfrak{f}}({\boldsymbol{X}}). (158)

This indicates that the term Γ\smash{\Gamma} defined in (118) is due to collisional effects. We postpone discussing these effects until section 6, so Γ\smash{\Gamma} will be ignored for now.

4.4 Summary of section 4

The main result of this section is that the Weyl symbol of the effective Green’s operator G^\smash{\widehat{G}} can be approximated within the assumed accuracy as follows:

G(𝑿,𝑲)G0(Ω(𝑿,𝑲)),Ω(𝑿,𝑲)𝑲𝑽(𝑿).\displaystyle G({\boldsymbol{X}},{\boldsymbol{K}})\approx G_{0}(\Omega({\boldsymbol{X}},{\boldsymbol{K}})),\qquad\Omega({\boldsymbol{X}},{\boldsymbol{K}})\doteq-{\boldsymbol{K}}\cdot{\boldsymbol{V}}({\boldsymbol{X}}). (159)

Here, 𝑽\smash{{\boldsymbol{V}}} is the unperturbed velocity in the 𝑿\smash{{\boldsymbol{X}}} space, so Ω(𝑿,𝑲)=ω𝒌𝒗+𝒪(ϵ)\smash{\Omega({\boldsymbol{X}},{\boldsymbol{K}})=\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}+\mathcal{O}(\epsilon)}, where 𝒗\smash{{\boldsymbol{v}}} is the unperturbed velocity in the 𝒙\smash{{\boldsymbol{x}}} space, and

G0(Ω)=\upiδ(Ω)+ipv1Ω,pv1Ωlimν0+Ων2+Ω2.\displaystyle G_{0}(\Omega)=\upi\,\delta(\Omega)+\mathrm{i}\,\operatorname{pv}\frac{1}{\Omega},\qquad\operatorname{pv}\frac{1}{\Omega}\doteq\lim_{\nu\to 0+}\frac{\Omega}{\nu^{2}+\Omega^{2}}. (160)

We also show that the term Γ\smash{\Gamma} defined in (118) is due to collisional effects. We postpone discussing these effects until section 6, so Γ\smash{\Gamma} will be ignored for now.

5 Interaction with prescribed fields

In this section, we explore the effect of the diffusion operator D^αβ\smash{\widehat{D}^{\alpha\beta}}. The oscillations will be described by W¯\smash{\overline{W}} as a prescribed function, so they are allowed (yet not required) to be ‘off-shell’, i.e. do not have to be constrained by a dispersion relation. Examples of off-shell fluctuations include driven near-field oscillations, evanescent waves, and microscopic fluctuations (see also section 6). We will first derive the symbol of D^αβ\smash{\widehat{D}^{\alpha\beta}} and, using this symbol, approximate the diffusion operator with a differential operator (section 5.1). Then, we will calculate the coefficients in the approximate expression for D^αβ\smash{\widehat{D}^{\alpha\beta}} (sections 5.2 and 5.3). Finally, we will introduce the concept of the OC distribution (section 5.4) and summarize and simplify the resulting equations (section 5.6).

5.1 Expansion of the dispersion operator

The (effective) Green’s operator can be represented through its symbol GG using (71):

G^=1(2\upi)Nd𝑿d𝑲d𝑺|𝑿+𝑺/2G(𝑿,𝑲)𝑿𝑺/2|ei𝑲𝑺.\displaystyle\widehat{G}=\frac{1}{(2\upi)^{N}}\int\mathrm{d}{\boldsymbol{X}}\,\mathrm{d}{\boldsymbol{K}}\,\mathrm{d}{\boldsymbol{S}}\ket{{\boldsymbol{X}}+{\boldsymbol{S}}/2}G({\boldsymbol{X}},{\boldsymbol{K}})\bra{{\boldsymbol{X}}-{\boldsymbol{S}}/2}\mathrm{e}^{\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}. (161)

The corresponding representation of u^α\widehat{u}^{\alpha} is even simpler, because the symbol of u^α\widehat{u}^{\alpha} is independent of 𝑲{\boldsymbol{K}}:202020One can also derive (162) formally from (71).

u^α=d𝑿|𝑿uα(𝑿)𝑿|.\displaystyle\widehat{u}^{\alpha}=\int\mathrm{d}{\boldsymbol{X}}\ket{{\boldsymbol{X}}}u^{\alpha}({\boldsymbol{X}})\bra{{\boldsymbol{X}}}. (162)

Let us also introduce the Wigner matrix of uαu^{\alpha}, denoted W𝒖αβ\smash{W_{{\boldsymbol{u}}}^{\alpha\beta}}, and its inverse Fourier transform C𝒖αβ\smash{C_{{\boldsymbol{u}}}^{\alpha\beta}} as in section 2.2.3. Using these together with (67), one obtains

(2\upi)Nu^αG^u^β\displaystyle(2\upi)^{N}\widehat{u}^{\alpha}\widehat{G}\widehat{u}^{\beta} =d𝑿d𝑿′′d𝑿d𝑲d𝑺uα(𝑿)uβ(𝑿′′)G(𝑿,𝑲)ei𝑲𝑺\displaystyle=\int\mathrm{d}{\boldsymbol{X}}^{\prime}\,\mathrm{d}{\boldsymbol{X}}^{\prime\prime}\,\mathrm{d}{\boldsymbol{X}}\,\mathrm{d}{\boldsymbol{K}}\,\mathrm{d}{\boldsymbol{S}}\,u^{\alpha}({\boldsymbol{X}}^{\prime})u^{\beta}({\boldsymbol{X}}^{\prime\prime})G({\boldsymbol{X}},{\boldsymbol{K}})\,\mathrm{e}^{\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}
×|𝑿𝑿|𝑿+𝑺/2𝑿𝑺/2|𝑿′′𝑿′′|\displaystyle\qquad\times\ket{{\boldsymbol{X}}^{\prime}}\braket{{\boldsymbol{X}}^{\prime}}{{\boldsymbol{X}}+{\boldsymbol{S}}/2}\braket{{\boldsymbol{X}}-{\boldsymbol{S}}/2}{{\boldsymbol{X}}^{\prime\prime}}\bra{{\boldsymbol{X}}^{\prime\prime}}
=d𝑿d𝑲d𝑺C𝒖αβ(𝑿,𝑺)G(𝑿,𝑲)ei𝑲𝑺|𝑿+𝑺/2𝑿𝑺/2|\displaystyle=\int\mathrm{d}{\boldsymbol{X}}\,\mathrm{d}{\boldsymbol{K}}\,\mathrm{d}{\boldsymbol{S}}\,{C}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{S}})G({\boldsymbol{X}},{\boldsymbol{K}})\,\mathrm{e}^{\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}\ket{{\boldsymbol{X}}+{\boldsymbol{S}}/2}\bra{{\boldsymbol{X}}-{\boldsymbol{S}}/2}
=d𝑿d𝑲d𝑲′′d𝑺W𝒖αβ(𝑿,𝑲)G(𝑿,𝑲′′)ei(𝑲+𝑲′′)𝑺\displaystyle=\int\mathrm{d}{\boldsymbol{X}}^{\prime}\,\mathrm{d}{\boldsymbol{K}}^{\prime}\,\mathrm{d}{\boldsymbol{K}}^{\prime\prime}\,\mathrm{d}{\boldsymbol{S}}^{\prime}\,{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}}^{\prime},{\boldsymbol{K}}^{\prime})G({\boldsymbol{X}}^{\prime},{\boldsymbol{K}}^{\prime\prime})\,\mathrm{e}^{\mathrm{i}({\boldsymbol{K}}^{\prime}+{\boldsymbol{K}}^{\prime\prime})\cdot{\boldsymbol{S}}^{\prime}}
×|𝑿+𝑺/2𝑿𝑺/2|.\displaystyle\qquad\times\ket{{\boldsymbol{X}}^{\prime}+{\boldsymbol{S}}^{\prime}/2}\bra{{\boldsymbol{X}}^{\prime}-{\boldsymbol{S}}^{\prime}/2}. (163)

Then, by taking symbX\smash{\text{symb}_{X}} of (163), one finds that the symbol of D^αβ\smash{\widehat{D}^{\alpha\beta}} is a convolution of W¯𝒖αβ\smash{\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}} and G\smash{G} (appendix B.3):

Dαβ(𝑿,𝑲)=d𝑲W¯𝒖αβ(𝑿,𝑲)G(𝑿,𝑲𝑲).\displaystyle D^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}})=\int\mathrm{d}{\boldsymbol{K}}^{\prime}\,\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}}^{\prime})G({\boldsymbol{X}},{\boldsymbol{K}}-{\boldsymbol{K}}^{\prime}). (164)

Let us Taylor-expand the symbol (164) in 𝑲{\boldsymbol{K}}:

Dαβ(𝑿,𝑲)\displaystyle D^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}})\approx d𝑲W¯𝒖αβ(𝑿,𝑲)G(𝑿,𝑲)\displaystyle\int\mathrm{d}{\boldsymbol{K}}^{\prime}\,\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}}^{\prime})G({\boldsymbol{X}},-{\boldsymbol{K}}^{\prime})
+Kcd𝑲W¯𝒖αβ(𝑿,𝑲)G|c(𝑿,𝑲)+𝒪(KaKbG|ab).\displaystyle+K_{c}\int\mathrm{d}{\boldsymbol{K}}^{\prime}\,\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}}^{\prime})G^{|c}({\boldsymbol{X}},-{\boldsymbol{K}}^{\prime})+\mathcal{O}(K_{a}K_{b}G^{|ab}). (165)

As a reminder, G|a(𝑿,𝑲)=aG(𝑿,𝑲)G^{|a}({\boldsymbol{X}},-{\boldsymbol{K}})=-\partial^{a}G({\boldsymbol{X}},-{\boldsymbol{K}}) denotes the derivative of GG with respect to (the aath component of) the whole second argument, Ka-K_{a}, and

KaGKa=ωGω+kiGki+riGri.\displaystyle K_{a}\,\frac{\partial G}{\partial K_{a}}=\omega\,\frac{\partial G}{\partial\omega}+k_{i}\,\frac{\partial G}{\partial k_{i}}+r^{i}\,\frac{\partial G}{\partial r^{i}}. (166)

Upon application of operX\smash{\text{oper}_{X}}, ω\omega gets replaced (roughly) with it=𝒪(ϵ)\mathrm{i}\partial_{t}=\mathcal{O}(\epsilon) and kik_{i} gets replaced (also roughly) with ii=𝒪(ϵ)-\mathrm{i}\partial_{i}=\mathcal{O}(\epsilon). By (144), the last term in (166) is of order ϵ\epsilon too. This means that the contribution of the whole KaaG\smash{K_{a}\partial^{a}G} term to the equation for f¯\smash{\overline{f}} is of order ϵ\epsilon. The standard QLT neglects this contribution entirely, i.e. adopts Dαβ(𝑿,𝑲)Dαβ(𝑿,𝟎)\smash{D^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}})\approx D^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{0}})}, in which case the diffusion operator becomes just a local function of phase-space variables, D^αβDαβ(𝑿,𝟎)\smash{\widehat{D}^{\alpha\beta}\approx D^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{0}})}. In this work, we retain corrections to the first order in 𝑿{\boldsymbol{X}}, i.e. keep the second term in (165) as well, while neglecting the higher-order terms as usual.

Within this model, one can rewrite (165) as follows:

Dαβ(𝑿,𝑲)D0αβ(𝑿)+KcΘαβc(𝑿).\displaystyle D^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}})\approx D_{0}^{\alpha\beta}({\boldsymbol{X}})+K_{c}\Theta^{\alpha\beta c}({\boldsymbol{X}}). (167)

Here, we used (141) and introduced

D0αβ(𝑿)d𝑲W¯𝒖αβ(𝑿,𝑲)G(𝑿,𝑲),\displaystyle\displaystyle D_{0}^{\alpha\beta}({\boldsymbol{X}})\doteq\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}})G^{*}({\boldsymbol{X}},{\boldsymbol{K}}), (168)
Θαβc(𝑿)d𝑲W¯𝒖αβ(𝑿,𝑲)(G|c(𝑿,𝑲)),\displaystyle\displaystyle\Theta^{\alpha\beta c}({\boldsymbol{X}})\doteq-\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}})(G^{|c}({\boldsymbol{X}},{\boldsymbol{K}}))^{*}, (169)

which satisfy (appendix B.4)

D0αβ(𝑿)=(D0αβ(𝑿)),Θαβc(𝑿)=(Θαβc(𝑿)).\displaystyle D_{0}^{\alpha\beta}({\boldsymbol{X}})=(D_{0}^{\alpha\beta}({\boldsymbol{X}}))^{*},\qquad\Theta^{\alpha\beta c}({\boldsymbol{X}})=-(\Theta^{\alpha\beta c}({\boldsymbol{X}}))^{*}. (170)

The first-order Weyl expansion of D^αβ\smash{\widehat{D}^{\alpha\beta}} is obtained by applying operX\smash{\text{oper}_{X}} to (167). Namely, for any ψ\psi, one has (cf. section 2.1.5)

D^αβψD0αβψiΘαβccψi2(cΘαβc)ψ.\displaystyle\widehat{D}^{\alpha\beta}\psi\approx D_{0}^{\alpha\beta}\psi-\mathrm{i}\Theta^{\alpha\beta c}\partial_{c}\psi-\frac{\mathrm{i}}{2}\,(\partial_{c}\Theta^{\alpha\beta c})\psi. (171)

What remains now is to calculate the functions D0αβ\smash{D_{0}^{\alpha\beta}} and Θαβc\smash{\Theta^{\alpha\beta c}} explicitly.

5.2 Wigner matrix of the velocity oscillations

To express D^αβ\smash{\widehat{D}^{\alpha\beta}} through the Wigner function WW of the perturbation Hamiltonian (section 3.1.1), we need to express W𝒖αβ\smash{W_{{\boldsymbol{u}}}^{\alpha\beta}} through WW. Recall that W𝒖αβ\smash{W_{{\boldsymbol{u}}}^{\alpha\beta}} is the symbol of the density operator W^𝒖αβ(2\upi)N|uαuβ|\smash{\widehat{W}_{{\boldsymbol{u}}}^{\alpha\beta}\doteq(2\upi)^{-N}\ket{u^{\alpha}}\bra{u^{\beta}}} (section 2.2.3). By definition (92), one has uα=iJαμq^μH~\smash{u^{\alpha}=\mathrm{i}J^{\alpha\mu}\widehat{q}_{\mu}\widetilde{H}}, where q^αiα\smash{\widehat{q}_{\alpha}\doteq-\mathrm{i}\partial_{\alpha}} (section 2.2.1). Then,

W^𝒖αβ=(2\upi)NJαμq^μ|H~H~|q^νJβν=JαμJβνq^μW^q^ν,\displaystyle\widehat{W}_{{\boldsymbol{u}}}^{\alpha\beta}=(2\upi)^{-N}J^{\alpha\mu}\widehat{q}_{\mu}\ket{\widetilde{H}}\bra{\widetilde{H}}\widehat{q}_{\nu}J^{\beta\nu}=J^{\alpha\mu}J^{\beta\nu}\widehat{q}_{\mu}\widehat{W}\widehat{q}_{\nu}, (172)

where W^\widehat{W} is the density operator whose symbol is WW. By applying symbX\smash{\text{symb}_{X}}, one obtains

W𝒖αβ=JαμJβν(qμ🟊W🟊qν),\displaystyle W_{{\boldsymbol{u}}}^{\alpha\beta}=J^{\alpha\mu}J^{\beta\nu}(q_{\mu}\bigstar W\bigstar q_{\nu}), (173)

where 🟊\bigstar is the Moyal product (72). Using formulas analogous to (33) in the (𝑿,𝑲)\smash{({\boldsymbol{X}},{\boldsymbol{K}})} space, one obtains

qμ🟊W🟊qν\displaystyle q_{\mu}\bigstar W\bigstar q_{\nu} =(qμWi2Wzμ)🟊qν\displaystyle=\left(q_{\mu}W-\frac{\mathrm{i}}{2}\frac{\partial W}{\partial z^{\mu}}\right)\bigstar q_{\nu}
=qμqνWi2qνWzμ+i2zν(qμWhi2Wzμ)\displaystyle=q_{\mu}q_{\nu}W-\frac{\mathrm{i}}{2}\,q_{\nu}\,\frac{\partial W}{\partial z^{\mu}}+\frac{\mathrm{i}}{2}\frac{\partial}{\partial z^{\nu}}\left(q_{\mu}W_{h}-\frac{\mathrm{i}}{2}\frac{\partial W}{\partial z^{\mu}}\right)
=qμqνWh+i2(qμWzνqνWzμ)+142Wzμzν.\displaystyle=q_{\mu}q_{\nu}W_{h}+\frac{\mathrm{i}}{2}\left(q_{\mu}\frac{\partial W}{\partial z^{\nu}}-q_{\nu}\frac{\partial W}{\partial z^{\mu}}\right)+\frac{1}{4}\frac{\partial^{2}W}{\partial z^{\mu}\partial z^{\nu}}. (174)

Hence, W𝒖αβW_{{\boldsymbol{u}}}^{\alpha\beta} and WW are connected via the following exact formula:

W𝒖αβ=JαμJβν(qμqνWi2(qνWzμqμWzν)+142Wzμzν).\displaystyle W_{{\boldsymbol{u}}}^{\alpha\beta}=J^{\alpha\mu}J^{\beta\nu}\left(q_{\mu}q_{\nu}W-\frac{\mathrm{i}}{2}\left(q_{\nu}\,\frac{\partial W}{\partial z^{\mu}}-q_{\mu}\,\frac{\partial W}{\partial z^{\nu}}\right)+\frac{1}{4}\frac{\partial^{2}W}{\partial z^{\mu}\partial z^{\nu}}\right). (175)

5.3 Nonlinear potentials

Due to (170), one has D0αβ=reD0αβD_{0}^{\alpha\beta}=\operatorname{re}D_{0}^{\alpha\beta}. Using this together with (168), (175), (151), and (145), one obtains

D0αβ\displaystyle D_{0}^{\alpha\beta} =JαμJβνred𝑲(\upiδ(Ω)ipv1Ω).\displaystyle=J^{\alpha\mu}J^{\beta\nu}\operatorname{re}\int\mathrm{d}{\boldsymbol{K}}\left(\upi\,\delta(\Omega)-\mathrm{i}\,\operatorname{pv}\frac{1}{\Omega}\right).
×(qμqνW¯i2(qνW¯zμqμW¯zν)+142W¯zμzν),\displaystyle\hskip 71.13188pt\times\left(q_{\mu}q_{\nu}\overline{W}-\frac{\mathrm{i}}{2}\left(q_{\nu}\,\frac{\partial\overline{W}}{\partial z^{\mu}}-q_{\mu}\,\frac{\partial\overline{W}}{\partial z^{\nu}}\right)+\frac{1}{4}\frac{\partial^{2}\overline{W}}{\partial z^{\mu}\partial z^{\nu}}\right), (176)

with notation as in (10). This can be written as D0αβ=𝖣αβ+ϱαβ+ςαβ\smash{D_{0}^{\alpha\beta}=\mathsf{D}^{\alpha\beta}+\varrho^{\alpha\beta}+\varsigma^{\alpha\beta}}, where

𝖣αβJαμJβνd𝑲\upiδ(Ω)qμqνW¯,\displaystyle\mathsf{D}^{\alpha\beta}\doteq J^{\alpha\mu}J^{\beta\nu}\int\mathrm{d}{\boldsymbol{K}}\,\upi\,\delta(\Omega)\,q_{\mu}q_{\nu}\overline{W}, (177)

and we also introduced

ϱαβ12JαμJβνd𝑲(qνW¯zμqμW¯zν)1Ω,\displaystyle\displaystyle\varrho^{\alpha\beta}\doteq-\frac{1}{2}\,J^{\alpha\mu}J^{\beta\nu}\fint\mathrm{d}{\boldsymbol{K}}\left(q_{\nu}\,\frac{\partial\overline{W}}{\partial z^{\mu}}-q_{\mu}\,\frac{\partial\overline{W}}{\partial z^{\nu}}\right)\frac{1}{\Omega}, (178)
ςαβ14JαμJβνd𝑲\upiδ(Ω)2W¯zμzν.\displaystyle\displaystyle\varsigma^{\alpha\beta}\doteq\frac{1}{4}\,J^{\alpha\mu}J^{\beta\nu}\int\mathrm{d}{\boldsymbol{K}}\,\upi\,\delta(\Omega)\,\frac{\partial^{2}\overline{W}}{\partial z^{\mu}\partial z^{\nu}}. (179)

As shown in appendix B.5, the contributions of these two functions to (117) are

zα(ϱαβf¯zβ)=𝒪(ϵε2),zα(ςαβf¯zβ)=𝒪(ϵ2ε2).\displaystyle\frac{\partial}{\partial z^{\alpha}}\left(\varrho^{\alpha\beta}\,\frac{\partial\overline{f}}{\partial z^{\beta}}\right)=\mathcal{O}(\epsilon\varepsilon^{2}),\qquad\frac{\partial}{\partial z^{\alpha}}\left(\varsigma^{\alpha\beta}\,\frac{\partial\overline{f}}{\partial z^{\beta}}\right)=\mathcal{O}(\epsilon^{2}\varepsilon^{2}). (180)

Thus, ϱαβ\smash{\varrho^{\alpha\beta}} must be retained and ςαβ\smash{\varsigma^{\alpha\beta}} must be neglected, which leads to

D0αβ𝖣αβ+ϱαβ.\displaystyle D_{0}^{\alpha\beta}\approx\mathsf{D}^{\alpha\beta}+\varrho^{\alpha\beta}. (181)

The function Θαβc=iimΘαβc\smash{\Theta^{\alpha\beta c}=\mathrm{i}\,\operatorname{im}\Theta^{\alpha\beta c}} can be written as follows:

Θαβc(𝑿)=iJαμJβνd𝑲qμqνW¯(𝑿,𝑲)Kcpv1Ω(𝑿,𝑲)=iVc(𝑿)Θαβ(𝑿),\displaystyle\Theta^{\alpha\beta c}({\boldsymbol{X}})=\mathrm{i}J^{\alpha\mu}J^{\beta\nu}\int\mathrm{d}{\boldsymbol{K}}\,q_{\mu}q_{\nu}\overline{W}({\boldsymbol{X}},{\boldsymbol{K}})\,\frac{\partial}{\partial K_{c}}\,\operatorname{pv}\frac{1}{\Omega({\boldsymbol{X}},{\boldsymbol{K}})}=-\mathrm{i}V^{c}({\boldsymbol{X}})\,\Uptheta^{\alpha\beta}({\boldsymbol{X}}),

where we introduced

ΘαβJαμJβνðΩd𝑲qμqνW¯Ω\displaystyle\Uptheta^{\alpha\beta}\doteq J^{\alpha\mu}J^{\beta\nu}\frac{\eth}{\partial\Omega}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\mu}q_{\nu}\overline{W}}{\Omega} (182)

and ð\smash{\eth} is defined as in (149). Then finally, one can rewrite (171) as follows:

D^αβψ\displaystyle\widehat{D}^{\alpha\beta}\psi (𝖣αβ+ϱαβ)ψΘαβVccψ12Vc(cΘαβ)ψ\displaystyle\approx(\mathsf{D}^{\alpha\beta}+\varrho^{\alpha\beta})\psi-\Uptheta^{\alpha\beta}V^{c}\partial_{c}\psi-\frac{1}{2}\,V^{c}(\partial_{c}\Uptheta^{\alpha\beta})\psi
=(𝖣αβ+ϱαβ)ψΘαβ(t+vλλ)ψ12((t+vλλ)Θαβ)ψ,\displaystyle=(\mathsf{D}^{\alpha\beta}+\varrho^{\alpha\beta})\psi-\Uptheta^{\alpha\beta}(\partial_{t}+v^{\lambda}\partial_{\lambda})\psi-\frac{1}{2}\,((\partial_{t}+v^{\lambda}\partial_{\lambda})\Uptheta^{\alpha\beta})\psi, (183)

where we used (99). With some algebra (appendix B.6), and assuming the notation

Φ=Jμνzμd𝑲qνW¯2Ω,\displaystyle\Phi=-J^{\mu\nu}\,\frac{\partial}{\partial z^{\mu}}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\nu}\overline{W}}{2\Omega}, (184)

one finds that (183) leads to

α(D^αββf¯)=α(𝖣αββf¯)12dtα(Θαββf¯)+{Φ,f¯}.\displaystyle\partial_{\alpha}(\widehat{D}^{\alpha\beta}\partial_{\beta}\overline{f})=\partial_{\alpha}(\mathsf{D}^{\alpha\beta}\partial_{\beta}\overline{f})-\frac{1}{2}\,\mathrm{d}_{t}\partial_{\alpha}(\Uptheta^{\alpha\beta}\partial_{\beta}\overline{f})+\{\Phi,\overline{f}\}. (185)

Hence, (117) becomes (to the extent that Γ\Gamma is negligible; see section 6.7)

dtf¯+12dtα(Θαββf¯){Φ,f¯}=α(𝖣αββf¯).\displaystyle\mathrm{d}_{t}\overline{f}+\frac{1}{2}\,\mathrm{d}_{t}\partial_{\alpha}(\Uptheta^{\alpha\beta}\partial_{\beta}\overline{f})-\{\Phi,\overline{f}\}=\partial_{\alpha}(\mathsf{D}^{\alpha\beta}\partial_{\beta}\overline{f}). (186)

The functions Θαβ\smash{\Uptheta^{\alpha\beta}}, Φ\smash{\Phi}, and 𝖣αβ\smash{\mathsf{D}^{\alpha\beta}} that determine the coefficients in this equation are fundamental and, for the lack of a better term, will be called nonlinear potentials.

5.4 Oscillation-center distribution

Let us introduce

Ff¯+12α(Θαββf¯).\displaystyle F\doteq\overline{f}+\frac{1}{2}\,\partial_{\alpha}(\Uptheta^{\alpha\beta}\partial_{\beta}\overline{f}). (187)

Then, using (185), one can rewrite (186) as212121The difference between F\smash{F} and f¯\smash{\overline{f}} is related to the concept of so-called adiabatic diffusion (Galeev & Sagdeev, 1985; Stix, 1992), which captures some but not all adiabatic effects.

tF{,F}=α(𝖣αββF),\displaystyle\partial_{t}F-\{\mathcal{H},F\}=\partial_{\alpha}(\mathsf{D}^{\alpha\beta}\partial_{\beta}F), (188)

where corrections 𝒪(ε4)\mathcal{O}(\varepsilon^{4}) have been neglected and we introduced H¯+Φ\smash{\mathcal{H}\doteq\overline{H}+\Phi}. As a reminder, the nonlinear potentials in (188) are as follows:

𝖣αβ\displaystyle\mathsf{D}^{\alpha\beta} =JαμJβνd𝑲\upiδ(Ω)qμqνW¯,\displaystyle=J^{\alpha\mu}J^{\beta\nu}\int\mathrm{d}{\boldsymbol{K}}\,\upi\,\delta(\Omega)\,q_{\mu}q_{\nu}\overline{W}, (189)
Θαβ\displaystyle\Uptheta^{\alpha\beta} =JαμJβνðΩd𝑲qμqνW¯Ω,\displaystyle=J^{\alpha\mu}J^{\beta\nu}\frac{\eth}{\partial\Omega}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\mu}q_{\nu}\overline{W}}{\Omega}, (190)
Φ\displaystyle\Phi =Jμνzμd𝑲qνW¯2Ω.\displaystyle=-J^{\mu\nu}\,\frac{\partial}{\partial z^{\mu}}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\nu}\overline{W}}{2\Omega}. (191)

Equations (187)–(191) form a closed model that describes the evolution of the average distribution f¯\smash{\overline{f}} in turbulence with prescribed W¯\overline{W}. In particular, (188) can be interpreted as a Liouville-type equation for FF as an effective, or ‘dressed’, distribution. The latter can be understood as the distribution of ‘dressed’ particles called OCs. Then, \mathcal{H} serves as the OC Hamiltonian, 𝖣αβ\smash{\mathsf{D}^{\alpha\beta}} is the phase-space diffusion coefficient, Φ\smash{\Phi} is the ponderomotive energy, Ω=ωqαvα\smash{\Omega=\omega-q_{\alpha}v^{\alpha}}, and vαJαββH¯\smash{v^{\alpha}\doteq J^{\alpha\beta}\partial_{\beta}\overline{H}}. Within the assumed accuracy, one can redefine vα\smash{v^{\alpha}} to be the OC velocity rather than the particle velocity; specifically,222222The advantage of the amended definition (192) is that it will lead to exact conservation laws of our theory, as to be discussed in section 7.5.

vαJαββ=JαββH¯+𝒪(ϵ2).\displaystyle v^{\alpha}\doteq J^{\alpha\beta}\partial_{\beta}\mathcal{H}=J^{\alpha\beta}\partial_{\beta}\overline{H}+\mathcal{O}(\epsilon^{2}). (192)

Then, the presence of δ(Ω)\smash{\delta(\Omega)} in (189) signifies that OCs diffuse in phase space in response to waves they are resonant with. Below, we use the terms ‘OCs’ and ‘particles’ interchangeably except where specified otherwise.

That said, the interpretation of OCs as particle-like objects is limited. Single-OC motion equations are not introduced in our approach. (They would have been singular for resonant interactions.) Accordingly, the transformation (187) of the distribution function f¯F\smash{\overline{f}\mapsto F} is not derived from a coordinate transformation but rather is fundamental. As a result, particles and OCs live in the same phase space, but the ‘dynamics of OCs’ can be irreversible (section 5.5). This qualitatively distinguishes our approach from the traditional OC theory (Dewar, 1973) and from the conceptually similar gyrokinetic theory (Littlejohn, 1981; Cary & Brizard, 2009), where coordinate transformations are central.

5.5 H{H}-theorem

Because W¯\smash{\overline{W}} is non-negative (section 2.1.6), 𝖣αβ\smash{\mathsf{D}^{\alpha\beta}} is positive-semidefinite; that is,

𝖣αβψαψβ=d𝑲\upiδ(Ω)a2W¯0,aJαμψαqμ\displaystyle\mathsf{D}^{\alpha\beta}\psi_{\alpha}\psi_{\beta}=\int\mathrm{d}{\boldsymbol{K}}\,\upi\,\delta(\Omega)\,a^{2}\overline{W}\geq 0,\qquad a\doteq J^{\alpha\mu}\psi_{\alpha}q_{\mu} (193)

for any real ψ\smash{\psi}. This leads to the following theorem. Consider the OC entropy defined as

𝒮d𝒛F(t,𝒛)lnF(t,𝒛).\displaystyle\mathscr{S}\doteq-\int\mathrm{d}{\boldsymbol{z}}\,F(t,{\boldsymbol{z}})\ln F(t,{\boldsymbol{z}}). (194)

According to (188), 𝒮\smash{\mathscr{S}} satisfies

d𝒮dt\displaystyle\frac{\mathrm{d}\mathscr{S}}{\mathrm{d}t} =d𝒛d(FlnF)dF({,F}+α(𝖣αββF))\displaystyle=-\int\mathrm{d}{\boldsymbol{z}}\,\frac{\mathrm{d}(F\ln F)}{\mathrm{d}F}\left(\{\mathcal{H},F\}+\partial_{\alpha}(\mathsf{D}^{\alpha\beta}\partial_{\beta}F)\right)
=d𝒛d(FlnF)dFJαβ(α)(βF)d𝒛(1+lnF)α(𝖣αββF)\displaystyle=-\int\mathrm{d}{\boldsymbol{z}}\,\frac{\mathrm{d}(F\ln F)}{\mathrm{d}F}\,J^{\alpha\beta}(\partial_{\alpha}\mathcal{H})(\partial_{\beta}F)-\int\mathrm{d}{\boldsymbol{z}}\,(1+\ln F)\,\partial_{\alpha}(\mathsf{D}^{\alpha\beta}\partial_{\beta}F)
=d𝒛Jαβ(α)β(FlnF)d𝒛lnFα(𝖣αββF)\displaystyle=-\int\mathrm{d}{\boldsymbol{z}}\,J^{\alpha\beta}(\partial_{\alpha}\mathcal{H})\partial_{\beta}(F\ln F)-\int\mathrm{d}{\boldsymbol{z}}\,\ln F\,\partial_{\alpha}(\mathsf{D}^{\alpha\beta}\partial_{\beta}F)
=d𝒛(Jαβαβ2)FlnF+d𝒛𝖣αβ(αlnF)(βlnF)F.\displaystyle=\int\mathrm{d}{\boldsymbol{z}}\,(J^{\alpha\beta}\partial^{2}_{\alpha\beta}\mathcal{H})\,F\ln F+\int\mathrm{d}{\boldsymbol{z}}\,\mathsf{D}^{\alpha\beta}(\partial_{\alpha}\ln F)(\partial_{\beta}\ln F)F. (195)

The first integral vanishes due to Jαβαβ2=0\smash{J^{\alpha\beta}\partial^{2}_{\alpha\beta}=0}. The second integral is non-negative due to (193). Thus,

d𝒮dt0,\displaystyle\frac{\mathrm{d}\mathscr{S}}{\mathrm{d}t}\geq 0, (196)

which is recognized as the H\smash{H}-theorem (Lifshitz & Pitaevskii, 1981, section 4) for QL OC dynamics.

5.6 Summary of section 5

From now on, we assume that the right-hand side of (188) scales not as 𝒪(ε2)\smash{\mathcal{O}(\varepsilon^{2})} but as 𝒪(ϵε2)\smash{\mathcal{O}(\epsilon\varepsilon^{2})}, either due to the scarcity of resonant particles or, for QL diffusion driven by microscopic fluctuations (section 6), due to the plasma parameter’s being large. Also, the spatial derivatives can be neglected within the assumed accuracy in the definition of FF (187) and on the right-hand side of (188). Using this together with (69), and with (56) for the Poisson bracket, our results can be summarized as follows.

QL evolution of a particle distribution in a prescribed wave field is governed by232323Remember that here we neglect Γ\smash{\Gamma} (118), which is a part of the collision operator to be reinstated in section 6.

Ft𝒙F𝒑+𝒑F𝒙=𝒑(𝗗F𝒑).\displaystyle\frac{\partial F}{\partial t}-\frac{\partial\mathcal{H}}{\partial{\boldsymbol{x}}}\cdot\frac{\partial F}{\partial{\boldsymbol{p}}}+\frac{\partial\mathcal{H}}{\partial{\boldsymbol{p}}}\cdot\frac{\partial F}{\partial{\boldsymbol{x}}}=\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left({\boldsymbol{\mathsf{D}}}\,\frac{\partial F}{\partial{\boldsymbol{p}}}\right). (197)

The OC distribution F\smash{F} is defined as

F=f¯+12𝒑(𝚯f¯𝒑),\displaystyle F=\overline{f}+\frac{1}{2}\,\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left({\boldsymbol{\Uptheta}}\,\frac{\partial\overline{f}}{\partial{\boldsymbol{p}}}\right), (198)

so the density of OCs is the same as the locally averages density of the true particles:

𝒩d𝒑F=d𝒑f¯.\displaystyle\mathcal{N}\doteq\int\mathrm{d}{\boldsymbol{p}}\,F=\int\mathrm{d}{\boldsymbol{p}}\,\overline{f}. (199)

The function \smash{\mathcal{H}} is understood as the OC Hamiltonian. It is given by

H¯+Φ,\displaystyle\mathcal{H}\doteq\overline{H}+\Phi, (200)

where H¯\smash{\overline{H}} is the average Hamiltonian, which may include interaction with background fields, and Φ\smash{\Phi} is the ponderomotive potential. The nonlinear potentials that enter (197) can be calculated to the zeroth order in ϵ\epsilon and are given by242424See section 9 for examples and section 6.6 for the explanation on how Φ\smash{\Phi} is related to Δ\smash{\Delta}, which is yet to be introduced. Also note that in combination with (200), equation (203) generalizes the related results from (Kentwell, 1987; Fraiman & Kostyukov, 1995; Dodin & Fisch, 2014).

𝗗\displaystyle{\boldsymbol{\mathsf{D}}} =dωd𝒌\upi𝒌𝒌𝖶¯(t,𝒌𝒗,𝒌;𝒑),\displaystyle=\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\upi\,{\boldsymbol{k}}{\boldsymbol{k}}\overline{\mathsf{W}}(t,{\boldsymbol{k}}\cdot{\boldsymbol{v}},{\boldsymbol{k}};{\boldsymbol{p}}), (201)
𝚯\displaystyle{\boldsymbol{\Uptheta}} =ϑdωd𝒌𝒌𝒌𝖶¯ω𝒌𝒗+ϑ|ϑ=0,\displaystyle=\frac{\partial}{\partial\vartheta}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\left.\frac{{\boldsymbol{k}}{\boldsymbol{k}}\overline{\mathsf{W}}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}+\vartheta}\right|_{\vartheta=0}, (202)
Φ\displaystyle\Phi =12𝒑dωd𝒌𝒌𝖶¯ω𝒌𝒗,\displaystyle=\frac{1}{2}\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\frac{{\boldsymbol{k}}\overline{\mathsf{W}}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}}, (203)

where 𝒌𝒌\smash{{\boldsymbol{k}}{\boldsymbol{k}}} is a dyadic matrix with two lower indices, and the same conventions apply as in section 2.1.2. Also, 𝒗\smash{{\boldsymbol{v}}} is hereby redefined as the OC spatial velocity, namely,

𝒗𝒑=𝒑H¯+𝒪(ϵ2).\displaystyle{\boldsymbol{v}}\doteq\partial_{\boldsymbol{p}}\mathcal{H}=\partial_{\boldsymbol{p}}\overline{H}+\mathcal{O}(\epsilon^{2}). (204)

The function 𝖶¯\smash{\overline{\mathsf{W}}} is defined as

𝖶¯(t,𝒙,ω,𝒌;𝒑)d𝒓W¯(t,𝒙,𝒑,ω,𝒌,𝒓),\displaystyle\overline{\mathsf{W}}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}})\doteq\int\mathrm{d}{\boldsymbol{r}}\,\overline{W}(t,{\boldsymbol{x}},{\boldsymbol{p}},\omega,{\boldsymbol{k}},{\boldsymbol{r}}), (205)

where W¯\overline{W} is the average Wigner function (88) of the perturbation Hamiltonian, i.e. the spectrum of its symmetrized autocorrelation function (89). Due to (78), it can be understood as the average of 𝖶symb𝗑𝖶^\smash{\mathsf{W}\doteq\text{symb}_{\mathsf{x}}\widehat{\mathsf{W}}} (where 𝖶^\smash{\widehat{\mathsf{W}}} is defined in (87)), i.e. as the Wigner function of the perturbation Hamiltonian with 𝒑\smash{{\boldsymbol{p}}} treated as a parameter. As such, 𝖶¯\smash{\overline{\mathsf{W}}} is non-negative, so 𝗗\smash{{\boldsymbol{\mathsf{D}}}} is positive-semidefinite. This leads to an H\smash{H}-theorem (proven similarly to (196)) for the entropy density σd𝒑FlnF\smash{\sigma\doteq-\int\mathrm{d}{\boldsymbol{p}}\,F\ln F}:

(dσdt)𝖣0,(ψt)𝖣𝒑(𝗗ψ𝒑).\displaystyle\left(\frac{\mathrm{d}\sigma}{\mathrm{d}t}\right)_{\mathsf{D}}\geq 0,\qquad\left(\frac{\partial\psi}{\partial t}\right)_{\mathsf{D}}\doteq\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left({\boldsymbol{\mathsf{D}}}\,\frac{\partial\psi}{\partial{\boldsymbol{p}}}\right). (206)

Also note that for homogeneous turbulence in particular, where 𝖶¯\smash{\overline{\mathsf{W}}} is independent of 𝒙\smash{{\boldsymbol{x}}}, (46) yields that

dω𝖶¯(t,𝒙,ω,𝒌;𝒑)=1𝒱ndωd𝒙𝖶¯(t,𝒙,ω,𝒌;𝒑)=1𝒱n|H~̊(t,𝒌,𝒑)|2¯,\displaystyle\int\mathrm{d}\omega\,\overline{\mathsf{W}}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}})=\frac{1}{\mathscr{V}_{n}}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{x}}\,\overline{\mathsf{W}}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}})=\frac{1}{\mathscr{V}_{n}}\,\overline{|\mathring{\widetilde{H}}(t,{\boldsymbol{k}},{\boldsymbol{p}})|^{2}}, (207)

where 𝒱n\smash{\mathscr{V}_{n}} is the plasma volume (the index n\smash{n} denotes the number of spatial dimensions) and H~̊\smash{\mathring{\widetilde{H}}} is the spatial spectrum of H~\smash{\widetilde{H}} as defined in (23).

Equation (197) can be used to calculate the ponderomotive force td𝒑f¯\smash{\partial_{t}\int\mathrm{d}{\boldsymbol{p}}\,\overline{f}} that a given wave field imparts on a plasma. This potentially resolves the controversies mentioned in (Kentwell & Jones, 1987). We will revisit this subject for on-shell waves in section 7.5.

6 Interaction with self-consistent fields

Here, we explain how to calculate the function 𝖶¯\overline{\mathsf{W}} in the presence of microscopic fluctuations (nonzero gg). In particular, we reinstate the term Γ\smash{\Gamma} that was omitted in section 5. We also show that a collision operator of the Balescu–Lenard type emerges from our theory within a general interaction model. This calculation can be considered as a generalization of that in (Rogister & Oberman, 1968) for homogeneous plasmas. Another related calculation was proposed in (Chavanis, 2012) in application to potential interactions in inhomogeneous systems using action–angle variables, with global averaging over the angles. (See also (Mynick, 1988) for a related calculation in action–angle variables based on the Fokker–Planck approach.) In contrast, our model holds for any Hamiltonian interactions via any vector fields and allows for weak inhomogeneities in canonical coordinates.

6.1 Interaction model

Let us assume that particles interact via an MM-component real field 𝚿(Ψ1,Ψ2,,ΨM){\boldsymbol{\Psi}}\equiv(\Psi^{1},\Psi^{2},\ldots,\Psi^{M})^{\intercal}. It is treated below as a column vector; hence the index . (A complex field can be accommodated by considering its real and imaginary parts as separate components.) We split this field into the average part 𝚿¯\smash{\overline{{\boldsymbol{\Psi}}}} and the oscillating part 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}}. The former is considered given. For the latter, we assume the action integral of this field without plasma in the form

S0=d𝘅𝔏0,𝔏0=12𝚿~𝚵^0𝚿~\displaystyle S_{0}=\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\mathfrak{L}_{0},\qquad\mathfrak{L}_{0}=\frac{1}{2}\,\smash{\widetilde{{\boldsymbol{\Psi}}}}^{\dagger}\widehat{\boldsymbol{\Xi}}_{0}\widetilde{{\boldsymbol{\Psi}}} (208)

(see section 9 for examples), where 𝚵^0\smash{\widehat{\boldsymbol{\Xi}}_{0}} is a Hermitian operator252525The field action often has the form S0=12d𝘅𝔤(𝖌𝚿~)𝚵^0𝚿~\smash{S_{0}=\frac{1}{2}\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\sqrt{\mathfrak{g}}\,({\boldsymbol{\mathfrak{g}}}\smash{\widetilde{{\boldsymbol{\Psi}}}}^{*})\widehat{\boldsymbol{\Xi}}_{0}\widetilde{{\boldsymbol{\Psi}}}}, where 𝖌(𝘅)\smash{{\boldsymbol{\mathfrak{g}}}({\boldsymbol{\mathsf{x}}})} is a spacetime metric, 𝔤|det𝖌|\smash{\mathfrak{g}\doteq|\det{\boldsymbol{\mathfrak{g}}}\,|}, and 𝚵^0\smash{\widehat{\boldsymbol{\Xi}}_{0}} is Hermitian with respect to the inner product 𝝃|𝝍𝔤d𝘅𝔤(𝖌𝝃)𝝍\smash{\langle{\boldsymbol{\xi}}|{\boldsymbol{\psi}}\rangle_{\mathfrak{g}}\doteq\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\sqrt{\mathfrak{g}}\,({\boldsymbol{\mathfrak{g}}}{\boldsymbol{\xi}}^{*}){\boldsymbol{\psi}}}. Using 𝚿~𝔤1/4𝚿~\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}^{\prime}\doteq\mathfrak{g}^{1/4}\widetilde{{\boldsymbol{\Psi}}}} and 𝚵^0𝔤1/4𝖌𝚵^0𝔤1/4\smash{\smash{\widehat{\boldsymbol{\Xi}}}_{0}^{\prime}\doteq\mathfrak{g}^{1/4}{\boldsymbol{\mathfrak{g}}}\widehat{\boldsymbol{\Xi}}_{0}\mathfrak{g}^{-1/4}}, one can cast this action in the form (208), with 𝚵^0\smash{\smash{\widehat{\boldsymbol{\Xi}}}_{0}^{\prime}} that is Hermitian with respect to the inner product (6). and 𝚿~=𝚿~\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}^{\dagger}=\smash{\widetilde{{\boldsymbol{\Psi}}}}^{\intercal}} is a row vector dual to 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}}. Plasma is allowed to consist of multiple species, henceforth denoted with index s\smash{s}. Because 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}} is assumed small, the generic Hamiltonian for each species s\smash{s} can be Taylor-expanded in 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}} and represented in a generic form

Hs(t,𝒙,𝒑)H0s+𝜶^s𝚿~+12(𝑳^s𝚿~)(𝑹^s𝚿~)\displaystyle H_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}})\approx H_{0s}+\widehat{\boldsymbol{\alpha}}_{s}^{\dagger}\widetilde{{\boldsymbol{\Psi}}}+\frac{1}{2}\,(\widehat{\boldsymbol{L}}_{s}\widetilde{{\boldsymbol{\Psi}}})^{\dagger}(\widehat{\boldsymbol{R}}_{s}\widetilde{{\boldsymbol{\Psi}}}) (209)

(see section 9 for examples), which can be considered as a second-order Taylor expansion of the full Hamiltonian in 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}}. Here, H0sH0s(t,𝒙,𝒑)\smash{H_{0s}\equiv H_{0s}(t,{\boldsymbol{x}},{\boldsymbol{p}})} is independent of 𝚿~𝚿~(t,𝒙)\smash{\widetilde{{\boldsymbol{\Psi}}}\equiv\widetilde{{\boldsymbol{\Psi}}}(t,{\boldsymbol{x}})}, 𝜶^s(α^s,1,α^s,2,,α^s,M)\smash{\widehat{\boldsymbol{\alpha}}_{s}\equiv(\widehat{\alpha}_{s,1},\widehat{\alpha}_{s,2},\ldots,\widehat{\alpha}_{s,M})^{\intercal}} is a column vector whose elements α^s,i\smash{\widehat{\alpha}_{s,i}} are linear operators on 𝗑\smash{\mathscr{H}_{\mathsf{x}}}, and the dagger is added so that 𝜶^s\smash{\widehat{\boldsymbol{\alpha}}_{s}^{\dagger}} could be understood as a row vector whose elements α^s,i\smash{\widehat{\alpha}_{s,i}^{\dagger}} act on the individual components of the field; i.e. 𝜶^s𝚿~α^s,iΨ~i\smash{\widehat{\boldsymbol{\alpha}}_{s}^{\dagger}\widetilde{{\boldsymbol{\Psi}}}\equiv\widehat{\alpha}^{\dagger}_{s,i}\widetilde{\Psi}^{i}}. We let 𝜶^s\smash{\widehat{\boldsymbol{\alpha}}_{s}} be nonlocal in tt and 𝒙{\boldsymbol{x}} (for example, 𝜶^s\smash{\widehat{\boldsymbol{\alpha}}_{s}} can be a spacetime derivative or a spacetime integral), and we also let 𝜶^s\smash{\widehat{\boldsymbol{\alpha}}_{s}} depend on 𝒑\smash{{\boldsymbol{p}}} parametrically, so

symb𝜶^s=𝜶s(t,𝒙,ω,𝒌;𝒑).\displaystyle\text{symb}\,\widehat{\boldsymbol{\alpha}}_{s}={\boldsymbol{\alpha}}_{s}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}}). (210)

The matrix operators 𝑳^s\smash{\widehat{\boldsymbol{L}}_{s}} and 𝑹^s\smash{\widehat{\boldsymbol{R}}_{s}} and their symbols 𝑳s\smash{{\boldsymbol{L}}_{s}} and 𝑹s\smash{{\boldsymbol{R}}_{s}} are understood similarly.

The Lagrangian density of the oscillating-field–plasma system is

𝔏p=12𝚿~𝚵^0𝚿~+sσs(𝒑σs𝒙˙σsHs(t,𝒙σs,𝒑σs))δ(𝒙𝒙σs(t)),\displaystyle\mathfrak{L}_{p}=\frac{1}{2}\,\smash{\widetilde{{\boldsymbol{\Psi}}}}^{\dagger}\widehat{\boldsymbol{\Xi}}_{0}\widetilde{{\boldsymbol{\Psi}}}+\sum_{s}\sum_{\sigma_{s}}({\boldsymbol{p}}_{\sigma_{s}}\cdot\dot{{\boldsymbol{x}}}_{\sigma_{s}}-H_{s}(t,{\boldsymbol{x}}_{\sigma_{s}},{\boldsymbol{p}}_{\sigma_{s}}))\delta({\boldsymbol{x}}-{\boldsymbol{x}}_{\sigma_{s}}(t)), (211)

where the sum is taken over individual particles. Note that

σsHs(t,𝒙σs,𝒑σs)δ(𝒙𝒙σs(t))\displaystyle\sum_{\sigma_{s}}H_{s}(t,{\boldsymbol{x}}_{\sigma_{s}},{\boldsymbol{p}}_{\sigma_{s}})\delta({\boldsymbol{x}}-{\boldsymbol{x}}_{\sigma_{s}}(t)) =d𝒑σsδ(𝒛𝒛σs(t))Hs(t,𝒙,𝒑)\displaystyle=\int\mathrm{d}{\boldsymbol{p}}\,\sum_{\sigma_{s}}\delta({\boldsymbol{z}}-{\boldsymbol{z}}_{\sigma_{s}}(t))H_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}})
=d𝒑fs(t,𝒙,𝒑)Hs(t,𝒙,𝒑),\displaystyle=\int\mathrm{d}{\boldsymbol{p}}\,f_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}})H_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}}), (212)

so the 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}}-dependent part of the system action can be written as S=d𝘅𝔏\smash{S=\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\mathfrak{L}} with

𝔏=12𝚿~𝚵^p𝚿~sd𝒑f~s𝜶^s𝚿~,\displaystyle\displaystyle\mathfrak{L}=\frac{1}{2}\,\smash{\widetilde{{\boldsymbol{\Psi}}}}^{\dagger}\widehat{\boldsymbol{\Xi}}_{p}\widetilde{{\boldsymbol{\Psi}}}-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\widetilde{f}_{s}\widehat{\boldsymbol{\alpha}}_{s}^{{\dagger}}\widetilde{{\boldsymbol{\Psi}}}, (213)
𝚵^p𝚵^0sd𝒑𝑳^sfs𝑹^s.\displaystyle\displaystyle\widehat{\boldsymbol{\Xi}}_{p}\doteq\widehat{\boldsymbol{\Xi}}_{0}-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\smash{\widehat{\boldsymbol{L}}}_{s}^{\dagger}f_{s}\widehat{\boldsymbol{R}}_{s}. (214)

(The contribution of f~s\smash{\widetilde{f}_{s}} to the second term in (213) has been omitted because it averages to zero at integration over spacetime and thus does not contribute to S\smash{S}.) This ‘abridged’ action is not sufficient to describe the particle motion, but it is sufficient to describe the dynamics of 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}} at given fs\smash{f_{s}}, as discussed below. The operator 𝚵^p\smash{\widehat{\boldsymbol{\Xi}}_{p}} can be considered Hermitian without loss of generality, because its anti-Hermitian part does not contribute to S\smash{S}. Also, we assume that unless either of 𝑳^\smash{\widehat{\boldsymbol{L}}} and 𝑹^\smash{\widehat{\boldsymbol{R}}} is zero, the high-frequency field has no three-wave resonances, so terms cubic in 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}} can be neglected in S\smash{S};262626This is tacitly assumed already in (209), where cubic terms are neglected. Also note that three-wave interactions that involve resonances between low-frequency oscillations of Fs\smash{F_{s}} and two high-frequency waves, like Raman scattering (Balakin et al., 2016), are still allowed. then,

𝚵^p𝚵^0sd𝒑(𝑳^sFs𝑹^s)H.\displaystyle\widehat{\boldsymbol{\Xi}}_{p}\approx\widehat{\boldsymbol{\Xi}}_{0}-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,(\smash{\widehat{\boldsymbol{L}}}_{s}^{\dagger}F_{s}\widehat{\boldsymbol{R}}_{s})_{\text{H}}. (215)

Using the same assumption, one can also adopt

H¯s=H0s+12(𝑳^s𝚿~)(𝑹^s𝚿~)¯,H~s𝜶^s𝚿~,\displaystyle\overline{H}_{s}=H_{0s}+\frac{1}{2}\,\overline{(\widehat{\boldsymbol{L}}_{s}\widetilde{{\boldsymbol{\Psi}}})^{\dagger}(\widehat{\boldsymbol{R}}_{s}\widetilde{{\boldsymbol{\Psi}}})},\qquad\widetilde{H}_{s}\approx\widehat{\boldsymbol{\alpha}}_{s}^{\dagger}\widetilde{{\boldsymbol{\Psi}}}, (216)

because in the absence of three-wave resonances, the oscillating part of (𝑳^s𝚿~)(𝑹^s𝚿~)\smash{(\widehat{\boldsymbol{L}}_{s}\widetilde{{\boldsymbol{\Psi}}})^{\dagger}(\widehat{\boldsymbol{R}}_{s}\widetilde{{\boldsymbol{\Psi}}})} contributes only 𝒪(ε4)\smash{\mathcal{O}(\varepsilon^{4})} terms to the equation for Fs\smash{F_{s}}.

6.2 Field equations

The Euler–Lagrange equation for 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}} derived from (213) is

𝚵^p𝚿~=sd𝒑𝜶^sf~s.\displaystyle\widehat{\boldsymbol{\Xi}}_{p}\widetilde{{\boldsymbol{\Psi}}}=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\widehat{\boldsymbol{\alpha}}_{s}\widetilde{f}_{s}. (217)

Then, to the extent that the linear approximation for f~s\smash{\tilde{f}_{s}} is sufficient (see below), one finds that the oscillating part of the field satisfies

𝚵^p𝚿~sd𝒑𝜶^sG^s{𝜶^s𝚿~,f¯s}=sd𝒑𝜶^sgs,\displaystyle\widehat{\boldsymbol{\Xi}}_{p}\widetilde{{\boldsymbol{\Psi}}}-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\widehat{\boldsymbol{\alpha}}_{s}\widehat{G}_{s}\{\widehat{\boldsymbol{\alpha}}_{s}^{\dagger}\widetilde{{\boldsymbol{\Psi}}},\overline{f}_{s}\}=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\widehat{\boldsymbol{\alpha}}_{s}g_{s}, (218)

where we used (108). Note that the right-hand side of (218) is determined by microscopic fluctuations gs(t,𝒙,𝒑)g_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}}) (section 4.3). Equation (218) can also be expressed as

𝚵^𝚿~=sd𝒑𝜶^sgs,\displaystyle\widehat{\boldsymbol{\Xi}}\widetilde{{\boldsymbol{\Psi}}}=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\widehat{\boldsymbol{\alpha}}_{s}g_{s}, (219)

where 𝚵^\smash{\widehat{\boldsymbol{\Xi}}} is understood as the plasma dispersion operator and is given by

𝚵^𝚵^psd𝒑𝜶^sG^s{𝜶^s,f¯s},\displaystyle\widehat{\boldsymbol{\Xi}}\doteq\widehat{\boldsymbol{\Xi}}_{p}-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\widehat{\boldsymbol{\alpha}}_{s}\widehat{G}_{s}\{\widehat{\boldsymbol{\alpha}}_{s}^{\dagger}\,\sqbullet\,,\overline{f}_{s}\}, (220)

where \smash{\sqbullet} is a placeholder. The general solution of (219) can be written as

𝚿~=𝚿¯~+ ~Ψ ~  , ~Ψ ~  =sd𝒑𝚵^1𝜶^sgs.\displaystyle\widetilde{{\boldsymbol{\Psi}}}=\underline{\widetilde{{\boldsymbol{\Psi}}}}+\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}},\qquad\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\smash{\widehat{\boldsymbol{\Xi}}}^{-1}\widehat{\boldsymbol{\alpha}}_{s}g_{s}. (221)

Here, 𝚵^1\smash{\smash{\widehat{\boldsymbol{\Xi}}}^{-1}} is the right inverse of 𝚵^\smash{\widehat{\boldsymbol{\Xi}}} (meaning 𝚵^𝚵^1=𝟏^\smash{\widehat{\boldsymbol{\Xi}}\smash{\widehat{\boldsymbol{\Xi}}}^{-1}=\widehat{\boldsymbol{1}}} yet 𝚵^1𝚵^𝟏^\smash{\smash{\widehat{\boldsymbol{\Xi}}}^{-1}\widehat{\boldsymbol{\Xi}}\neq\widehat{\boldsymbol{1}}}) such that 𝚿~\textstyle\widetilde{{\boldsymbol{\Psi}}}  ~\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt} vanishes at zero gg.272727Most generally, the problem of finding 𝚵^1\smash{\smash{\widehat{\boldsymbol{\Xi}}}^{-1}} is the standard problem of calculating the field produced by a given radiation source. The rest of the solution, 𝚿¯~\smash{\underline{\widetilde{{\boldsymbol{\Psi}}}}}, is the macroscopic field that satisfies

𝚵^𝚿¯~=𝟎.\displaystyle\widehat{\boldsymbol{\Xi}}\underline{\widetilde{{\boldsymbol{\Psi}}}}={\boldsymbol{0}}. (222)

In the special case when the dispersion operator is Hermitian (𝚵^=𝚵^H\smash{\widehat{\boldsymbol{\Xi}}=\widehat{\boldsymbol{\Xi}}_{\text{H}}}), (222) also flows from the ‘adiabatic’ macroscopic part of the action S\smash{S}, namely,

Sad12d𝘅𝚿¯~𝚵^H𝚿¯~.\displaystyle S_{\text{ad}}\doteq\frac{1}{2}\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\smash{\underline{\widetilde{{\boldsymbol{\Psi}}}}}^{\dagger}\widehat{\boldsymbol{\Xi}}_{\text{H}}\underline{\widetilde{{\boldsymbol{\Psi}}}}. (223)

Because we have assumed a linear model for fs\smash{f_{s}} in (218), 𝚿¯~\smash{\underline{\widetilde{{\boldsymbol{\Psi}}}}} is decoupled from 𝚿~\textstyle\widetilde{{\boldsymbol{\Psi}}}  ~\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt} , and hence the dynamics of 𝚿¯~\smash{\underline{\widetilde{{\boldsymbol{\Psi}}}}} turns out to be collisionless. This is justified, because collisional dissipation is assumed to be much slower that collisionless dissipation (section 3.3). One can reinstate collisions in (222) by modifying G^s\smash{\widehat{G}_{s}} ad hoc, if necessary. Alternatively, one can avoid separating 𝚿¯~\smash{\underline{\widetilde{{\boldsymbol{\Psi}}}}} and 𝚿~\textstyle\widetilde{{\boldsymbol{\Psi}}}  ~\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt} and, instead, derive an equation for the average Wigner matrix of the whole 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}} (McDonald, 1991). However, this approach is beyond QLT, so it is not considered in this paper.

6.3 Dispersion matrix

As readily seen from the definition (220), the operator 𝚵^\smash{\widehat{\boldsymbol{\Xi}}} can be expressed as

𝚵^=𝚵^pik^jsd𝒑𝜶^sG^s𝜶^sFspj+𝒪(ϵ,ε2).\displaystyle\widehat{\boldsymbol{\Xi}}=\widehat{\boldsymbol{\Xi}}_{p}-\mathrm{i}\widehat{k}_{j}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\widehat{\boldsymbol{\alpha}}_{s}\widehat{G}_{s}\widehat{\boldsymbol{\alpha}}_{s}^{\dagger}\,\frac{\partial F_{s}}{\partial p_{j}}+\mathcal{O}(\epsilon,\varepsilon^{2}). (224)

The corrections caused by nonzero ϵ\epsilon and ε\varepsilon in this formula will be insignificant for our purposes, so they will be neglected. In particular, this means that GssymbXG^s\smash{G_{s}\doteq\text{symb}_{X}\widehat{G}_{s}} can be adopted in the form independent of 𝒓\smash{{\boldsymbol{r}}} (section 4.2):

Gsiω𝒌𝒗s+i0=\upiδ(ω𝒌𝒗s)+ipv1ω𝒌𝒗s.\displaystyle G_{s}\approx\frac{\mathrm{i}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\mathrm{i}0}=\upi\,\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})+\mathrm{i}\,\operatorname{pv}\frac{1}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}. (225)

Then, G^s\smash{\widehat{G}_{s}} can be considered as an operator on 𝗑\smash{\mathscr{H}_{\mathsf{x}}} with 𝒑{\boldsymbol{p}} as a parameter, and symbG^s=Gs(t,𝒙,ω,𝒌;𝒑)\smash{\text{symb}\,\widehat{G}_{s}=G_{s}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}})}. Also,

symb(𝜶^sG^s𝜶^s)=𝜶sGs𝜶s𝜶sGs𝜶s.\displaystyle\text{symb}\,(\widehat{\boldsymbol{\alpha}}_{s}\widehat{G}_{s}\widehat{\boldsymbol{\alpha}}_{s}^{\dagger})={\boldsymbol{\alpha}}_{s}\star G_{s}\star{\boldsymbol{\alpha}}_{s}^{\dagger}\approx{\boldsymbol{\alpha}}_{s}G_{s}{\boldsymbol{\alpha}}_{s}^{\dagger}. (226)

This readily yields the ‘dispersion matrix’ 𝚵symb𝗑𝚵^\smash{{\boldsymbol{\Xi}}\doteq\text{symb}_{\mathsf{x}}\widehat{\boldsymbol{\Xi}}}:

𝚵(ω,𝒌)𝚵p(ω,𝒌)+sd𝒑𝜶s(ω,𝒌;𝒑)𝜶s(ω,𝒌;𝒑)ω𝒌𝒗s+i0𝒌Fs(𝒑)𝒑,\displaystyle\displaystyle{\boldsymbol{\Xi}}(\omega,{\boldsymbol{k}})\approx{\boldsymbol{\Xi}}_{p}(\omega,{\boldsymbol{k}})+\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\frac{{\boldsymbol{\alpha}}_{s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}){\boldsymbol{\alpha}}_{s}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\mathrm{i}0}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}, (227)
𝚵p(ω,𝒌)𝚵0(ω,𝒌)sd𝒑s(ω,𝒌;𝒑)Fs(𝒑)\displaystyle\displaystyle{\boldsymbol{\Xi}}_{p}(\omega,{\boldsymbol{k}})\approx{\boldsymbol{\Xi}}_{0}(\omega,{\boldsymbol{k}})-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{\wp}}_{s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})F_{s}({\boldsymbol{p}}) (228)

(see section 9 for examples). Here, 𝜶s𝜶s\smash{{\boldsymbol{\alpha}}_{s}{\boldsymbol{\alpha}}_{s}^{\dagger}} is a dyadic matrix, and the arguments tt and 𝒙{\boldsymbol{x}} are henceforth omitted for brevity. Also, we introduced the operators ^s=^s\smash{\widehat{\boldsymbol{\wp}}_{s}=\widehat{\boldsymbol{\wp}}_{s}^{\dagger}} and their symbols s=s\smash{{\boldsymbol{\wp}}_{s}={\boldsymbol{\wp}}_{s}^{\dagger}} as

^s(𝑳^s𝑹^)H,ssymb^s(𝑳s𝑹s)H.\displaystyle\widehat{\boldsymbol{\wp}}_{s}\doteq(\smash{\widehat{\boldsymbol{L}}}_{s}^{\dagger}\widehat{\boldsymbol{R}})_{\text{H}},\qquad{\boldsymbol{\wp}}_{s}\doteq\text{symb}\,\widehat{\boldsymbol{\wp}}_{s}\approx({\boldsymbol{L}}_{s}^{\dagger}{\boldsymbol{R}}_{s})_{\text{H}}. (229)

The appearance of +i0+\mathrm{i}0 in the denominator in (227) is related to the Landau rule. (Remember that as arguments of Weyl symbols, ω\omega and 𝒌{\boldsymbol{k}} are real by definition.) The Hermitian and anti-Hermitian parts of the dispersion matrix are

𝚵H(ω,𝒌)\displaystyle{\boldsymbol{\Xi}}_{\text{H}}(\omega,{\boldsymbol{k}}) 𝚵p(ω,𝒌)+sd𝒑𝜶s(ω,𝒌;𝒑)𝜶s(ω,𝒌;𝒑)ω𝒌𝒗s𝒌Fs(𝒑)𝒑,\displaystyle\approx{\boldsymbol{\Xi}}_{p}(\omega,{\boldsymbol{k}})+\sum_{s}\fint\mathrm{d}{\boldsymbol{p}}\,\frac{{\boldsymbol{\alpha}}_{s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}){\boldsymbol{\alpha}}_{s}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}, (230)
𝚵A(ω,𝒌)\displaystyle{\boldsymbol{\Xi}}_{\text{A}}(\omega,{\boldsymbol{k}}) \upisd𝒑𝜶s(ω,𝒌;𝒑)𝜶s(ω,𝒌;𝒑)δ(ω𝒌𝒗s)𝒌Fs(𝒑)𝒑.\displaystyle\approx-\upi\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{\alpha}}_{s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}){\boldsymbol{\alpha}}_{s}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}. (231)

Assuming the notation 𝚵(𝚵)1\smash{{\boldsymbol{\Xi}}^{-{\dagger}}\doteq({\boldsymbol{\Xi}}^{\dagger})^{-1}}, the inverse dispersion matrix can be expressed as

𝚵1=𝚵1𝚵𝚵=𝚵1𝚵H𝚵i𝚵1𝚵A𝚵.\displaystyle{\boldsymbol{\Xi}}^{-1}={\boldsymbol{\Xi}}^{-1}{\boldsymbol{\Xi}}^{{\dagger}}{\boldsymbol{\Xi}}^{-{\dagger}}={\boldsymbol{\Xi}}^{-1}{\boldsymbol{\Xi}}_{\text{H}}{\boldsymbol{\Xi}}^{-{\dagger}}-\mathrm{i}{\boldsymbol{\Xi}}^{-1}{\boldsymbol{\Xi}}_{\text{A}}{\boldsymbol{\Xi}}^{-{\dagger}}. (232)

Because 𝚵=(𝚵1)\smash{{\boldsymbol{\Xi}}^{-{\dagger}}=({\boldsymbol{\Xi}}^{-1})^{\dagger}}, this leads to the following formulas, which we will need later:

(𝚵1)H=𝚵1𝚵H𝚵,(𝚵1)A=𝚵1𝚵A𝚵.\displaystyle({\boldsymbol{\Xi}}^{-1})_{\text{H}}={\boldsymbol{\Xi}}^{-1}{\boldsymbol{\Xi}}_{\text{H}}{\boldsymbol{\Xi}}^{-{\dagger}},\qquad({\boldsymbol{\Xi}}^{-1})_{\text{A}}=-{\boldsymbol{\Xi}}^{-1}{\boldsymbol{\Xi}}_{\text{A}}{\boldsymbol{\Xi}}^{-{\dagger}}. (233)

6.4 Spectrum of microscopic fluctuations

Other objects to be used below are the density operators of the oscillating fields:

𝗪^𝚿¯~(2\upi)𝗇|𝚿¯~𝚿¯~|,𝗪^ ~Ψ ~  (2\upi)𝗇| ~Ψ ~  ~Ψ ~  |,\displaystyle\widehat{\boldsymbol{{\boldsymbol{\mathsf{W}}}}}_{\underline{\widetilde{{\boldsymbol{\Psi}}}}}\doteq(2\upi)^{-{\mathsf{n}}}\,\ket{\underline{\widetilde{{\boldsymbol{\Psi}}}}}\bra{\underline{\widetilde{{\boldsymbol{\Psi}}}}},\qquad\widehat{\boldsymbol{{\boldsymbol{\mathsf{W}}}}}_{\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}}\doteq(2\upi)^{-{\mathsf{n}}}\,\ket{\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}}\bra{\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}}, (234)

and the corresponding average Wigner matrices on (𝘅,𝗸)\smash{({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})}. The former, 𝗨𝗪¯𝚿¯~\smash{{\boldsymbol{\mathsf{U}}}\doteq\overline{{\boldsymbol{\mathsf{W}}}}_{\underline{\widetilde{{\boldsymbol{\Psi}}}}}}, is readily found by definition (51), and the latter, 𝖂𝗪¯ ~Ψ ~  \smash{{\boldsymbol{\mathfrak{W}}}\doteq\overline{{\boldsymbol{\mathsf{W}}}}_{\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}}}, is calculated as follows. Let us consider gs(t,𝒙,𝒑)\smash{g_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}})} as a ket in 𝗑\smash{\mathscr{H}_{\mathsf{x}}}, with 𝒑\smash{{\boldsymbol{p}}} as a parameter. Then, (221) readily yields

𝗪^ ~Ψ ~  =1(2\upi)𝗇s,sd𝒑d𝒑𝚵^1𝜶^s(𝒑)|gs(𝒑)gs(𝒑)|𝜶^s(𝒑)𝚵^.\displaystyle\widehat{\boldsymbol{{\boldsymbol{\mathsf{W}}}}}_{\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}}=\frac{1}{(2\upi)^{\mathsf{n}}}\sum_{s,s^{\prime}}\int\mathrm{d}{\boldsymbol{p}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,\smash{\widehat{\boldsymbol{\Xi}}}^{-1}\widehat{\boldsymbol{\alpha}}_{s}({\boldsymbol{p}})\ket{g_{s}({\boldsymbol{p}})}\bra{g_{s^{\prime}}({\boldsymbol{p}}^{\prime})}\widehat{\boldsymbol{\alpha}}^{\dagger}_{s^{\prime}}({\boldsymbol{p}}^{\prime})\smash{\widehat{\boldsymbol{\Xi}}}^{-{\dagger}}. (235)

By applying symb𝗑\smash{\text{symb}_{\mathsf{x}}} to this, one obtains

𝖂(ω,𝒌)=s,sd𝒑d𝒑𝚵1𝜶s(𝒑)𝕲ss(𝒑,𝒑)𝜶s(𝒑)𝚵,\displaystyle{\boldsymbol{\mathfrak{W}}}(\omega,{\boldsymbol{k}})=\sum_{s,s^{\prime}}\int\mathrm{d}{\boldsymbol{p}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,{\boldsymbol{\Xi}}^{-1}\star{\boldsymbol{\alpha}}_{s}({\boldsymbol{p}})\star{\boldsymbol{\mathfrak{G}}}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\star{\boldsymbol{\alpha}}^{\dagger}_{s^{\prime}}({\boldsymbol{p}}^{\prime})\star{\boldsymbol{\Xi}}^{-{\dagger}}, (236)

where most arguments are omitted for brevity and (appendix B.7)

𝔊ss(𝒑,𝒑)\displaystyle\mathfrak{G}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime}) 1(2\upi)𝗇d𝘀ei𝗸𝘀gs(𝘅+𝘀/2,𝒑)gs(𝘅𝘀/2,𝒑)¯\displaystyle\doteq\frac{1}{(2\upi)^{\mathsf{n}}}\int\mathrm{d}{\boldsymbol{\mathsf{s}}}\,\mathrm{e}^{-\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{s}}}}\,\overline{g_{s}({\boldsymbol{\mathsf{x}}}+{\boldsymbol{\mathsf{s}}}/2,{\boldsymbol{p}})\,g_{s^{\prime}}({\boldsymbol{\mathsf{x}}}-{\boldsymbol{\mathsf{s}}}/2,{\boldsymbol{p}}^{\prime})}
1(2\upi)nδssδ(𝒑𝒑)δ(ω𝒌𝒗s)Fs(𝒑),\displaystyle\approx\frac{1}{(2\upi)^{n}}\,\delta_{ss^{\prime}}\delta({\boldsymbol{p}}-{\boldsymbol{p}}^{\prime})\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})F_{s}({\boldsymbol{p}}), (237)

assuming corrections due to inter-particle correlations are negligible. Then, (236) gives

𝖂(ω,𝒌)=1(2\upi)nsd𝒑\displaystyle{\boldsymbol{\mathfrak{W}}}(\omega,{\boldsymbol{k}})=\frac{1}{(2\upi)^{n}}\sum_{s^{\prime}}\int\mathrm{d}{\boldsymbol{p}}^{\prime}\, δ(ω𝒌𝒗s)Fs(𝒑)\displaystyle\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})F_{s^{\prime}}({\boldsymbol{p}}^{\prime})
×𝚵1(ω,𝒌)(𝜶s𝜶s)(ω,𝒌;𝒑)𝚵(ω,𝒌),\displaystyle\times{\boldsymbol{\Xi}}^{-1}(\omega,{\boldsymbol{k}})({\boldsymbol{\alpha}}_{s^{\prime}}{\boldsymbol{\alpha}}_{s^{\prime}}^{\dagger})(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime}){\boldsymbol{\Xi}}^{-{\dagger}}(\omega,{\boldsymbol{k}}), (238)

where 𝒗s𝒗s(t,𝒙,𝒑)\smash{{\boldsymbol{v}}^{\prime}_{s^{\prime}}\doteq{\boldsymbol{v}}_{s^{\prime}}(t,{\boldsymbol{x}},{\boldsymbol{p}}^{\prime})}. It is readily seen from (238) that 𝖂\smash{{\boldsymbol{\mathfrak{W}}}} is positive-semidefinite. One can also recognize (238) as a manifestation of the dressed-particle superposition principle (Rostoker, 1964). Specifically, (238) shows that the contributions of individual particles to 𝖂\smash{{\boldsymbol{\mathfrak{W}}}} are additive and affected by the plasma collective response, i.e. by the difference between 𝚵\smash{{\boldsymbol{\Xi}}} and the vacuum dispersion matrix 𝚵0\smash{{\boldsymbol{\Xi}}_{0}}.

Using (238), one can also find other averages quadratic in the field via (cf. (53a))

(𝗟^ ~Ψ ~  )(𝗥^ ~Ψ ~  )¯dωd𝒌(𝗟𝖂𝗥)(ω,𝒌),\displaystyle\overline{(\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}})(\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}})^{\dagger}}\approx\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,({\boldsymbol{\mathsf{L}}}{\boldsymbol{\mathfrak{W}}}{\boldsymbol{\mathsf{R}}}^{\dagger})(\omega,{\boldsymbol{k}}), (239)

where 𝗟^\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}} and 𝗥^\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}} are any linear operators and 𝗟\smash{{\boldsymbol{\mathsf{L}}}} and 𝗥\smash{{\boldsymbol{\mathsf{R}}}} are their symbols; for example,

~Ψ ~  (t,𝒙) ~Ψ ~  (t,𝒙)¯dωd𝒌𝖂(ω,𝒌).\displaystyle\overline{\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}(t,{\boldsymbol{x}})\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}^{\dagger}(t,{\boldsymbol{x}})}\approx\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,{\boldsymbol{\mathfrak{W}}}(\omega,{\boldsymbol{k}}). (240)

Because of this, we loosely attribute 𝖂\smash{{\boldsymbol{\mathfrak{W}}}} as the spectrum of microscopic oscillations, but see also section 8.2, where an alternative notation is introduced and a fluctuation–dissipation theorem is derived from (238) for plasma in thermal equilibrium. See also section 9 for specific examples.

6.5 Nonlinear potentials

From (221), the oscillating part of the Hamiltonian (216) can be split into the macroscopic part and the microscopic part as H~s=H¯~s+ ~H ~  s\smash{\widetilde{H}_{s}=\underline{\widetilde{H}}_{s}+\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{H}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{H}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{H}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{H}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}_{s}}, H¯~s=𝜶^s𝚿¯~\smash{\underline{\widetilde{H}}_{s}=\widehat{\boldsymbol{\alpha}}_{s}^{\dagger}\underline{\widetilde{{\boldsymbol{\Psi}}}}}, and

~H ~  s(𝒑)=sd𝒑𝒳^ss(𝒑,𝒑)gs(𝒑).\displaystyle\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{H}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{H}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{H}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{H}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}_{s}({\boldsymbol{p}})=\sum_{s^{\prime}}\int\mathrm{d}{\boldsymbol{p}}^{\prime}\,\widehat{\mathcal{X}}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})g_{s^{\prime}}({\boldsymbol{p}}^{\prime}). (241)

Here, 𝒳^ss\smash{\widehat{\mathcal{X}}_{ss^{\prime}}} is an operator on 𝗑\smash{\mathscr{H}_{\mathsf{x}}} given by

𝒳^ss(𝒑,𝒑)𝜶^s(𝒑)𝚵^1𝜶^s(𝒑),\displaystyle\widehat{\mathcal{X}}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\doteq\widehat{\boldsymbol{\alpha}}_{s}^{\dagger}({\boldsymbol{p}})\smash{\widehat{\boldsymbol{\Xi}}}^{-1}\widehat{\boldsymbol{\alpha}}_{s^{\prime}}({\boldsymbol{p}}^{\prime}), (242)

with the symbol

𝒳ss(ω,𝒌;𝒑,𝒑)𝜶s(ω,𝒌;𝒑)𝚵1(ω,𝒌)𝜶s(ω,𝒌;𝒑)\displaystyle\mathcal{X}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\approx{\boldsymbol{\alpha}}_{s}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}){\boldsymbol{\Xi}}^{-1}(\omega,{\boldsymbol{k}}){\boldsymbol{\alpha}}_{s^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime}) (243)

(see section 9 for examples). The corresponding average Wigner functions on (𝘅,𝗸)\smash{({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})} are 𝖶¯s=𝖶¯s(m)+𝖶¯s(μ)\smash{\overline{\mathsf{W}}_{s}=\overline{\mathsf{W}}_{s}^{\text{(m)}}+\overline{\mathsf{W}}_{s}^{(\mu)}}, where the index ‘m’ stands for ‘macroscopic’ and the index ‘μ\smash{\mu}’ stands for ‘microscopic’. Because the dependence on t\smash{t} and 𝒙\smash{{\boldsymbol{x}}} is slow, one can approximate them as follows:

𝖶¯s(m)𝜶s(ω,𝒌;𝒑)𝗨(ω,𝒌)𝜶s(ω,𝒌;𝒑),\displaystyle\overline{\mathsf{W}}_{s}^{\text{(m)}}\approx{\boldsymbol{\alpha}}_{s}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}){\boldsymbol{\mathsf{U}}}(\omega,{\boldsymbol{k}}){\boldsymbol{\alpha}}_{s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}), (244a)
𝖶¯s(μ)𝜶s(ω,𝒌;𝒑)𝖂(ω,𝒌)𝜶s(ω,𝒌;𝒑).\displaystyle\overline{\mathsf{W}}_{s}^{(\mu)}\approx{\boldsymbol{\alpha}}_{s}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}){\boldsymbol{\mathfrak{W}}}(\omega,{\boldsymbol{k}}){\boldsymbol{\alpha}}_{s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}). (244b)

The matrix 𝗨\smash{{\boldsymbol{\mathsf{U}}}} is positive-semidefinite as an average Wigner tensor (section 2.1.7), and so is 𝖂\smash{{\boldsymbol{\mathfrak{W}}}} (section 6.4). Hence, both 𝖶¯s(m)\smash{\overline{\mathsf{W}}_{s}^{\text{(m)}}} and 𝖶¯s(μ)\smash{\overline{\mathsf{W}}_{s}^{(\mu)}} are non-negative. Using (238), one can also rewrite the Wigner function of ~H ~  s\smash{\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{H}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{H}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{H}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{H}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}_{s}} more compactly as

𝖶¯s(μ)(ω,𝒌;𝒑)=(2\upi)nsd𝒑δ(ω𝒌𝒗s)|𝒳ss(ω,𝒌;𝒑,𝒑)|2Fs(𝒑).\displaystyle\overline{\mathsf{W}}_{s}^{(\mu)}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})=(2\upi)^{-n}\sum_{s^{\prime}}\int\mathrm{d}{\boldsymbol{p}}^{\prime}\,\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})|\mathcal{X}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}\,F_{s^{\prime}}({\boldsymbol{p}}^{\prime}). (245)

Now we can represent the nonlinear potentials (201)–(203) as

𝗗s=𝗗s(m)+𝗗s(μ),𝚯s=𝚯s(m)+𝚯s(μ),Φs\displaystyle{\boldsymbol{\mathsf{D}}}_{s}={\boldsymbol{\mathsf{D}}}_{s}^{\text{(m)}}+{\boldsymbol{\mathsf{D}}}_{s}^{(\mu)},\qquad{\boldsymbol{\Uptheta}}_{s}={\boldsymbol{\Uptheta}}_{s}^{\text{(m)}}+{\boldsymbol{\Uptheta}}_{s}^{(\mu)},\qquad\Phi_{s} =Φs(m)+Φs(μ).\displaystyle=\Phi_{s}^{\text{(m)}}+\Phi_{s}^{(\mu)}. (246)

Here, the index (m)\smash{{}^{\text{(m)}}} denotes contributions from 𝖶¯s(m)\smash{\overline{\mathsf{W}}_{s}^{\text{(m)}}} and the index (μ) denotes contributions from 𝖶¯s(μ)\smash{\overline{\mathsf{W}}_{s}^{(\mu)}}. Specifically,

𝗗s(m)\displaystyle{\boldsymbol{\mathsf{D}}}_{s}^{\text{(m)}} =d𝒌\upi𝒌𝒌𝖶¯s(m)(𝒌𝒗s,𝒌;𝒑),\displaystyle=\int\mathrm{d}{\boldsymbol{k}}\,\upi\,{\boldsymbol{k}}{\boldsymbol{k}}\overline{\mathsf{W}}_{s}^{\text{(m)}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}}), (247)
𝚯s(m)\displaystyle{\boldsymbol{\Uptheta}}_{s}^{\text{(m)}} =ϑdωd𝒌𝒌𝒌𝖶¯s(m)(ω,𝒌;𝒑)ω𝒌𝒗s+ϑ|ϑ=0,\displaystyle=\frac{\partial}{\partial\vartheta}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\left.\frac{{\boldsymbol{k}}{\boldsymbol{k}}\overline{\mathsf{W}}_{s}^{\text{(m)}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\right|_{\vartheta=0}, (248)
Φs(m)\displaystyle\Phi_{s}^{\text{(m)}} =12𝒑dωd𝒌𝒌𝖶¯s(m)(ω,𝒌;𝒑)ω𝒌𝒗s.\displaystyle=\frac{1}{2}\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\frac{{\boldsymbol{k}}\overline{\mathsf{W}}_{s}^{\text{(m)}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}. (249)

Here, 𝖶¯s(m)\smash{\overline{\mathsf{W}}_{s}^{\text{(m)}}} is a non-negative function (244a), so 𝗗s(m)\smash{{\boldsymbol{\mathsf{D}}}_{s}^{\text{(m)}}} is positive-semidefinite and leads to an H\smash{H}-theorem similar to (206). One also has

𝗗s(μ)\displaystyle{\boldsymbol{\mathsf{D}}}_{s}^{(\mu)} =sd𝒌(2\upi)nd𝒑\upi𝒌𝒌δ(𝒌𝒗s𝒌𝒗s)|𝒳ss(𝒌𝒗s,𝒌;𝒑,𝒑)|2Fs(𝒑),\displaystyle=\sum_{s^{\prime}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,\upi\,{\boldsymbol{k}}{\boldsymbol{k}}\,\delta({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,|\mathcal{X}_{ss^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}\,F_{s^{\prime}}({\boldsymbol{p}}^{\prime}), (250)
𝚯s(μ)\displaystyle{\boldsymbol{\Uptheta}}_{s}^{(\mu)} =sϑd𝒌(2\upi)nd𝒑𝒌𝒌Fs(𝒑)𝒌(𝒗s𝒗s)+ϑ|𝒳ss(𝒌𝒗s,𝒌;𝒑,𝒑)|2|ϑ=0,\displaystyle=\sum_{s^{\prime}}\frac{\partial}{\partial\vartheta}\fint\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,\left.\frac{{\boldsymbol{k}}{\boldsymbol{k}}F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{{\boldsymbol{k}}\cdot({\boldsymbol{v}}^{\prime}_{s^{\prime}}-{\boldsymbol{v}}_{s})+\vartheta}\,|\mathcal{X}_{ss^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}},{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}\right|_{\vartheta=0}, (251)
Φs(μ)\displaystyle\Phi_{s}^{(\mu)} =s𝒑d𝒌(2\upi)nd𝒑𝒌Fs(𝒑)2𝒌(𝒗s𝒗s)|𝒳ss(𝒌𝒗s,𝒌;𝒑,𝒑)|2.\displaystyle=\sum_{s^{\prime}}\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\fint\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,\frac{{\boldsymbol{k}}F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{2{\boldsymbol{k}}\cdot({\boldsymbol{v}}^{\prime}_{s^{\prime}}-{\boldsymbol{v}}_{s})}\,|\mathcal{X}_{ss^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}},{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}. (252)

The functions 𝚯s(μ)\smash{{\boldsymbol{\Uptheta}}_{s}^{(\mu)}} and Φs(μ)\smash{\Phi_{s}^{(\mu)}} scale as 𝖶¯s(μ)\smash{\overline{\mathsf{W}}_{s}^{(\mu)}}, i.e. as ϵε2\smash{\epsilon\varepsilon^{2}} (section 3.3). Their contribution to (197) is of order ϵ𝚯s(μ)\smash{\epsilon{\boldsymbol{\Uptheta}}_{s}^{(\mu)}} and ϵΦs(μ)\smash{\epsilon\Phi_{s}^{(\mu)}}, respectively, so it scales as ϵ2ε2\smash{\epsilon^{2}\varepsilon^{2}} and therefore is negligible within our model. In contrast, 𝗗s(μ)\smash{{\boldsymbol{\mathsf{D}}}_{s}^{(\mu)}} must be retained alongside with 𝗗s(m)\smash{{\boldsymbol{\mathsf{D}}}_{s}^{\text{(m)}}}. This is because although weak, macroscopic fluctuations can resonate with particles from the bulk distribution, while the stronger macroscopic fluctuations are assumed to resonate only with particles from the tail distribution, which are few.

6.6 Oscillation-center Hamiltonian

Within the assumed accuracy, the OC Hamiltonian is s=H¯s+Φs(m)\smash{\mathcal{H}_{s}=\overline{H}_{s}+\Phi_{s}^{\text{(m)}}}, and H¯s\smash{\overline{H}_{s}} is given by (216). Combined with the general theorem (53c), the latter readily yields H¯s=H0s+ϕs\smash{\overline{H}_{s}=H_{0s}+\phi_{s}}, where

ϕs12dωd𝒌tr(𝗨s)\displaystyle\phi_{s}\approx\frac{1}{2}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\operatorname{tr}\big{(}{\boldsymbol{\mathsf{U}}}{\boldsymbol{\wp}}_{s}\big{)} (253)

and the contribution of 𝖂\smash{{\boldsymbol{\mathfrak{W}}}} has been neglected. Because both Φs(m)\smash{\Phi_{s}^{\text{(m)}}} and ϕs\smash{\phi_{s}} are quadratic in 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}} and enter s\smash{\mathcal{H}_{s}} only in the combination ΔsΦs(m)+ϕs\smash{\Delta_{s}\doteq\Phi_{s}^{\text{(m)}}+\phi_{s}}, it is convenient to attribute the latter as the ‘total’ ponderomotive energy. Using (249) in combination with (244a), one can express it as follows:

Δs=12𝒑dωd𝒌𝒌𝜶s𝗨𝜶sω𝒌𝒗s+12dωd𝒌tr(𝗨s)\displaystyle\Delta_{s}=\frac{1}{2}\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,{\boldsymbol{k}}\,\frac{{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\mathsf{U}}}{\boldsymbol{\alpha}}_{s}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}+\frac{1}{2}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\operatorname{tr}\big{(}{\boldsymbol{\mathsf{U}}}{\boldsymbol{\wp}}_{s}\big{)} (254)

(see section 9 for examples). Notably,

Δs=12δδFsdωd𝒌tr(𝚵H𝗨)=δSadδFs,\displaystyle\Delta_{s}=-\frac{1}{2}\frac{\delta}{\delta F_{s}}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\operatorname{tr}({\boldsymbol{\Xi}}_{\text{H}}{\boldsymbol{\mathsf{U}}})=-\frac{\delta S_{\text{ad}}}{\delta F_{s}}, (255)

where δ/δFs\smash{\delta/\delta F_{s}} denotes a functional derivative and Sad\smash{S_{\text{ad}}} is the adiabatic action defined in (223). Equation (255) is a generalization of the well-known ‘K\smash{K}χ\smash{\chi} theorem’ (Kaufman & Holm, 1984; Kaufman, 1987). Loosely speaking, it says that the coefficient connecting Δs\smash{\Delta_{s}} with 𝗨\smash{{\boldsymbol{\mathsf{U}}}} is proportional to the linear polarizability of an individual particle of type s\smash{s} (Dodin et al., 2017; Dodin & Fisch, 2010a). (‘K\smash{K}’ in the name of this theorem is the same as our Δs\smash{\Delta_{s}}, and ‘χ\smash{\chi}’ is the linear susceptibility.) Also, the OC Hamiltonian and the OC velocity can be expressed as

s=H0s+Δs,𝒗=𝒑H0s+𝒑Δs.\displaystyle\mathcal{H}_{s}=H_{0s}+\Delta_{s},\qquad{\boldsymbol{v}}=\partial_{\boldsymbol{p}}H_{0s}+\partial_{\boldsymbol{p}}\Delta_{s}. (256)

6.7 Polarization drag

Within the assumed accuracy, the OC distribution can be expressed as

Fs=f¯s+12𝒑(𝚯s(m)f¯s𝒑),\displaystyle F_{s}=\overline{f}_{s}+\frac{1}{2}\,\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left({\boldsymbol{\Uptheta}}_{s}^{\text{(m)}}\,\frac{\partial\overline{f}_{s}}{\partial{\boldsymbol{p}}}\right), (257)

and (197) becomes

Fsts𝒙Fs𝒑+s𝒑Fs𝒙=𝒑(𝗗s(m)Fs𝒑)+𝒞s,\displaystyle\displaystyle\frac{\partial F_{s}}{\partial t}-\frac{\partial\mathcal{H}_{s}}{\partial{\boldsymbol{x}}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}+\frac{\partial\mathcal{H}_{s}}{\partial{\boldsymbol{p}}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{x}}}=\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left({\boldsymbol{\mathsf{D}}}_{s}^{\text{(m)}}\,\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}\right)+\mathcal{C}_{s}, (258)
𝒞s𝒑(𝗗s(μ)Fs𝒑)+Γs,\displaystyle\displaystyle\mathcal{C}_{s}\doteq\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left({\boldsymbol{\mathsf{D}}}_{s}^{(\mu)}\,\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}\right)+\Gamma_{s}, (259)

where we have reinstated the term Γs\smash{\Gamma_{s}} introduced in section 3.3. As a collisional term, Γs\smash{\Gamma_{s}} is needed only to the zeroth order in ϵ\smash{\epsilon}, so

Γs={H~s,gs}¯𝒑(gs𝒙H~s¯)𝒑𝜻s,𝜻s=i𝘅|𝒌^H~s𝘅|gs¯.\displaystyle\displaystyle\Gamma_{s}=\overline{\{\widetilde{H}_{s},g_{s}\}}\approx\partial_{{\boldsymbol{p}}}\cdot(\overline{g_{s}\partial_{{\boldsymbol{x}}}\widetilde{H}_{s}})\equiv\partial_{{\boldsymbol{p}}}\cdot{\boldsymbol{\zeta}}_{s},\qquad{\boldsymbol{\zeta}}_{s}=\mathrm{i}\overline{\braket{{\boldsymbol{\mathsf{x}}}}{\widehat{\boldsymbol{k}}\widetilde{H}_{s}}\braket{{\boldsymbol{\mathsf{x}}}}{g_{s}}}. (260)

Correlating with gs\smash{g_{s}} is only the microscopic part of H~s\smash{\widetilde{H}_{s}}, so using (241) one obtains

𝜻s=isd𝒑𝘅|𝒌^𝒳^ss(𝒑,𝒑)|gs(𝒑)gs(𝒑)|¯𝘅.\displaystyle{\boldsymbol{\zeta}}_{s}=\mathrm{i}\sum_{s^{\prime}}\int\mathrm{d}{\boldsymbol{p}}^{\prime}\,\langle{\boldsymbol{\mathsf{x}}}|\widehat{\boldsymbol{k}}\widehat{\mathcal{X}}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\overline{\ket{g_{s^{\prime}}({\boldsymbol{p}}^{\prime})}\bra{g_{s}({\boldsymbol{p}})}}{\boldsymbol{\mathsf{x}}}\rangle. (261)

Next, let us use (28) and 𝜻=re𝜻\smash{{\boldsymbol{\zeta}}=\operatorname{re}{\boldsymbol{\zeta}}} to express this result as follows:

𝜻s\displaystyle{\boldsymbol{\zeta}}_{s} =isdωd𝒌d𝒑𝒌𝒳ss(ω,𝒌,𝒑,𝒑)𝕲ss(ω,𝒌,𝒑,𝒑)\displaystyle=\mathrm{i}\sum_{s^{\prime}}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,{\boldsymbol{k}}\star\mathcal{X}_{ss^{\prime}}(\omega,{\boldsymbol{k}},{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\star{\boldsymbol{\mathfrak{G}}}_{s^{\prime}s}(\omega,{\boldsymbol{k}},{\boldsymbol{p}}^{\prime},{\boldsymbol{p}})
idωd𝒌d𝒑𝒌𝒳ss(ω,𝒌,𝒑,𝒑)(2\upi)nδ(𝒑𝒑)δ(ω𝒌𝒗s)Fs(𝒑)\displaystyle\approx\mathrm{i}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,{\boldsymbol{k}}\mathcal{X}_{ss^{\prime}}(\omega,{\boldsymbol{k}},{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\,(2\upi)^{-n}\delta({\boldsymbol{p}}-{\boldsymbol{p}}^{\prime})\,\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})\,F_{s}({\boldsymbol{p}})
=d𝒌(2\upi)n𝒌im𝒳ss(𝒌𝒗s,𝒌,𝒑,𝒑)Fs(𝒑),\displaystyle=-\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,{\boldsymbol{k}}\operatorname{im}\mathcal{X}_{ss^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}},{\boldsymbol{p}},{\boldsymbol{p}})\,F_{s}({\boldsymbol{p}}), (262)

where we have approximated \smash{\star} with the usual product and substituted (237). Hence,

Γs𝒑(𝕱sFs),\displaystyle\Gamma_{s}\approx-\partial_{\boldsymbol{p}}\cdot({\boldsymbol{\mathfrak{F}}}_{s}F_{s}), (263)

where 𝕱s\smash{{\boldsymbol{\mathfrak{F}}}_{s}} can be interpreted as the polarization drag (i.e. the average force that is imposed on an OC by its dress) and is given by

𝕱s=d𝒌(2\upi)n𝒌im𝒳ss(𝒌𝒗s,𝒌,𝒑,𝒑).\displaystyle{\boldsymbol{\mathfrak{F}}}_{s}=\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,{\boldsymbol{k}}\operatorname{im}\mathcal{X}_{ss^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}},{\boldsymbol{p}},{\boldsymbol{p}}). (264)

Using (243), one also rewrite this as follows:

𝕱s\displaystyle{\boldsymbol{\mathfrak{F}}}_{s} d𝒌(2\upi)n𝒌(𝜶s(𝚵1)A𝜶s)(𝒌𝒗s,𝒌;𝒑)\displaystyle\approx\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,{\boldsymbol{k}}\,({\boldsymbol{\alpha}}_{s}^{\dagger}({\boldsymbol{\Xi}}^{-1})_{\text{A}}{\boldsymbol{\alpha}}_{s})({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}}) (265a)
d𝒌(2\upi)n𝒌(𝜶s𝚵1𝚵A𝚵𝜶s)(𝒌𝒗s,𝒌;𝒑),\displaystyle\approx-\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,{\boldsymbol{k}}\,({\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\Xi}}^{-1}{\boldsymbol{\Xi}}_{\text{A}}{\boldsymbol{\Xi}}^{-{\dagger}}{\boldsymbol{\alpha}}_{s})({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}}), (265b)

where we have substituted (233) for (𝚵1)A\smash{({\boldsymbol{\Xi}}^{-1})_{\text{A}}}. With (231) for 𝚵A\smash{{\boldsymbol{\Xi}}_{\text{A}}}, this yields

𝕱ssd𝒌(2\upi)nd𝒑\displaystyle{\boldsymbol{\mathfrak{F}}}_{s}\approx\sum_{s^{\prime}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\, \upiδ(𝒌𝒗s𝒌𝒗s)𝒌𝒌Fs(𝒑)𝒑\displaystyle\upi\,\delta({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,{\boldsymbol{k}}{\boldsymbol{k}}\cdot\frac{\partial F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial{\boldsymbol{p}}^{\prime}}
×𝜶s(𝒌𝒗s,𝒌;𝒑)𝚵1(𝒌𝒗s,𝒌)𝜶s(𝒌𝒗s,𝒌;𝒑)\displaystyle\times{\boldsymbol{\alpha}}_{s}^{\dagger}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}}){\boldsymbol{\Xi}}^{-1}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}}){\boldsymbol{\alpha}}_{s^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}}^{\prime})
×𝜶s(𝒌𝒗s,𝒌;𝒑)𝚵(𝒌𝒗s,𝒌)𝜶s(𝒌𝒗s,𝒌;𝒑).\displaystyle\times{\boldsymbol{\alpha}}_{s^{\prime}}^{\dagger}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}}^{\prime}){\boldsymbol{\Xi}}^{-{\dagger}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}}){\boldsymbol{\alpha}}_{s}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}}). (266)

The product of the last two lines equals |𝒳ss(𝒌𝒗s,𝒌;𝒑,𝒑)|2\smash{|\mathcal{X}_{ss^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}}. Hence,

𝕱s=sd𝒌(2\upi)nd𝒑\upiδ(𝒌𝒗s𝒌𝒗s)|𝒳ss(𝒌𝒗s,𝒌;𝒑,𝒑)|2𝒌𝒌Fs(𝒑)𝒑.\displaystyle{\boldsymbol{\mathfrak{F}}}_{s}=\sum_{s^{\prime}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,\upi\,\delta({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,|\mathcal{X}_{ss^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}\,{\boldsymbol{k}}{\boldsymbol{k}}\cdot\frac{\partial F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial{\boldsymbol{p}}^{\prime}}. (267)

6.8 Collision operator

By combining (263) for Γs\smash{\Gamma_{s}} with (250) for 𝗗s(μ)\smash{{\boldsymbol{\mathsf{D}}}_{s}^{(\mu)}}, one can express 𝒞s\smash{\mathcal{C}_{s}} as

𝒞s=𝒑sd𝒌(2\upi)nd𝒑\displaystyle\mathcal{C}_{s}=\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\sum_{s^{\prime}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\, \upiδ(𝒌𝒗s𝒌𝒗s)|𝒳ss(𝒌𝒗s,𝒌;𝒑,𝒑)|2\displaystyle\upi\,\delta({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,|\mathcal{X}_{ss^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}
×𝒌𝒌(Fs(𝒑)𝒑Fs(𝒑)Fs(𝒑)Fs(𝒑)𝒑),\displaystyle\times{\boldsymbol{k}}{\boldsymbol{k}}\cdot\left(\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}\,F_{s^{\prime}}({\boldsymbol{p}}^{\prime})-F_{s}({\boldsymbol{p}})\,\frac{\partial F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial{\boldsymbol{p}}^{\prime}}\right), (268)

where 𝒳ss\mathcal{X}_{ss^{\prime}} is given by (243). One can recognize this as a generalization of the Balescu–Lenard collision operator (Krall & Trivelpiece, 1973, section 11.11) to interactions via a general multi-component field 𝚿{\boldsymbol{\Psi}}. Specific examples can be found in section 9.

It is readily seen that 𝒞s\smash{\mathcal{C}_{s}} conserves particles, i.e.

d𝒑𝒞s=0,\displaystyle\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{C}_{s}=0, (269)

and vanishes in thermal equilibrium (section 8.1). Other properties of 𝒞s\smash{\mathcal{C}_{s}} are determined by the properties of the coupling coefficient 𝒳ss\smash{\mathcal{X}_{ss^{\prime}}}, which are as follows. Note that

|𝒳ss(ω,𝒌;𝒑,𝒑)|2=𝒬ss(ω,𝒌;𝒑,𝒑)+ss(ω,𝒌;𝒑,𝒑)/2,\displaystyle|\mathcal{X}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}=\mathcal{Q}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})+\mathcal{R}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})/2, (270)

where we introduced

𝒬ss(ω,𝒌;𝒑,𝒑)\displaystyle\mathcal{Q}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime}) (|𝒳ss(ω,𝒌;𝒑,𝒑)|2+|𝒳ss(ω,𝒌;𝒑,𝒑)|2)/2,\displaystyle\doteq(|\mathcal{X}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}+|\mathcal{X}_{s^{\prime}s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime},{\boldsymbol{p}})|^{2})/2, (271)
ss(ω,𝒌;𝒑,𝒑)\displaystyle\mathcal{R}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime}) |𝒳ss(ω,𝒌;𝒑,𝒑)|2|𝒳ss(ω,𝒌;𝒑,𝒑)|2.\displaystyle\doteq|\mathcal{X}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}-|\mathcal{X}_{s^{\prime}s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime},{\boldsymbol{p}})|^{2}. (272)

To calculate ss\smash{\mathcal{R}_{ss^{\prime}}}, note that (243) yields

|𝒳ss(ω,𝒌;𝒑,𝒑)|2\displaystyle|\mathcal{X}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2} |𝜶s(ω,𝒌;𝒑)𝚵1(ω,𝒌)𝜶s(ω,𝒌;𝒑)|2,\displaystyle\approx|{\boldsymbol{\alpha}}_{s}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}){\boldsymbol{\Xi}}^{-1}(\omega,{\boldsymbol{k}}){\boldsymbol{\alpha}}_{s^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime})|^{2},
|𝒳ss(ω,𝒌;𝒑,𝒑)|2\displaystyle|\mathcal{X}_{s^{\prime}s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime},{\boldsymbol{p}})|^{2} |𝜶s(ω,𝒌;𝒑)𝚵(ω,𝒌)𝜶s(ω,𝒌;𝒑)|2,\displaystyle\approx|{\boldsymbol{\alpha}}_{s}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}){\boldsymbol{\Xi}}^{-{\dagger}}(\omega,{\boldsymbol{k}}){\boldsymbol{\alpha}}_{s^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime})|^{2}, (273)

whence one obtains

ss(ω,𝒌;𝒑,𝒑)4im(\displaystyle\mathcal{R}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\approx 4\operatorname{im}\big{(} 𝜶s(ω,𝒌;𝒑)(𝚵1)H(ω,𝒌)𝜶s(ω,𝒌;𝒑)\displaystyle{\boldsymbol{\alpha}}_{s}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})({\boldsymbol{\Xi}}^{-1})_{\text{H}}(\omega,{\boldsymbol{k}}){\boldsymbol{\alpha}}_{s^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime})
𝜶s(ω,𝒌;𝒑)(𝚵1)A(ω,𝒌)𝜶s(ω,𝒌;𝒑)).\displaystyle{\boldsymbol{\alpha}}_{s^{\prime}}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime})({\boldsymbol{\Xi}}^{-1})_{\text{A}}(\omega,{\boldsymbol{k}}){\boldsymbol{\alpha}}_{s^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime})\big{)}. (274)

The operators (𝚵1)H\smash{({\boldsymbol{\Xi}}^{-1})_{\text{H}}}, (𝚵1)A\smash{({\boldsymbol{\Xi}}^{-1})_{\text{A}}}, and 𝜶^s\smash{\widehat{\boldsymbol{\alpha}}_{s}} (for all s\smash{s}) have been introduced for real fields, so their matrix elements in the coordinate representation are real. Then, the corresponding symbols satisfy 𝑨(ω,𝒌)=𝑨(ω,𝒌)\smash{{\boldsymbol{A}}(-\omega,-{\boldsymbol{k}})={\boldsymbol{A}}^{*}(\omega,{\boldsymbol{k}})}, where 𝑨\smash{{\boldsymbol{A}}} is any of the three symbols. This gives

ss(𝒌;𝒑,𝒑)ss(𝒌𝒗s,𝒌;𝒑,𝒑)=ss(𝒌;𝒑,𝒑).\displaystyle\mathcal{R}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\doteq\mathcal{R}_{ss^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})=-\mathcal{R}_{ss^{\prime}}(-{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime}). (275)

Because the rest of the integrand in (6.8) is even in 𝒌\smash{{\boldsymbol{k}}}, (275) signifies that ss\smash{\mathcal{R}_{ss^{\prime}}} does not contribute to 𝒞s\smash{\mathcal{C}_{s}}. Thus, 𝒳ss\smash{\mathcal{X}_{ss^{\prime}}} in (6.8) can as well be replaced with 𝒬ss\smash{\mathcal{Q}_{ss^{\prime}}}:

𝒞s=𝒑sd𝒌(2\upi)nd𝒑\displaystyle\mathcal{C}_{s}=\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\sum_{s^{\prime}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\, \upiδ(𝒌𝒗s𝒌𝒗s)𝒬ss(𝒌𝒗s,𝒌;𝒑,𝒑)\displaystyle\upi\,\delta({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,\mathcal{Q}_{ss^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})
×𝒌𝒌(Fs(𝒑)𝒑Fs(𝒑)Fs(𝒑)Fs(𝒑)𝒑).\displaystyle\times{\boldsymbol{k}}{\boldsymbol{k}}\cdot\left(\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}\,F_{s^{\prime}}({\boldsymbol{p}}^{\prime})-F_{s}({\boldsymbol{p}})\,\frac{\partial F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial{\boldsymbol{p}}^{\prime}}\right). (276)

In this representation, the coupling coefficient in 𝒞s\smash{\mathcal{C}_{s}} is manifestly symmetric,

𝒬ss(ω,𝒌;𝒑,𝒑)=𝒬ss(ω,𝒌;𝒑,𝒑),\displaystyle\mathcal{Q}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})=\mathcal{Q}_{s^{\prime}s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime},{\boldsymbol{p}}), (277)

which readily leads to momentum and energy conservation (appendix C):282828Remember that 𝒗s\smash{{\boldsymbol{v}}_{s}} is defined as the OC velocity in the above formulas (section 5.6). If 𝒗s\smash{{\boldsymbol{v}}_{s}} is treated as the particle velocity instead, then s\smash{\mathcal{H}_{s}} in (278) should be replaced with H¯s\smash{\overline{H}_{s}}. Both options are admissible within the assumed accuracy, but the former option is preferable because it leads to other conservation laws that are exact within our model (section 7.5).

sd𝒑𝒑𝒞s=0,sd𝒑s𝒞s=0.\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{p}}\mathcal{C}_{s}=0,\qquad\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\mathcal{C}_{s}=0. (278)

The collision operator 𝒞s\smash{\mathcal{C}_{s}} also satisfies the H\smash{H}-theorem (appendix C.3):

(dσdt)coll0,\displaystyle\left(\frac{\mathrm{d}\sigma}{\mathrm{d}t}\right)_{\text{coll}}\geq 0, (279)

where the entropy density σ\smash{\sigma} is defined as

σsd𝒑Fs(𝒑)lnFs(𝒑)\displaystyle\sigma\doteq-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,F_{s}({\boldsymbol{p}})\ln F_{s}({\boldsymbol{p}}) (280)

and (tFs)coll𝒞s\smash{(\partial_{t}F_{s})_{\text{coll}}\doteq\mathcal{C}_{s}}. Note that these properties are not restricted to any particular s\smash{\mathcal{H}_{s}}. Also note that if applied in proper variables (section 3.1.2), our formula (6.8) can describe collisions in strong background fields. This topic, including comparison with the relevant literature, is left to future work.

6.9 Summary of section 6

Let us summarize the above general results (for examples, see section 9). We consider species s\smash{s} governed by a Hamiltonian of the form

Hs=H0s+𝜶^s(𝒑)𝚿~+12(𝑳^s𝚿~)(𝑹^s𝚿~),\displaystyle H_{s}=H_{0s}+\widehat{\boldsymbol{\alpha}}_{s}^{\dagger}({\boldsymbol{p}})\widetilde{{\boldsymbol{\Psi}}}+\frac{1}{2}\,(\widehat{\boldsymbol{L}}_{s}\widetilde{{\boldsymbol{\Psi}}})^{\dagger}(\widehat{\boldsymbol{R}}_{s}\widetilde{{\boldsymbol{\Psi}}}), (281)

where 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}} is a real oscillating field (of any dimension M\smash{M}), which generally consists of a macroscopic part 𝚿¯~\smash{\underline{\widetilde{{\boldsymbol{\Psi}}}}} and a microscopic part 𝚿~\textstyle\widetilde{{\boldsymbol{\Psi}}}  ~\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt} . The term H0s\smash{H_{0s}} is independent of 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}}, and the operators 𝜶^s\smash{\smash{\widehat{\boldsymbol{\alpha}}}_{s}^{\dagger}}, 𝑳^s\smash{\widehat{\boldsymbol{L}}_{s}}, and 𝑹^s\smash{\widehat{\boldsymbol{R}}_{s}} may be nonlocal in tt and 𝒙{\boldsymbol{x}} and may depend on the momentum 𝒑{\boldsymbol{p}} parametrically. The dynamics of this system averaged over the fast oscillations can be described in terms of the OC distribution function

Fs=f¯s+12𝒑(𝚯sf¯s𝒑)\displaystyle F_{s}=\overline{f}_{s}+\frac{1}{2}\,\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left({\boldsymbol{\Uptheta}}_{s}\,\frac{\partial\overline{f}_{s}}{\partial{\boldsymbol{p}}}\right) (282)

(the index (m)\smash{{}^{\text{(m)}}} is henceforth omitted for brevity), which is governed by the following equation of the Fokker–Planck type:

Fsts𝒙Fs𝒑+s𝒑Fs𝒙=𝒑(𝗗sFs𝒑)+𝒞s.\displaystyle\frac{\partial F_{s}}{\partial t}-\frac{\partial\mathcal{H}_{s}}{\partial{\boldsymbol{x}}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}+\frac{\partial\mathcal{H}_{s}}{\partial{\boldsymbol{p}}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{x}}}=\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left({\boldsymbol{\mathsf{D}}}_{s}\,\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}\right)+\mathcal{C}_{s}. (283)

Here, s=H0s+Δs\smash{\mathcal{H}_{s}=H_{0s}+\Delta_{s}} is the OC Hamiltonian, 𝚯s\smash{{\boldsymbol{\Uptheta}}_{s}} is the dressing function, and Δs\smash{\Delta_{s}} is the total ponderomotive energy (i.e. the part of the OC Hamiltonian that is quadratic in 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}}), so 𝒗s(t,𝒙,𝒑)𝒑s\smash{{\boldsymbol{v}}_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}})\doteq\partial_{\boldsymbol{p}}\mathcal{H}_{s}} is the OC velocity. Specifically,

𝗗s\displaystyle{\boldsymbol{\mathsf{D}}}_{s} =d𝒌\upi𝒌𝒌𝖶¯s(t,𝒙,𝒌𝒗s,𝒌;𝒑),\displaystyle=\int\mathrm{d}{\boldsymbol{k}}\,\upi\,{\boldsymbol{k}}{\boldsymbol{k}}\overline{\mathsf{W}}_{s}(t,{\boldsymbol{x}},{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}}), (284a)
𝚯s\displaystyle{\boldsymbol{\Uptheta}}_{s} =ϑdωd𝒌𝒌𝒌𝖶¯s(t,𝒙,ω,𝒌;𝒑)ω𝒌𝒗s+ϑ|ϑ=0,\displaystyle=\frac{\partial}{\partial\vartheta}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,{\boldsymbol{k}}{\boldsymbol{k}}\left.\frac{\overline{\mathsf{W}}_{s}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}})}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\right|_{\vartheta=0}, (284b)
Δs\displaystyle\Delta_{s} =12𝒑dωd𝒌𝒌𝖶¯s(t,𝒙,ω,𝒌;𝒑)ω𝒌𝒗s\displaystyle=\frac{1}{2}\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,{\boldsymbol{k}}\,\frac{\overline{\mathsf{W}}_{s}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}})}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}
+12dωd𝒌tr(𝗨s)(t,𝒙,ω,𝒌;𝒑).\displaystyle\hphantom{\,=\,}+\frac{1}{2}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\operatorname{tr}\big{(}{\boldsymbol{\mathsf{U}}}{\boldsymbol{\wp}}_{s}\big{)}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}}). (284c)

Here, 𝖶¯s=𝜶s𝗨𝜶s\smash{\overline{\mathsf{W}}_{s}={\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\mathsf{U}}}{\boldsymbol{\alpha}}_{s}} is a scalar function, the average Wigner matrix 𝗨\smash{{\boldsymbol{\mathsf{U}}}} is understood as the Fourier spectrum of the symmetrized autocorrelation matrix of the macroscopic oscillations:

𝗨(t,𝒙,ω,𝒌)=dτ2\upid𝒔(2\upi)n𝚿¯~(t+τ/2,𝒙+𝒔/2)𝚿¯~(tτ/2,𝒙𝒔/2)¯eiωτi𝒌𝒔,{\boldsymbol{\mathsf{U}}}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}})=\int\frac{\mathrm{d}\tau}{2\upi}\,\frac{\mathrm{d}{\boldsymbol{s}}}{(2\upi)^{n}}\,\,\overline{{\underline{\widetilde{{\boldsymbol{\Psi}}}}}(t+\tau/2,{\boldsymbol{x}}+{\boldsymbol{s}}/2)\,\smash{\underline{\widetilde{{\boldsymbol{\Psi}}}}}^{\dagger}(t-\tau/2,{\boldsymbol{x}}-{\boldsymbol{s}}/2)}\,\mathrm{e}^{\mathrm{i}\omega\tau-\mathrm{i}{\boldsymbol{k}}\cdot{\boldsymbol{s}}}, (285)

with ndim𝒙\smash{n\doteq\dim{\boldsymbol{x}}}. Also, the vector 𝜶s(t,𝒙,ω,𝒌;𝒑)\smash{{\boldsymbol{\alpha}}_{s}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}})} is the Weyl symbol of 𝜶^s\smash{\widehat{\boldsymbol{\alpha}}_{s}} as defined in (26), s(t,𝒙,ω,𝒌;𝒑)(𝑳s𝑹s)H\smash{{\boldsymbol{\wp}}_{s}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}})\approx({\boldsymbol{L}}_{s}^{\dagger}{\boldsymbol{R}}_{s})_{\text{H}}}, 𝑳s\smash{{\boldsymbol{L}}_{s}} and 𝑹s\smash{{\boldsymbol{R}}_{s}} are the Weyl symbols of 𝑳^s\smash{\widehat{\boldsymbol{L}}_{s}} and 𝑹^s\smash{\widehat{\boldsymbol{R}}_{s}}, respectively, and H\smash{{}_{\text{H}}} denotes the Hermitian part. The matrix 𝗗s\smash{{\boldsymbol{\mathsf{D}}}_{s}} is positive-semidefinite and satisfies an H\smash{H}-theorem of the form (206). Also, Δs\smash{\Delta_{s}} satisfies the ‘K\smash{K}χ\smash{\chi} theorem’:

Δs=12δδFsdωd𝒌tr(𝚵H𝗨).\displaystyle\Delta_{s}=-\frac{1}{2}\frac{\delta}{\delta F_{s}}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\operatorname{tr}({\boldsymbol{\Xi}}_{\text{H}}{\boldsymbol{\mathsf{U}}}). (286)

The matrix 𝚵{\boldsymbol{\Xi}} characterizes the collective plasma response to the field 𝚿~\widetilde{{\boldsymbol{\Psi}}} and is given by

𝚵𝚵0+sd𝒑(𝜶s(𝒑)𝜶s(𝒑)ω𝒌𝒗s(𝒑)+i0𝒌Fs(𝒑)𝒑s(𝒑)Fs(𝒑)).\displaystyle{\boldsymbol{\Xi}}\approx{\boldsymbol{\Xi}}_{0}+\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\Bigg{(}\frac{{\boldsymbol{\alpha}}_{s}({\boldsymbol{p}})\,{\boldsymbol{\alpha}}_{s}^{\dagger}({\boldsymbol{p}})}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}({\boldsymbol{p}})+\mathrm{i}0}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}-{\boldsymbol{\wp}}_{s}({\boldsymbol{p}})F_{s}({\boldsymbol{p}})\Bigg{)}. (287)

Here, the arguments (t,𝒙,ω,𝒌)\smash{(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}})} are omitted for brevity, 𝜶s𝜶s\smash{{\boldsymbol{\alpha}}_{s}{\boldsymbol{\alpha}}_{s}^{\dagger}} is a dyadic matrix, and 𝚵0\smash{{\boldsymbol{\Xi}}_{0}} is the symbol of the Hermitian dispersion operator 𝚵^0\smash{\widehat{\boldsymbol{\Xi}}_{0}} that governs the field 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}} in the absence of plasma. Specifically, 𝚵^0\smash{\widehat{\boldsymbol{\Xi}}_{0}} is defined such that the field Lagrangian density without plasma is 𝔏0=𝚿~𝚵^0𝚿~/2\smash{\mathfrak{L}_{0}=\smash{\widetilde{{\boldsymbol{\Psi}}}}^{\dagger}\widehat{\boldsymbol{\Xi}}_{0}\widetilde{{\boldsymbol{\Psi}}}/2}.

The spectrum of microscopic fluctuations (specifically, the spectrum of the symmetrized autocorrelation function of the microscopic field 𝚿~\textstyle\widetilde{{\boldsymbol{\Psi}}}  ~\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt} ) is a positive-semidefinite matrix function and given by

𝖂(ω,𝒌)=1(2\upi)nsd𝒑\displaystyle{\boldsymbol{\mathfrak{W}}}(\omega,{\boldsymbol{k}})=\frac{1}{(2\upi)^{n}}\sum_{s^{\prime}}\int\mathrm{d}{\boldsymbol{p}}^{\prime}\, δ(ω𝒌𝒗s)Fs(𝒑)\displaystyle\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})F_{s^{\prime}}({\boldsymbol{p}}^{\prime})
×𝚵1(ω,𝒌)(𝜶s𝜶s)(ω,𝒌;𝒑)𝚵(ω,𝒌),\displaystyle\times{\boldsymbol{\Xi}}^{-1}(\omega,{\boldsymbol{k}})({\boldsymbol{\alpha}}_{s^{\prime}}{\boldsymbol{\alpha}}_{s^{\prime}}^{\dagger})(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime}){\boldsymbol{\Xi}}^{-{\dagger}}(\omega,{\boldsymbol{k}}), (288)

where 𝒗s𝒗s(𝒑)\smash{{\boldsymbol{v}}^{\prime}_{s^{\prime}}\doteq{\boldsymbol{v}}_{s^{\prime}}({\boldsymbol{p}}^{\prime})}. (The dependence on t\smash{t} and 𝒙\smash{{\boldsymbol{x}}} is assumed too but not emphasized.) The microscopic fluctuations give rise to a collision operator of the Balescu–Lenard type:

𝒞s=𝒑s\displaystyle\mathcal{C}_{s}=\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\sum_{s^{\prime}}\int d𝒌(2\upi)nd𝒑\upiδ(𝒌𝒗s𝒌𝒗s)𝒬ss(𝒌𝒗s,𝒌;𝒑,𝒑)\displaystyle\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,\upi\,\delta({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,\mathcal{Q}_{ss^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})
×𝒌𝒌(Fs(𝒑)𝒑Fs(𝒑)Fs(𝒑)Fs(𝒑)𝒑),\displaystyle\times{\boldsymbol{k}}{\boldsymbol{k}}\cdot\left(\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}\,F_{s^{\prime}}({\boldsymbol{p}}^{\prime})-F_{s}({\boldsymbol{p}})\,\frac{\partial F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial{\boldsymbol{p}}^{\prime}}\right), (289)

where the coupling coefficient 𝒬ss(ω,𝒌;𝒑,𝒑)=𝒬ss(ω,𝒌;𝒑,𝒑)\smash{\mathcal{Q}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})=\mathcal{Q}_{s^{\prime}s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime},{\boldsymbol{p}})} is given by

𝒬ss(ω,𝒌;𝒑,𝒑)=(|𝒳ss(ω,𝒌;𝒑,𝒑)|2+|𝒳ss(ω,𝒌;𝒑,𝒑)|2)/2,\displaystyle\displaystyle\mathcal{Q}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})=(|\mathcal{X}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}+|\mathcal{X}_{s^{\prime}s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime},{\boldsymbol{p}})|^{2})/2, (290)
𝒳ss(ω,𝒌;𝒑,𝒑)𝜶s(ω,𝒌;𝒑)𝚵1(ω,𝒌)𝜶s(ω,𝒌;𝒑).\displaystyle\mathcal{X}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\approx{\boldsymbol{\alpha}}_{s}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}){\boldsymbol{\Xi}}^{-1}(\omega,{\boldsymbol{k}}){\boldsymbol{\alpha}}_{s^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime}). (291)

The operator 𝒞s\smash{\mathcal{C}_{s}} satisfies the H\smash{H}-theorem and conserves particles, momentum, and energy:

d𝒑𝒞s=0,sd𝒑𝒑𝒞s=0,sd𝒑s𝒞s=0.\displaystyle\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{C}_{s}=0,\qquad\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{p}}\mathcal{C}_{s}=0,\qquad\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\mathcal{C}_{s}=0.

7 Interaction with on-shell waves

Here, we discuss QL interaction of plasma with ‘on-shell’ waves, i.e. waves constrained by dispersion relations. To motivate the assumptions that will be adopted, and also to systematically introduce our notation, we start with briefly overviewing theory of linear waves in dispersive media (Tracy et al., 2014; Whitham, 1974), including monochromatic waves (section 7.1), conservative eikonal waves (section 7.2), general eikonal waves (section 7.3), and general broadband waves described by the WKE (section 7.4). After that, we derive conservation laws for the total momentum and energy, which are exact within our model (section 7.5). All waves in this section are considered macroscopic, so we adopt a simplified notation 𝚿¯~𝚿~\smash{\underline{\widetilde{{\boldsymbol{\Psi}}}}\equiv\widetilde{{\boldsymbol{\Psi}}}} and the index (m)\smash{{}^{(\text{m})}} will be omitted.

7.1 Monochromatic waves

Conservative (nondissipative) waves can be described using the least-action principle δS=0\delta S=0. Assuming the notation as in section 6.2, the action integral can be expressed as S=d𝘅𝔏S=\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\mathfrak{L} with the Lagrangian density given by

𝔏=12𝚿~𝚵^H𝚿~.\displaystyle\mathfrak{L}=\frac{1}{2}\,\smash{\widetilde{{\boldsymbol{\Psi}}}}^{\dagger}\widehat{\boldsymbol{\Xi}}_{\text{H}}\widetilde{{\boldsymbol{\Psi}}}. (292)

First, let us assume a homogeneous stationary medium, so 𝚵H(t,𝒙,ω,𝒌)=𝚵H(ω,𝒌)\smash{{\boldsymbol{\Xi}}_{\text{H}}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}})={\boldsymbol{\Xi}}_{\text{H}}(\omega,{\boldsymbol{k}})}. Because we assume real fields,292929A complex field can be accommodated by considering its real and imaginary parts as separate components. t1,𝒙1|𝚵^|t2,𝒙2\smash{\braket{t_{1},{\boldsymbol{x}}_{1}}{\widehat{\boldsymbol{\Xi}}}{t_{2},{\boldsymbol{x}}_{2}}} is real for all (t1,𝒙1,t2,𝒙2)\smash{(t_{1},{\boldsymbol{x}}_{1},t_{2},{\boldsymbol{x}}_{2})}, one also has

𝚵H(ω,𝒌)=𝚵H(ω,𝒌)=𝚵H(ω,𝒌),\displaystyle{\boldsymbol{\Xi}}_{\text{H}}(-\omega,-{\boldsymbol{k}})={\boldsymbol{\Xi}}_{\text{H}}^{*}(\omega,{\boldsymbol{k}})={\boldsymbol{\Xi}}_{\text{H}}^{\intercal}(\omega,{\boldsymbol{k}}), (293)

where the latter equality is due to 𝚵H(ω,𝒌)=𝚵H(ω,𝒌)\smash{{\boldsymbol{\Xi}}_{\text{H}}^{\dagger}(\omega,{\boldsymbol{k}})={\boldsymbol{\Xi}}_{\text{H}}(\omega,{\boldsymbol{k}})}.

Because 𝚵H(ω,𝒌){\boldsymbol{\Xi}}_{\text{H}}(\omega,{\boldsymbol{k}}) is Hermitian, it has Mdim𝚵HM\doteq\dim{\boldsymbol{\Xi}}_{\text{H}} orthonormal eigenvectors 𝜼b{\boldsymbol{\eta}}_{b}:

𝚵H(ω,𝒌)𝜼b(ω,𝒌)=Λb(ω,𝒌)𝜼b(ω,𝒌),𝜼b(ω,𝒌)𝜼b(ω,𝒌)=δb,b.\displaystyle{\boldsymbol{\Xi}}_{\text{H}}(\omega,{\boldsymbol{k}}){\boldsymbol{\eta}}_{b}(\omega,{\boldsymbol{k}})=\Lambda_{b}(\omega,{\boldsymbol{k}}){\boldsymbol{\eta}}_{b}(\omega,{\boldsymbol{k}}),\qquad{\boldsymbol{\eta}}_{b}^{\dagger}(\omega,{\boldsymbol{k}}){\boldsymbol{\eta}}_{b^{\prime}}(\omega,{\boldsymbol{k}})=\delta_{b,b^{\prime}}. (294)

Here Λb\smash{\Lambda_{b}} are the corresponding eigenvalues, which are real and satisfy

Λb(ω,𝒌)=𝜼b(ω,𝒌)𝚵H(ω,𝒌)𝜼b(ω,𝒌).\displaystyle\Lambda_{b}(\omega,{\boldsymbol{k}})={\boldsymbol{\eta}}_{b}^{\dagger}(\omega,{\boldsymbol{k}}){\boldsymbol{\Xi}}_{\text{H}}(\omega,{\boldsymbol{k}}){\boldsymbol{\eta}}_{b}(\omega,{\boldsymbol{k}}). (295)

Due to (293), one has

Λb(ω,𝒌)=eigvb(𝚵H(ω,𝒌))=eigvb𝚵H(ω,𝒌)=Λb(ω,𝒌),\displaystyle\Lambda_{b}(-\omega,-{\boldsymbol{k}})=\text{eigv}_{b}\,({\boldsymbol{\Xi}}_{\text{H}}(\omega,{\boldsymbol{k}}))^{\intercal}=\text{eigv}_{b}\,{\boldsymbol{\Xi}}_{\text{H}}(\omega,{\boldsymbol{k}})=\Lambda_{b}(\omega,{\boldsymbol{k}}), (296)

where eigvb\text{eigv}_{b} stands for the bbth eigenvalue. Using this together with (293), one obtains from (294) that

𝚵H(ω,𝒌)𝜼b(ω,𝒌)=Λb(ω,𝒌)𝜼b(ω,𝒌),\displaystyle{\boldsymbol{\Xi}}_{\text{H}}^{*}(\omega,{\boldsymbol{k}}){\boldsymbol{\eta}}_{b}(-\omega,-{\boldsymbol{k}})=\Lambda_{b}(\omega,{\boldsymbol{k}}){\boldsymbol{\eta}}_{b}(-\omega,-{\boldsymbol{k}}), (297)

whence

𝜼b(ω,𝒌)=𝜼b(ω,𝒌).\displaystyle{\boldsymbol{\eta}}_{b}(-\omega,-{\boldsymbol{k}})={\boldsymbol{\eta}}_{b}^{*}(\omega,{\boldsymbol{k}}). (298)

Let us consider a monochromatic wave of the form

𝚿~(t,𝒙)=re(eiω¯t+i𝒌¯𝒙𝚿˘),\displaystyle\widetilde{{\boldsymbol{\Psi}}}(t,{\boldsymbol{x}})=\operatorname{re}(\mathrm{e}^{-\mathrm{i}\overline{\omega}t+\mathrm{i}\overline{{\boldsymbol{k}}}\cdot{\boldsymbol{x}}}\,{\breve{{\boldsymbol{\Psi}}}}), (299)

with real frequency ω¯\overline{\omega}, real wavevector 𝒌¯\overline{{\boldsymbol{k}}}, and complex amplitude 𝚿˘{\breve{{\boldsymbol{\Psi}}}}. For such a wave, the action integral can be expressed as S=d𝘅𝔏¯\smash{S=\int\mathrm{d}{\boldsymbol{\mathsf{x}}}\,\overline{\mathfrak{L}}}, where the average Lagrangian density 𝔏¯\overline{\mathfrak{L}} is given by303030Here we use that for any oscillating a=re(eiθa˘)\smash{a=\operatorname{re}(\mathrm{e}^{\mathrm{i}\theta}{\breve{a}})} and b=re(eiθb˘)\smash{b=\operatorname{re}(\mathrm{e}^{\mathrm{i}\theta}{\breve{b}})}, one has ab¯=re(a˘b˘)/2\smash{\overline{ab}=\operatorname{re}({\breve{a}}^{*}{\breve{b}})/2} and that 𝚿˘𝚵H(ω¯,𝒌¯)𝚿˘\smash{\smash{{\breve{{\boldsymbol{\Psi}}}}}^{\dagger}{\boldsymbol{\Xi}}_{\text{H}}(\overline{\omega},\overline{{\boldsymbol{k}}}){\breve{{\boldsymbol{\Psi}}}}} is real because 𝚵H(ω¯,𝒌¯)\smash{{\boldsymbol{\Xi}}_{\text{H}}(\overline{\omega},\overline{{\boldsymbol{k}}})} is Hermitian.

𝔏¯=12𝚿~𝚵^H𝚿~¯=14re(𝚿˘𝚵H(ω¯,𝒌¯)𝚿˘)=14𝚿˘𝚵H(ω¯,𝒌¯)𝚿˘.\displaystyle\overline{\mathfrak{L}}=\frac{1}{2}\,\overline{\smash{\widetilde{{\boldsymbol{\Psi}}}}^{\dagger}\widehat{\boldsymbol{\Xi}}_{\text{H}}\widetilde{{\boldsymbol{\Psi}}}}=\frac{1}{4}\,\operatorname{re}(\smash{{\breve{{\boldsymbol{\Psi}}}}}^{\dagger}{\boldsymbol{\Xi}}_{\text{H}}(\overline{\omega},\overline{{\boldsymbol{k}}}){\breve{{\boldsymbol{\Psi}}}})=\frac{1}{4}\,\smash{{\breve{{\boldsymbol{\Psi}}}}}^{\dagger}{\boldsymbol{\Xi}}_{\text{H}}(\overline{\omega},\overline{{\boldsymbol{k}}}){\breve{{\boldsymbol{\Psi}}}}. (300)

Let us decompose 𝚿˘{\breve{{\boldsymbol{\Psi}}}} in the basis formed by the eigenvectors 𝜼b\smash{{\boldsymbol{\eta}}_{b}}, that is, as

𝚿˘=b𝜼ba˘b.\displaystyle{\breve{{\boldsymbol{\Psi}}}}=\sum_{b}{\boldsymbol{\eta}}_{b}{\breve{a}}^{b}. (301)

Then, (300) becomes

𝔏¯=14bΛb(ω¯,𝒌¯)|a˘b|2.\displaystyle\overline{\mathfrak{L}}=\frac{1}{4}\sum_{b}\Lambda_{b}(\overline{\omega},\overline{{\boldsymbol{k}}})\,|{\breve{a}}^{b}|^{2}. (302)

The real and imaginary parts of the amplitudes a˘b{\breve{a}}^{b} can be treated as independent variables. This is equivalent to treating a˘b{\breve{a}}^{b*} and a˘b{\breve{a}}^{b} as independent variables, so one arrives at the following Euler–Lagrange equations:

0=δS[𝒂˘,𝒂˘]δa˘b=14Λb(ω¯,𝒌¯)a˘b,0=δS[𝒂˘,𝒂˘]δa˘b=14a˘bΛb(ω¯,𝒌¯).\displaystyle 0=\frac{\delta S[{\breve{{\boldsymbol{a}}}}^{*},{\breve{{\boldsymbol{a}}}}]}{\delta{\breve{a}}^{b*}}=\frac{1}{4}\,\Lambda_{b}(\overline{\omega},\overline{{\boldsymbol{k}}}){\breve{a}}^{b},\qquad 0=\frac{\delta S[{\breve{{\boldsymbol{a}}}}^{*},{\breve{{\boldsymbol{a}}}}]}{\delta{\breve{a}}^{b}}=\frac{1}{4}\,{\breve{a}}^{b*}\Lambda_{b}(\overline{\omega},\overline{{\boldsymbol{k}}}). (303)

Hence the bbth mode with a nonzero amplitude a˘b{\breve{a}}^{b} satisfies the dispersion relation

0=Λb(ω¯,𝒌¯)=Λb(ω¯,𝒌¯).\displaystyle 0=\Lambda_{b}(\overline{\omega},\overline{{\boldsymbol{k}}})=\Lambda_{b}(-\overline{\omega},-\overline{{\boldsymbol{k}}}). (304)

Equation (304) determines a dispersion surface in the 𝗸{\boldsymbol{\mathsf{k}}} space where the waves can have nonzero amplitude. This surface is sometimes called a shell, so waves constrained by a dispersion relation are called on-shell. Also note that combining (304) with (294) yields that on-shell waves satisfy

𝚵H(ω¯,𝒌¯)𝜼b(ω¯,𝒌¯)=𝟎,𝜼b(ω¯,𝒌¯)𝚵H(ω¯,𝒌¯)=𝟎,\displaystyle{\boldsymbol{\Xi}}_{\text{H}}(\overline{\omega},\overline{{\boldsymbol{k}}}){\boldsymbol{\eta}}_{b}(\overline{\omega},\overline{{\boldsymbol{k}}})={\boldsymbol{0}},\qquad{\boldsymbol{\eta}}_{b}^{\dagger}(\overline{\omega},\overline{{\boldsymbol{k}}}){\boldsymbol{\Xi}}_{\text{H}}(\overline{\omega},\overline{{\boldsymbol{k}}})={\boldsymbol{0}}, (305)

which are two mutually adjoint representations of the same equation.

Below, we consider the case when (304) is satisfied only for one mode at a time, so summation over bb and the index bb itself can be omitted. (A more general case is discussed, for example, in (Dodin et al., 2019).) Then, 𝚿˘=𝜼(ω¯,𝒌¯)a˘\smash{{\breve{{\boldsymbol{\Psi}}}}={\boldsymbol{\eta}}(\overline{\omega},\overline{{\boldsymbol{k}}}){\breve{a}}},

𝔏¯=14Λ(ω¯,𝒌¯)|a˘|2,\displaystyle\overline{\mathfrak{L}}=\frac{1}{4}\,\Lambda(\overline{\omega},\overline{{\boldsymbol{k}}})|{\breve{a}}|^{2}, (306)

and ω¯\overline{\omega} is connected with 𝒌¯\overline{{\boldsymbol{k}}} via ω¯=w(𝒌¯)\smash{\overline{\omega}=w(\overline{{\boldsymbol{k}}})}, where w(𝒌)=w(𝒌)w({\boldsymbol{k}})=-w(-{\boldsymbol{k}}) is the function that solves Λ(w(𝒌),𝒌)=0\Lambda(w({\boldsymbol{k}}),{\boldsymbol{k}})=0. Also importantly, (305) ensures that

Λ(ω¯,𝒌¯)\displaystyle\partial_{\sqbullet}\Lambda(\overline{\omega},\overline{{\boldsymbol{k}}}) =((𝜼)𝚵H𝜼+𝜼(𝚵H)𝜼+𝜼𝚵H(𝜼))|(ω,𝒌)=(ω¯,𝒌¯)\displaystyle=((\partial_{\sqbullet}{\boldsymbol{\eta}}^{\dagger}){\boldsymbol{\Xi}}_{\text{H}}{\boldsymbol{\eta}}+{\boldsymbol{\eta}}^{\dagger}(\partial_{\sqbullet}{\boldsymbol{\Xi}}_{\text{H}}){\boldsymbol{\eta}}+{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}_{\text{H}}(\partial_{\sqbullet}{\boldsymbol{\eta}}))\big{|}_{(\omega,{\boldsymbol{k}})=(\overline{\omega},\overline{{\boldsymbol{k}}})}
=(𝜼(𝚵H)𝜼)|(ω,𝒌)=(ω¯,𝒌¯),\displaystyle=({\boldsymbol{\eta}}^{\dagger}(\partial_{\sqbullet}{\boldsymbol{\Xi}}_{\text{H}}){\boldsymbol{\eta}})\big{|}_{(\omega,{\boldsymbol{k}})=(\overline{\omega},\overline{{\boldsymbol{k}}})}, (307)

where \smash{\sqbullet} can be replaced with any variable.

7.2 Conservative eikonal waves

7.2.1 Basic properties

In case of a quasimonochromatic eikonal wave and, possibly, inhomogeneous non-stationary plasma, one can apply the same arguments as in section 7.1 except the above equalities are now satisfied up to 𝒪(ϵ)\smash{\mathcal{O}(\epsilon)}. For a single-mode wave, one has

𝚿~(t,𝒙)=re(𝚿~c(t,𝒙))+𝒪(ϵ),𝚿~c=eiθ(t,𝒙)𝜼(t,𝒙)a˘(t,𝒙),\displaystyle\widetilde{{\boldsymbol{\Psi}}}(t,{\boldsymbol{x}})=\operatorname{re}(\widetilde{{\boldsymbol{\Psi}}}_{\text{c}}(t,{\boldsymbol{x}}))+\mathcal{O}(\epsilon),\qquad\widetilde{{\boldsymbol{\Psi}}}_{\text{c}}=\mathrm{e}^{\mathrm{i}\theta(t,{\boldsymbol{x}})}{\boldsymbol{\eta}}(t,{\boldsymbol{x}}){\breve{a}}(t,{\boldsymbol{x}}), (308)

where the local frequency and the wavevector,

ω¯tθ,𝒌¯𝒙θ\displaystyle\overline{\omega}\doteq-\partial_{t}\theta,\qquad\overline{{\boldsymbol{k}}}\doteq\partial_{{\boldsymbol{x}}}\theta (309)

are slow functions of (t,𝒙)\smash{(t,{\boldsymbol{x}})}, and so is 𝜼(t,𝒙)𝜼(t,𝒙,ω¯(t,𝒙),𝒌¯(t,𝒙))\smash{{\boldsymbol{\eta}}(t,{\boldsymbol{x}})\doteq{\boldsymbol{\eta}}(t,{\boldsymbol{x}},\overline{\omega}(t,{\boldsymbol{x}}),\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}}))}, which satisfies (294). Then,

𝔏¯=14Λ(t,𝒙,ω¯,𝒌¯)|a˘(t,𝒙)|2+𝒪(ϵ).\displaystyle\overline{\mathfrak{L}}=\frac{1}{4}\,\Lambda(t,{\boldsymbol{x}},\overline{\omega},\overline{{\boldsymbol{k}}})\,|{\breve{a}}(t,{\boldsymbol{x}})|^{2}+\mathcal{O}(\epsilon). (310)

Within the leading-order theory, the term 𝒪(ϵ)\smash{\mathcal{O}(\epsilon)} is neglected.313131Corrections to the lowest-order dispersion relation produce the so-called spin Hall effect; see (Dodin et al., 2019; Ruiz & Dodin, 2017a) for an overview and (Bliokh et al., 2015; Ruiz & Dodin, 2015a; Oancea et al., 2020; Andersson et al., 2021) for examples. These corrections are beyond the accuracy of the model considered, so they will be ignored. Then, the least action principle

0=δS[θ,𝒂˘,𝒂˘]δa˘b14Λb(ω¯,𝒌¯)a˘b,0=δS[θ,𝒂˘,𝒂˘]δa˘b14a˘bΛb(ω¯,𝒌¯)\displaystyle 0=\frac{\delta S[\theta,{\breve{{\boldsymbol{a}}}}^{*},{\breve{{\boldsymbol{a}}}}]}{\delta{\breve{a}}^{b*}}\approx\frac{1}{4}\,\Lambda_{b}(\overline{\omega},\overline{{\boldsymbol{k}}}){\breve{a}}^{b},\qquad 0=\frac{\delta S[\theta,{\breve{{\boldsymbol{a}}}}^{*},{\breve{{\boldsymbol{a}}}}]}{\delta{\breve{a}}^{b}}\approx\frac{1}{4}\,{\breve{a}}^{b*}\Lambda_{b}(\overline{\omega},\overline{{\boldsymbol{k}}}) (311)

leads to the same (but now local) dispersion relation as for monochromatic waves, Λ(t,𝒙,ω¯,𝒌¯)=0\smash{\Lambda(t,{\boldsymbol{x}},\overline{\omega},\overline{{\boldsymbol{k}}})=0}. This shows that quasimonochromatic waves are also on-shell, and thus they satisfy (307) as well. Also notice that the dispersion relation can now be understood as a Hamilton–Jacobi equation for the eikonal phase θ\smash{\theta}:

Λ(t,𝒙,tθ,𝒙θ)=0.\displaystyle\Lambda(t,{\boldsymbol{x}},-\partial_{t}\theta,\partial_{{\boldsymbol{x}}}\theta)=0. (312)

Like in the previous section, let us introduce the function w\smash{w} that solves

Λ(t,𝒙,w(t,𝒙,𝒌),𝒌)=0\displaystyle\Lambda(t,{\boldsymbol{x}},w(t,{\boldsymbol{x}},{\boldsymbol{k}}),{\boldsymbol{k}})=0 (313)

and therefore satisfies

w(t,𝒙,𝒌)=w(t,𝒙,𝒌).\displaystyle w(t,{\boldsymbol{x}},{\boldsymbol{k}})=-w(t,{\boldsymbol{x}},-{\boldsymbol{k}}). (314)

Differentiating (313) with respect to tt, 𝒙{\boldsymbol{x}}, and 𝒌{\boldsymbol{k}} leads to

tΛ+(ωΛ)tw=0,\displaystyle\partial_{t}\Lambda+(\partial_{\omega}\Lambda)\partial_{t}w=0, (315a)
𝒙Λ+(ωΛ)𝒙w=0,\displaystyle\partial_{{\boldsymbol{x}}}\Lambda+(\partial_{\omega}\Lambda)\partial_{{\boldsymbol{x}}}w=0, (315b)
𝒌Λ+(ωΛ)𝒌w=0,\displaystyle\partial_{{\boldsymbol{k}}}\Lambda+(\partial_{\omega}\Lambda)\partial_{{\boldsymbol{k}}}w=0, (315c)

where the derivatives of Λ\Lambda are evaluated at (t,𝒙,w(t,𝒙,𝒌),𝒌)\smash{(t,{\boldsymbol{x}},w(t,{\boldsymbol{x}},{\boldsymbol{k}}),{\boldsymbol{k}})}. In particular, (315c) gives

𝒗gw𝒌=𝒌ΛωΛ,\displaystyle{\boldsymbol{v}}_{\text{g}}\doteq\frac{\partial w}{\partial{\boldsymbol{k}}}=-\frac{\partial_{{\boldsymbol{k}}}\Lambda}{\partial_{\omega}\Lambda}, (316)

for the group velocity 𝒗g\smash{{\boldsymbol{v}}_{\text{g}}}, whose physical meaning is to be specified shortly.

Because θ\theta is now an additional dynamical variable, one also obtains an additional Euler–Lagrange equation:

0=δθS[θ,a˘,a˘]=t+𝒙𝓙,\displaystyle 0=\delta_{\theta}S[\theta,{\breve{a}}^{*},{\breve{a}}]=\partial_{t}\mathcal{I}+\partial_{{\boldsymbol{x}}}\cdot{\boldsymbol{\mathcal{J}}}, (317)

where \mathcal{I} is called the action density and 𝓙{\boldsymbol{\mathcal{J}}} is the action flux density:

𝔏ω=|a˘|24Λω=|a˘|24𝜼𝚵Hω𝜼,\displaystyle\mathcal{I}\doteq\frac{\partial\mathfrak{L}}{\partial\omega}=\frac{|{\breve{a}}|^{2}}{4}\frac{\partial\Lambda}{\partial\omega}=\frac{|{\breve{a}}|^{2}}{4}\,{\boldsymbol{\eta}}^{\dagger}\,\frac{\partial{\boldsymbol{\Xi}}_{\text{H}}}{\partial\omega}\,{\boldsymbol{\eta}}, (318)
𝒥i𝔏ki=|a˘|24Λki=|a˘|24𝜼𝚵Hki𝜼,\displaystyle\mathcal{J}^{i}\doteq-\frac{\partial\mathfrak{L}}{\partial k_{i}}=-\frac{|{\breve{a}}|^{2}}{4}\frac{\partial\Lambda}{\partial k_{i}}=-\frac{|{\breve{a}}|^{2}}{4}\,{\boldsymbol{\eta}}^{\dagger}\,\frac{\partial{\boldsymbol{\Xi}}_{\text{H}}}{\partial k_{i}}\,{\boldsymbol{\eta}}, (319)

where we used (307) and the derivatives are evaluated on (t,𝒙,w(t,𝒙,𝒌¯(t,𝒙)),𝒌¯(t,𝒙))\smash{(t,{\boldsymbol{x}},w(t,{\boldsymbol{x}},\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}})),\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}}))}. Using (316), one can also rewrite (319) as

𝓙=𝒗¯g,𝒗¯g(t,𝒙)𝒗g(t,𝒙,𝒌¯(t,𝒙)).\displaystyle{\boldsymbol{\mathcal{J}}}=\overline{{\boldsymbol{v}}}_{\text{g}}\mathcal{I},\qquad\overline{{\boldsymbol{v}}}_{\text{g}}(t,{\boldsymbol{x}})\doteq{\boldsymbol{v}}_{\text{g}}(t,{\boldsymbol{x}},\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}})). (320)

(The arguments (t,𝒙)(t,{\boldsymbol{x}}) will be omitted from now on for brevity. We will also use (𝒌)({\boldsymbol{k}}) as a shorthand for (w(𝒌),𝒌)(w({\boldsymbol{k}}),{\boldsymbol{k}}) where applicable.) Then, (317) becomes

t+𝒙(𝒗¯g)=0,\displaystyle\partial_{t}\mathcal{I}+\partial_{{\boldsymbol{x}}}\cdot(\overline{{\boldsymbol{v}}}_{\text{g}}\mathcal{I})=0, (321)

which can be a understood as a continuity equation for quasiparticles (‘photons’ or, more generally, ‘wave quanta’) with density \smash{\mathcal{I}} and fluid velocity 𝒗¯g\smash{\overline{{\boldsymbol{v}}}_{\text{g}}} (see also section 7.2.2). Thus, if an eikonal wave satisfies the least-action principle, its total action d𝒙\smash{\int\mathrm{d}{\boldsymbol{x}}\,\mathcal{I}} (‘number of quanta’) is an invariant. This conservation law can be attributed to the fact that the wave Lagrangian density 𝔏¯\smash{\overline{\mathfrak{L}}} depends on derivatives of θ\smash{\theta} but not on θ\smash{\theta} per se.

Also notice the following. By expanding (310) in tθ\smash{\partial_{t}\theta} around tθ=w(t,𝒙,𝒙θ)\smash{\partial_{t}\theta=-w(t,{\boldsymbol{x}},\partial_{{\boldsymbol{x}}}\theta)}, which is satisfied on any solution, one obtains

𝔏¯14(tθ+w(t,𝒙,𝒙θ))ωΛ|a˘|2=(tθ+w(t,𝒙,𝒙θ)),\displaystyle\overline{\mathfrak{L}}\approx-\frac{1}{4}\,(\partial_{t}\theta+w(t,{\boldsymbol{x}},\partial_{{\boldsymbol{x}}}\theta))\partial_{\omega}\Lambda\,|{\breve{a}}|^{2}=-(\partial_{t}\theta+w(t,{\boldsymbol{x}},\partial_{{\boldsymbol{x}}}\theta))\mathcal{I}, (322)

where we used that 𝔏¯(t,𝒙,tθ,𝒙θ)=0\smash{\overline{\mathfrak{L}}(t,{\boldsymbol{x}},-\partial_{t}\theta,\partial_{{\boldsymbol{x}}}\theta)=0} due to (312). Then, one arrives at the canonical form of the action integral (Hayes, 1973)

S[,θ]=dtd𝒙(tθ+w(t,𝒙,𝒌)).\displaystyle S[\mathcal{I},\theta]=-\int\mathrm{d}t\,\mathrm{d}{\boldsymbol{x}}\,(\partial_{t}\theta+w(t,{\boldsymbol{x}},{\boldsymbol{k}}))\mathcal{I}. (323)

From here, δS=0\smash{\delta_{\mathcal{I}}S=0} yields the dispersion relation in the Hamilton–Jacobi form tθ+w(t,𝒙,𝒌)=0\smash{\partial_{t}\theta+w(t,{\boldsymbol{x}},{\boldsymbol{k}})=0}, and δθS=0\smash{\delta_{\theta}S=0} yields the action conservation (321).

7.2.2 Ray equations

By (309), one has the so-called consistency relations:

tk¯i+iω¯=0,ik¯j=jk¯i.\displaystyle\partial_{t}\overline{k}_{i}+\partial_{i}\overline{\omega}=0,\qquad\partial_{i}\overline{k}_{j}=\partial_{j}\overline{k}_{i}. (324)

These lead to

(t\displaystyle\bigg{(}\frac{\partial}{\partial t} +𝒗¯g𝒙)k¯i(t,𝒙)=w(t,𝒙,𝒌¯(t,𝒙))xi+𝒗¯gk¯i(t,𝒙)𝒙\displaystyle+\overline{{\boldsymbol{v}}}_{\text{g}}\cdot\frac{\partial}{\partial{\boldsymbol{x}}}\bigg{)}\overline{k}_{i}(t,{\boldsymbol{x}})=-\frac{\partial w(t,{\boldsymbol{x}},\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}}))}{\partial x^{i}}+\overline{{\boldsymbol{v}}}_{\text{g}}\cdot\frac{\partial\overline{k}_{i}(t,{\boldsymbol{x}})}{\partial{\boldsymbol{x}}}
=(w(t,𝒙,𝒌)xi)𝒌=𝒌¯(t,𝒙)v¯gjk¯j(t,𝒙)xi+v¯gjk¯i(t,𝒙)xj\displaystyle=-\left(\frac{\partial w(t,{\boldsymbol{x}},{\boldsymbol{k}})}{\partial x^{i}}\right)_{{\boldsymbol{k}}=\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}})}-\overline{v}_{\text{g}}^{j}\,\frac{\partial\overline{k}_{j}(t,{\boldsymbol{x}})}{\partial x^{i}}+\overline{v}_{\text{g}}^{j}\,\frac{\partial\overline{k}_{i}(t,{\boldsymbol{x}})}{\partial x^{j}}
=(w(t,𝒙,𝒌)xi)𝒌=𝒌¯(t,𝒙),\displaystyle=-\left(\frac{\partial w(t,{\boldsymbol{x}},{\boldsymbol{k}})}{\partial x^{i}}\right)_{{\boldsymbol{k}}=\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}})}, (325)

and similarly,

(t\displaystyle\bigg{(}\frac{\partial}{\partial t} +𝒗¯g𝒙)ω¯(t,𝒙)=(t+𝒗¯g𝒙)w(t,𝒙,𝒌¯(t,𝒙))\displaystyle+\overline{{\boldsymbol{v}}}_{\text{g}}\cdot\frac{\partial}{\partial{\boldsymbol{x}}}\bigg{)}\overline{\omega}(t,{\boldsymbol{x}})=\bigg{(}\frac{\partial}{\partial t}+\overline{{\boldsymbol{v}}}_{\text{g}}\cdot\frac{\partial}{\partial{\boldsymbol{x}}}\bigg{)}w(t,{\boldsymbol{x}},\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}}))
=(w(t,𝒙,𝒌)t+v¯giw(t,𝒙,𝒌)xi)𝒌=𝒌¯(t,𝒙)+v¯gi(t+𝒗¯g𝒙)k¯i(t,𝒙)\displaystyle=\left(\frac{\partial w(t,{\boldsymbol{x}},{\boldsymbol{k}})}{\partial t}+\overline{v}_{\text{g}}^{i}\,\frac{\partial w(t,{\boldsymbol{x}},{\boldsymbol{k}})}{\partial x^{i}}\right)_{{\boldsymbol{k}}=\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}})}+\overline{v}_{\text{g}}^{i}\left(\frac{\partial}{\partial t}+\overline{{\boldsymbol{v}}}_{\text{g}}\cdot\frac{\partial}{\partial{\boldsymbol{x}}}\right)\overline{k}_{i}(t,{\boldsymbol{x}})
=(w(t,𝒙,𝒌)t)𝒌=𝒌¯(t,𝒙),\displaystyle=\left(\frac{\partial w(t,{\boldsymbol{x}},{\boldsymbol{k}})}{\partial t}\right)_{{\boldsymbol{k}}=\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}})}, (326)

where we used (325). Using the convective derivative associated with the group velocity,

d/dtdtt+(𝒗¯g𝒙),\displaystyle\mathrm{d}/\mathrm{d}t\equiv\mathrm{d}_{t}\doteq\partial_{t}+(\overline{{\boldsymbol{v}}}_{\text{g}}\cdot\partial_{\boldsymbol{x}}), (327)

one can rewrite these compactly as

dk¯i(t,𝒙)dt=(w(t,𝒙,𝒌)xi)𝒌=𝒌¯(t,𝒙),dω¯(t,𝒙)dt=(w(t,𝒙,𝒌)t)𝒌=𝒌¯(t,𝒙).\displaystyle\frac{\mathrm{d}\overline{k}_{i}(t,{\boldsymbol{x}})}{\mathrm{d}t}=-\left(\frac{\partial w(t,{\boldsymbol{x}},{\boldsymbol{k}})}{\partial x^{i}}\right)_{{\boldsymbol{k}}=\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}})},\quad\frac{\mathrm{d}\overline{\omega}(t,{\boldsymbol{x}})}{\mathrm{d}t}=\left(\frac{\partial w(t,{\boldsymbol{x}},{\boldsymbol{k}})}{\partial t}\right)_{{\boldsymbol{k}}=\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}})}. (328)

One can also represent (328) as ordinary differential equations for 𝒌¯(t)𝒌¯(t,𝒙¯(t))\smash{\overline{{\boldsymbol{k}}}(t)\doteq\overline{{\boldsymbol{k}}}(t,\overline{{\boldsymbol{x}}}(t))} and ω¯(t)ω¯(t,𝒙¯(t))\smash{\overline{\omega}(t)\doteq\overline{\omega}(t,\overline{{\boldsymbol{x}}}(t))}, where 𝒙¯(t)\smash{\overline{{\boldsymbol{x}}}(t)} are the ‘ray trajectories’ governed by

dx¯i(t)dt=vgi(t,𝒙¯(t),𝒌¯(t)).\displaystyle\frac{\mathrm{d}\overline{x}^{i}(t)}{\mathrm{d}t}=v_{\text{g}}^{i}(t,\overline{{\boldsymbol{x}}}(t),\overline{{\boldsymbol{k}}}(t)). (329)

Specifically, together with (329), equations (328) become Hamilton’s equations also known as the ray equations:

dx¯idt=w(t,𝒙¯,𝒌¯)k¯i,dk¯idt=w(t,𝒙¯,𝒌¯)x¯i,dω¯dt=w(t,𝒙¯,𝒌¯)t,\displaystyle\frac{\mathrm{d}\overline{x}^{i}}{\mathrm{d}t}=\frac{\partial w(t,\overline{{\boldsymbol{x}}},\overline{{\boldsymbol{k}}})}{\partial\overline{k}_{i}},\qquad\frac{\mathrm{d}\overline{k}_{i}}{\mathrm{d}t}=-\frac{\partial w(t,\overline{{\boldsymbol{x}}},\overline{{\boldsymbol{k}}})}{\partial\overline{x}^{i}},\qquad\frac{\mathrm{d}\overline{\omega}}{\mathrm{d}t}=\frac{\partial w(t,\overline{{\boldsymbol{x}}},\overline{{\boldsymbol{k}}})}{\partial t}, (330)

where 𝒙¯\smash{\overline{{\boldsymbol{x}}}} is the coordinate, 𝒌¯\smash{\hbar\overline{{\boldsymbol{k}}}} is the momentum, ω¯\smash{\hbar\overline{\omega}} is the energy, w\smash{\hbar w} is the Hamiltonian, and the constant factor \smash{\hbar} can be anything. If \smash{\hbar} is chosen to be the Planck constant, then (330) can be interpreted as the motion equations of individual wave quanta, for example, photons. Hamilton’s equations for ‘true’ particles, such as electrons and ions, are also subsumed under (330) in that they can be understood as the ray equations of the particles considered as quantum-matter waves in the semiclassical limit.

Also notably, (330) can be obtained by considering the point-particle limit of (323) (Ruiz & Dodin, 2015b). Specifically, adopting (t,𝒙)δ(𝒙𝒙¯(t))\smash{\mathcal{I}(t,{\boldsymbol{x}})\,\propto\,\delta({\boldsymbol{x}}-\overline{{\boldsymbol{x}}}(t))} and taking the integral in (323) by parts leads to a canonical action Sdt(𝒌¯𝒙¯˙w(t,𝒙¯,𝒌¯))\smash{S\,\propto\,\int\mathrm{d}t\,(\overline{{\boldsymbol{k}}}\cdot\dot{\overline{{\boldsymbol{x}}}}-w(t,\overline{{\boldsymbol{x}}},\overline{{\boldsymbol{k}}}))}, whence Hamilton’s equations follow as usual.

7.2.3 Wave momentum and energy

Using (321) and (327), one arrives at the following equality for any given field 𝖷\smash{\mathsf{X}}:

t(𝖷)+𝒙(𝖷𝒗¯g)\displaystyle\partial_{t}(\mathsf{X}\mathcal{I})+\partial_{\boldsymbol{x}}\cdot(\mathsf{X}\mathcal{I}\overline{{\boldsymbol{v}}}_{\text{g}}) =(t𝖷)+𝖷(t)+[𝒙(𝒗¯g)]𝖷+(𝒗¯g𝒙)𝖷\displaystyle=(\partial_{t}\mathsf{X})\mathcal{I}+\mathsf{X}(\partial_{t}\mathcal{I})+[\partial_{\boldsymbol{x}}\cdot(\mathcal{I}\overline{{\boldsymbol{v}}}_{\text{g}})]\mathsf{X}+\mathcal{I}(\overline{{\boldsymbol{v}}}_{\text{g}}\cdot\partial_{\boldsymbol{x}})\mathsf{X}
=[t+(𝒗¯g𝒙)]𝖷+𝖷[t+𝒙(𝒗¯g)]\displaystyle=\mathcal{I}[\partial_{t}+(\overline{{\boldsymbol{v}}}_{\text{g}}\cdot\partial_{\boldsymbol{x}})]\mathsf{X}+\mathsf{X}[\partial_{t}\mathcal{I}+\partial_{\boldsymbol{x}}\cdot(\mathcal{I}\overline{{\boldsymbol{v}}}_{\text{g}})]
=dt𝖷.\displaystyle=\mathcal{I}\,\mathrm{d}_{t}\mathsf{X}. (331)

For 𝖷=k¯i\smash{\mathsf{X}=\overline{k}_{i}} and 𝖷=ω¯\smash{\mathsf{X}=\overline{\omega}}, (331) yields, respectively,

tPw,i+𝒙(𝒗¯gPw,i)\displaystyle\partial_{t}P_{\text{w},i}+\partial_{{\boldsymbol{x}}}\cdot(\overline{{\boldsymbol{v}}}_{\text{g}}P_{\text{w},i}) =iw,\displaystyle=-\mathcal{I}\partial_{i}w, (332a)
tw+𝒙(𝒗¯gw)\displaystyle\partial_{t}\mathcal{E}_{\text{w}}+\partial_{{\boldsymbol{x}}}\cdot(\overline{{\boldsymbol{v}}}_{\text{g}}\mathcal{E}_{\text{w}}) =tw,\displaystyle=\mathcal{I}\partial_{t}w, (332b)

where we used (328) and introduced the following notation:

𝑷w𝒌¯,wω¯.\displaystyle{\boldsymbol{P}}_{\text{w}}\doteq\overline{{\boldsymbol{k}}}\mathcal{I},\qquad\mathcal{E}_{\text{w}}\doteq\overline{\omega}\mathcal{I}. (333)

When a medium is homogeneous along xi\smash{x^{i}}, (332a) yields d𝒙Pw,i=const\smash{\int\mathrm{d}{\boldsymbol{x}}\,P_{\text{w},i}=\text{const}}. Likewise, when a medium is stationary, (332b) yields d𝒙w=const\smash{\int\mathrm{d}{\boldsymbol{x}}\,\mathcal{E}_{\text{w}}=\text{const}}. Hence, by definition, 𝑷w\smash{{\boldsymbol{P}}_{\text{w}}} and w\smash{\mathcal{E}_{\text{w}}} are the densities of the wave canonical momentum and energy, at least up to a constant factor κ\smash{\kappa}.323232Therefore, in a zero-dimensional wave, where d𝒙\smash{\int\mathrm{d}{\boldsymbol{x}}} can be omitted, conservation of the total action \smash{\mathcal{I}} implies conservation of w/ω\smash{\mathcal{E}_{\text{w}}/\omega}, which is a well-known adiabatic invariant of a discrete harmonic oscillator with a slowly varying frequency (Landau & Lifshitz, 1976, section 49). A proof that κ=1\smash{\kappa=1} can be found, for example, in (Dodin & Fisch, 2012). In section 7.5, we will show this using different arguments.

7.3 Non-conservative eikonal waves

In a medium with nonzero 𝚵A\smash{{\boldsymbol{\Xi}}_{\text{A}}}, where waves are non-conservative, the wave properties are defined as in the previous section but the wave action evolves differently. The variational principle is not easy to apply in this case (however, see (Dodin et al., 2017)), so a different approach will be used to derive the action equation. A more straightforward but less intuitive approach can be found in (Dodin et al., 2019; McDonald, 1988).

7.3.1 Monochromatic waves

First, consider a homogeneous stationary medium and a ‘monochromatic’ (exponentially growing at a constant rate) wave field in the form

𝚿~(t,𝒙)=re(eiω¯t+i𝒌¯𝒙𝚿˘c),𝚿˘c=eγ¯t×const,\displaystyle\widetilde{{\boldsymbol{\Psi}}}(t,{\boldsymbol{x}})=\operatorname{re}(\mathrm{e}^{-\mathrm{i}\overline{\omega}t+\mathrm{i}\overline{{\boldsymbol{k}}}\cdot{\boldsymbol{x}}}\,{\breve{{\boldsymbol{\Psi}}}}_{\text{c}}),\qquad{\breve{{\boldsymbol{\Psi}}}}_{\text{c}}=\mathrm{e}^{\overline{\gamma}t}\times\text{const}, (334)

where the constants ω¯\smash{\overline{\omega}} and 𝒌¯\smash{\overline{{\boldsymbol{k}}}} are, as usual, the real frequency and wavenumber, and γ¯\smash{\overline{\gamma}} is the linear growth rate, which can have either sign. Then, (222) becomes

𝟎=𝚵(ω¯+iγ¯,𝒌¯)𝚿˘c=𝚵H(ω¯,𝒌¯)𝚿˘c+i(γ¯ω𝚵H(ω¯,𝒌¯)+𝚵A(ω¯,𝒌¯))𝚿˘c+𝒪(ϵ2),\displaystyle{\boldsymbol{0}}={\boldsymbol{\Xi}}(\overline{\omega}+\mathrm{i}\overline{\gamma},\overline{{\boldsymbol{k}}}){\breve{{\boldsymbol{\Psi}}}}_{\text{c}}={\boldsymbol{\Xi}}_{\text{H}}(\overline{\omega},\overline{{\boldsymbol{k}}}){\breve{{\boldsymbol{\Psi}}}}_{\text{c}}+\mathrm{i}(\overline{\gamma}\partial_{\omega}{\boldsymbol{\Xi}}_{\text{H}}(\overline{\omega},\overline{{\boldsymbol{k}}})+{\boldsymbol{\Xi}}_{\text{A}}(\overline{\omega},\overline{{\boldsymbol{k}}})){\breve{{\boldsymbol{\Psi}}}}_{\text{c}}+\mathcal{O}(\epsilon^{2}), (335)

where we assume that 𝚵\smash{{\boldsymbol{\Xi}}} is a smooth function of ω\smash{\omega} and also that both 𝚵A\smash{{\boldsymbol{\Xi}}_{\text{A}}} and γ¯\smash{\overline{\gamma}} are 𝒪(ϵ)\smash{\mathcal{O}(\epsilon)}. Like in section 7.2.1, we adopt 𝚿˘c=𝜼a˘+𝒪(ϵ)\smash{{\breve{{\boldsymbol{\Psi}}}}_{\text{c}}={\boldsymbol{\eta}}{\breve{a}}+\mathcal{O}(\epsilon)}, where the polarization vector 𝜼\smash{{\boldsymbol{\eta}}} is the relevant eigenvector of 𝚵H\smash{{\boldsymbol{\Xi}}_{\text{H}}}. Then, by projecting (335) on 𝜼\smash{{\boldsymbol{\eta}}}, one obtains

0=Λ(ω¯,𝒌¯)a˘+i(γ¯ωΛ+𝜼𝚵A𝜼)|(ω,𝒌)=(ω¯,𝒌¯)a˘+𝒪(ϵ2),\displaystyle 0=\Lambda(\overline{\omega},\overline{{\boldsymbol{k}}}){\breve{a}}+\mathrm{i}(\overline{\gamma}\partial_{\omega}\Lambda+{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}_{\text{A}}{\boldsymbol{\eta}})\big{|}_{(\omega,{\boldsymbol{k}})=(\overline{\omega},\overline{{\boldsymbol{k}}})}{\breve{a}}+\mathcal{O}(\epsilon^{2}), (336)

where Λ=𝜼𝚵H𝜼\smash{\Lambda={\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}_{\text{H}}{\boldsymbol{\eta}}} is the corresponding eigenvalue of 𝚵H\smash{{\boldsymbol{\Xi}}_{\text{H}}} and we used (307). Let us neglect 𝒪(ϵ2)\smash{\mathcal{O}(\epsilon^{2})}, divide (336) by a˘\smash{{\breve{a}}}, and consider the real and imaginary parts of the resulting equation separately:

Λ(ω¯,𝒌¯)=0,(γ¯ωΛ+𝜼𝚵A𝜼)|(ω,𝒌)=0.\displaystyle\Lambda(\overline{\omega},\overline{{\boldsymbol{k}}})=0,\qquad(\overline{\gamma}\partial_{\omega}\Lambda+{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}_{\text{A}}{\boldsymbol{\eta}})\big{|}_{(\omega,{\boldsymbol{k}})}=0. (337)

The former is the same dispersion relation for ω¯\smash{\overline{\omega}} as for conservative waves, and the latter yields γ¯=γ(𝒌¯)\overline{\gamma}=\gamma({\boldsymbol{\overline{k}}}), where

γ(𝒌)𝜼𝚵A𝜼ωΛ.\displaystyle\gamma({\boldsymbol{k}})\doteq-\frac{{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}_{\text{A}}{\boldsymbol{\eta}}}{\partial_{\omega}\Lambda}. (338)

Because |a˘|eγ¯t\smash{|{\breve{a}}|\,\propto\,\mathrm{e}^{\overline{\gamma}t}}, one can write the amplitude equation as

t|a˘|2=2γ¯|a˘|2.\displaystyle\partial_{t}|{\breve{a}}|^{2}=2\overline{\gamma}|{\breve{a}}|^{2}. (339)

One can also define the action density \smash{\mathcal{I}} as in section 7.2.1 and rewrite (339) in terms of that. Because =|a˘|2×const\smash{\mathcal{I}=|{\breve{a}}|^{2}\times\text{const}}, one obtains

t=2γ¯.\displaystyle\partial_{t}\mathcal{I}=2\overline{\gamma}\mathcal{I}. (340)

7.3.2 Non-monochromatic waves

When weak inhomogeneity and weak dissipation coexist, their effect on the action density is additive, so (321) and (340) merge into a general equation

t+𝒙(𝒗¯g)=2γ¯.\displaystyle\partial_{t}\mathcal{I}+\partial_{\boldsymbol{x}}(\overline{{\boldsymbol{v}}}_{\text{g}}\mathcal{I})=2\overline{\gamma}\mathcal{I}. (341)

(A formal derivation of (341), which uses the Weyl expansion (41) and projection of the field equation on the polarization vector, can be found in (Dodin et al., 2019).) Then, (331) is modified as follows:

t(𝖷)+𝒙(𝖷𝒗¯g)=dt𝖷+2γ¯𝖷,\displaystyle\partial_{t}(\mathsf{X}\mathcal{I})+\partial_{\boldsymbol{x}}\cdot(\mathsf{X}\mathcal{I}\overline{{\boldsymbol{v}}}_{\text{g}})=\mathcal{I}\,\mathrm{d}_{t}\mathsf{X}+2\overline{\gamma}\mathsf{X}\mathcal{I}, (342)

and the equations (332) for the wave momentum and energy (333) become

tPw,i+𝒙(𝒗¯gPw,i)\displaystyle\partial_{t}P_{\text{w},i}+\partial_{{\boldsymbol{x}}}\cdot(\overline{{\boldsymbol{v}}}_{\text{g}}P_{\text{w},i}) =2γ¯Pw,iiw,\displaystyle=2\overline{\gamma}P_{\text{w},i}-\mathcal{I}\partial_{i}w, (343a)
tw+𝒙(𝒗¯gw)\displaystyle\partial_{t}\mathcal{E}_{\text{w}}+\partial_{{\boldsymbol{x}}}\cdot(\overline{{\boldsymbol{v}}}_{\text{g}}\mathcal{E}_{\text{w}}) =2γ¯w+tw.\displaystyle=2\overline{\gamma}\mathcal{E}_{\text{w}}+\mathcal{I}\partial_{t}w. (343b)

A comment is due here regarding the relation between (341) and the amplitude equation (339) that is commonly used in the standard QLT for homogeneous plasma (for example, see (2.21) in (Drummond & Pines, 1962)). In a nutshell, the latter is incorrect, even when 𝒙=0\smash{\partial_{\boldsymbol{x}}=0}. Because f¯\smash{\overline{f}} is time-dependent, waves do not grow or decay exponentially. Rather, they can be considered as geometrical-optics (WKB) waves, and unlike in section 7.3.1, the ratio |a˘|2/\smash{|{\breve{a}}|^{2}/\mathcal{I}} generally evolves at a rate comparable to γ¯\smash{\overline{\gamma}}. The standard QLT remains conservative only because it also incorrectly replaces (103) with its stationary-plasma limit (ϵ=0\smash{\epsilon=0}) and the two errors cancel each other. These issues are less of a problem for waves in not-too-hot plasmas (\egLangmuir waves), because in such plasmas, changing the distribution functions does not significantly affect the dispersion relations and thus |a˘|2/\smash{|{\breve{a}}|^{2}/\mathcal{I}} does in fact approximately remain constant. See also the discussion in section 9.1.4.

7.4 General waves

Let us now discuss a more general case that includes broadband waves. The evolution of such waves can be described statistically in terms of their average Wigner matrix 𝗨\smash{{\boldsymbol{\mathsf{U}}}}. This matrix also determines the function 𝖶¯s\smash{\overline{\mathsf{W}}_{s}} that is given by (244a) and enters the nonlinear potentials (284). Below, we derive the general form of 𝗨\smash{{\boldsymbol{\mathsf{U}}}} in terms of the phase-space action density J\smash{J} and the governing equation for J\smash{J} (sections 7.4.17.4.3). Then, we also express the function 𝖶¯s\smash{\overline{\mathsf{W}}_{s}} through J\smash{J} (section 7.4.4). Related calculations can also be found in (McDonald & Kaufman, 1985; Ruiz, 2017).

7.4.1 Average Wigner matrix of an eikonal wave field

Let us start with calculating the average Wigner matrix of an eikonal field 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}} of the form (308) (see also appendix A.2). Using 𝚿~=(𝚿~c+𝚿~c)/2\smash{\widetilde{{\boldsymbol{\Psi}}}=(\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}+\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}^{*})/2}, it can be readily expressed through the average Wigner functions of the complexified field333333Field complexification is discussed, for example, in (Brizard et al., 1993). 𝚿~c\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}} and of its complex conjugate:

𝗨(𝗪¯𝚿~c+𝗪¯𝚿~c)/4(𝗨c++𝗨c)/4.\displaystyle{\boldsymbol{\mathsf{U}}}\approx(\overline{{\boldsymbol{\mathsf{W}}}}_{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}}+\overline{{\boldsymbol{\mathsf{W}}}}_{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}^{*}})/4\equiv({\boldsymbol{\mathsf{U}}}_{\text{c}+}+{\boldsymbol{\mathsf{U}}}_{\text{c}-})/4. (344)

For 𝚿~c=a˘𝜼(ω¯,𝒌¯)eiθ\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}={\breve{a}}{\boldsymbol{\eta}}(\overline{\omega},\overline{{\boldsymbol{k}}})\mathrm{e}^{\mathrm{i}\theta}}, where the arguments (t,𝒙)(t,{\boldsymbol{x}}) are omitted for brevity, one has

𝗨c𝗨c+(𝜼𝜼)(ω¯,𝒌¯)|a˘|2dτd𝒔(2\upi)𝗇eiθ(t+τ/2,𝒙+𝒔/2)eiθ(tτ/2,𝒙𝒔/2)eiωτi𝒌𝒔,\displaystyle{\boldsymbol{\mathsf{U}}}_{\text{c}}\equiv{\boldsymbol{\mathsf{U}}}_{\text{c}+}\approx({\boldsymbol{\eta}}{\boldsymbol{\eta}}^{\dagger})(\overline{\omega},\overline{{\boldsymbol{k}}})|{\breve{a}}|^{2}\int\frac{\mathrm{d}\tau\,\mathrm{d}{\boldsymbol{s}}}{(2\upi)^{\mathsf{n}}}\,\mathrm{e}^{\mathrm{i}\theta(t+\tau/2,{\boldsymbol{x}}+{\boldsymbol{s}}/2)}\mathrm{e}^{-\mathrm{i}\theta(t-\tau/2,{\boldsymbol{x}}-{\boldsymbol{s}}/2)}\mathrm{e}^{\mathrm{i}\omega\tau-\mathrm{i}{\boldsymbol{k}}\cdot{\boldsymbol{s}}}, (345)

where we neglected the dependence of a˘\smash{{\breve{a}}} and 𝜼\smash{{\boldsymbol{\eta}}} on (t,𝒙)\smash{(t,{\boldsymbol{x}})} because it is weak compared to that of e±iθ\smash{\mathrm{e}^{\pm\mathrm{i}\theta}}. By Taylor-expanding θ\smash{\theta}, one obtains

𝗨c(𝜼𝜼)(ω¯,𝒌¯)|a˘|2dτd𝒔(2\upi)𝗇ei(ωω¯)τi(𝒌𝒌¯)𝒔=(𝜼𝜼)(ω¯,𝒌¯)|a˘|2δ(ωω¯)δ(𝒌𝒌¯).\displaystyle{\boldsymbol{\mathsf{U}}}_{\text{c}}\approx({\boldsymbol{\eta}}{\boldsymbol{\eta}}^{\dagger})(\overline{\omega},\overline{{\boldsymbol{k}}})|{\breve{a}}|^{2}\int\frac{\mathrm{d}\tau\,\mathrm{d}{\boldsymbol{s}}}{(2\upi)^{\mathsf{n}}}\,\mathrm{e}^{\mathrm{i}(\omega-\overline{\omega})\tau-\mathrm{i}({\boldsymbol{k}}-\overline{{\boldsymbol{k}}})\cdot{\boldsymbol{s}}}=({\boldsymbol{\eta}}{\boldsymbol{\eta}}^{\dagger})(\overline{\omega},\overline{{\boldsymbol{k}}})|{\breve{a}}|^{2}\delta(\omega-\overline{\omega})\delta({\boldsymbol{k}}-\overline{{\boldsymbol{k}}}).

For 𝚿~c=a˘𝜼(ω¯,𝒌¯)eiθ\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}^{*}={\breve{a}}^{*}{\boldsymbol{\eta}}^{*}(\overline{\omega},\overline{{\boldsymbol{k}}})\mathrm{e}^{-\mathrm{i}\theta}}, which can also be written as 𝚿~c=a˘𝜼(ω¯,𝒌¯)eiθ\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}^{*}={\breve{a}}^{*}{\boldsymbol{\eta}}(-\overline{\omega},-\overline{{\boldsymbol{k}}})\mathrm{e}^{-\mathrm{i}\theta}} due to (298), the result is the same up to replacing ω¯ω¯\smash{\overline{\omega}\to-\overline{\omega}} and 𝒌¯𝒌¯\smash{\overline{{\boldsymbol{k}}}\to-\overline{{\boldsymbol{k}}}}. Also notice that

δ(ωω¯)δ(𝒌𝒌¯)\displaystyle\delta(\omega\mp\overline{\omega})\delta({\boldsymbol{k}}\mp\overline{{\boldsymbol{k}}}) =δ(ωw(𝒌¯))δ(𝒌𝒌¯)\displaystyle=\delta(\omega\mp w(\overline{{\boldsymbol{k}}}))\delta({\boldsymbol{k}}\mp\overline{{\boldsymbol{k}}})
=δ(ωw(±𝒌))δ(𝒌𝒌¯)\displaystyle=\delta(\omega\mp w(\pm{\boldsymbol{k}}))\delta({\boldsymbol{k}}\mp\overline{{\boldsymbol{k}}})
=δ(ωw(𝒌))δ(𝒌𝒌¯),\displaystyle=\delta(\omega-w({\boldsymbol{k}}))\delta({\boldsymbol{k}}\mp\overline{{\boldsymbol{k}}}), (346)

so one can rewrite 𝗨c±\smash{{\boldsymbol{\mathsf{U}}}_{\text{c}\pm}} as follows:

𝗨c±=𝜼(𝒌)𝜼(𝒌)|a˘|2δ(ωw(𝒌))δ(𝒌𝒌¯),\displaystyle{\boldsymbol{\mathsf{U}}}_{\text{c}\pm}={\boldsymbol{\eta}}({\boldsymbol{k}}){\boldsymbol{\eta}}^{\dagger}({\boldsymbol{k}})|{\breve{a}}|^{2}\delta(\omega-w({\boldsymbol{k}}))\delta({\boldsymbol{k}}\mp\overline{{\boldsymbol{k}}}), (347)

where (𝒌)(w(𝒌),𝒌)\smash{({\boldsymbol{k}})\equiv(w({\boldsymbol{k}}),{\boldsymbol{k}})}. Thus finally,

𝗨(ω,𝒌)𝜼(𝒌)𝜼(𝒌)|a˘|2(δ(𝒌𝒌¯)+δ(𝒌+𝒌¯))δ(ωw(𝒌))/4.\displaystyle{\boldsymbol{\mathsf{U}}}(\omega,{\boldsymbol{k}})\approx{\boldsymbol{\eta}}({\boldsymbol{k}}){\boldsymbol{\eta}}^{\dagger}({\boldsymbol{k}})\,|{\breve{a}}|^{2}(\delta({\boldsymbol{k}}-\overline{{\boldsymbol{k}}})+\delta({\boldsymbol{k}}+\overline{{\boldsymbol{k}}}))\delta(\omega-w({\boldsymbol{k}}))/4. (348)

7.4.2 Average Wigner matrix of a general wave

Assuming the background medium is sufficiently smooth, a general wave field can be represented as a superposition of eikonal fields:

𝚿~=re𝚿~c,𝚿~c=σ𝚿~σ,c,𝚿~σ,c=a˘σeiθσ.\displaystyle\textstyle\widetilde{{\boldsymbol{\Psi}}}=\operatorname{re}\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}},\qquad\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}=\sum_{\sigma}\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\sigma,\text{c}},\qquad\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\sigma,\text{c}}={\breve{a}}_{\sigma}\mathrm{e}^{\mathrm{i}\theta_{\sigma}}. (349)

As a quadratic functional, its average Wigner matrix 𝗨\smash{{\boldsymbol{\mathsf{U}}}} equals the sum of the average Wigner matrices 𝗨σ\smash{{\boldsymbol{\mathsf{U}}}_{\sigma}} of the individual eikonal waves:

𝗨=σ𝗨σ=σ(𝗨σ,c++𝗨σ,c)/4,\displaystyle\textstyle{\boldsymbol{\mathsf{U}}}=\sum_{\sigma}{\boldsymbol{\mathsf{U}}}_{\sigma}=\sum_{\sigma}({\boldsymbol{\mathsf{U}}}_{\sigma,\text{c}+}+{\boldsymbol{\mathsf{U}}}_{\sigma,\text{c}-})/4, (350)

where 𝗨σ,c+𝗨σ,c\smash{{\boldsymbol{\mathsf{U}}}_{\sigma,\text{c}+}\equiv{\boldsymbol{\mathsf{U}}}_{\sigma,\text{c}}} and 𝗨σ,c\smash{{\boldsymbol{\mathsf{U}}}_{\sigma,\text{c}-}} are the average Wigner matrices of 𝚿~σ,c\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\sigma,\text{c}}} and 𝚿~σ,c\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\sigma,\text{c}}^{*}}, respectively:

𝗨σ,c±=𝜼(𝒌)𝜼(𝒌)|a˘σ|2δ(𝒌𝒌¯σ)δ(ωw(𝒌)).\displaystyle{\boldsymbol{\mathsf{U}}}_{\sigma,\text{c}\pm}={\boldsymbol{\eta}}({\boldsymbol{k}}){\boldsymbol{\eta}}^{\dagger}({\boldsymbol{k}})|{\breve{a}}_{\sigma}|^{2}\delta({\boldsymbol{k}}\mp\overline{{\boldsymbol{k}}}_{\sigma})\delta(\omega-w({\boldsymbol{k}})). (351)

Equation (350) can also be expressed as

𝗨=(𝗨c++𝗨c)/4,𝗨c±=σ𝗨σ,c±,\displaystyle\textstyle{\boldsymbol{\mathsf{U}}}=({\boldsymbol{\mathsf{U}}}_{\text{c}+}+{\boldsymbol{\mathsf{U}}}_{\text{c}-})/4,\qquad{\boldsymbol{\mathsf{U}}}_{\text{c}\pm}=\sum_{\sigma}{\boldsymbol{\mathsf{U}}}_{\sigma,\text{c}\pm}, (352)

where 𝗨c±\smash{{\boldsymbol{\mathsf{U}}}_{\text{c}\pm}} are the average Wigner matrices of 𝚿~c\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}} and 𝚿~c\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}^{*}}, respectively:

𝗨c±=𝜼(𝒌)𝜼(𝒌)hc±(𝒌)δ(ωw(𝒌)),hc±(𝒌)σ|a˘σ|2δ(𝒌𝒌¯σ).\displaystyle{\boldsymbol{\mathsf{U}}}_{\text{c}\pm}={\boldsymbol{\eta}}({\boldsymbol{k}}){\boldsymbol{\eta}}^{\dagger}({\boldsymbol{k}})h_{\text{c}\pm}({\boldsymbol{k}})\delta(\omega-w({\boldsymbol{k}})),\qquad h_{\text{c}\pm}({\boldsymbol{k}})\doteq\textstyle\sum_{\sigma}|{\breve{a}}_{\sigma}|^{2}\delta({\boldsymbol{k}}\mp\overline{{\boldsymbol{k}}}_{\sigma}). (353)

Because hc(𝒌)=hc+(𝒌)hc(𝒌)\smash{h_{\text{c}-}({\boldsymbol{k}})=h_{\text{c}+}(-{\boldsymbol{k}})\equiv h_{\text{c}}(-{\boldsymbol{k}})}, the matrix 𝗨\smash{{\boldsymbol{\mathsf{U}}}} can also be written as follows:

𝗨(ω,𝒌)=(𝜼𝜼)(𝒌)(h(𝒌)+h(𝒌))δ(ωw(𝒌)),\displaystyle\textstyle{\boldsymbol{\mathsf{U}}}(\omega,{\boldsymbol{k}})=({\boldsymbol{\eta}}{\boldsymbol{\eta}}^{\dagger})({\boldsymbol{k}})\,(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\delta(\omega-w({\boldsymbol{k}})), (354)

where h(𝒌)hc(𝒌)/4\smash{h({\boldsymbol{k}})\doteq h_{\text{c}}({\boldsymbol{k}})/4} is given by

h(𝒌)=14σ|a˘σ|2δ(𝒌𝒌¯σ)0.\displaystyle\textstyle h({\boldsymbol{k}})=\frac{1}{4}\sum_{\sigma}|{\breve{a}}_{\sigma}|^{2}\delta({\boldsymbol{k}}-\overline{{\boldsymbol{k}}}_{\sigma})\geq 0. (355)

This shows that for broadband waves comprised of eikonal waves, 𝗨\smash{{\boldsymbol{\mathsf{U}}}} has the same form as for an eikonal wave except h(𝒌)\smash{h({\boldsymbol{k}})} is not necessarily delta-shaped.

7.4.3 Phase-space action density and the wave-kinetic equation

The wave equation for the complexified field 𝚿~c\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}} can be written in the invariant form as 𝚵^|𝚿~c=|𝟎\smash{\widehat{\boldsymbol{\Xi}}\ket{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}}=\ket{{\boldsymbol{0}}}}. Multiplying it by 𝚿~c|\smash{\bra{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}}} from the right leads to

𝚵^𝗨^c=𝟎^,𝗨^c(2\upi)𝗇|𝚿~c𝚿~c|¯.\widehat{\boldsymbol{\Xi}}\widehat{\boldsymbol{{\boldsymbol{\mathsf{U}}}}}_{\text{c}}=\widehat{\boldsymbol{0}},\qquad\widehat{\boldsymbol{{\boldsymbol{\mathsf{U}}}}}_{\text{c}}\doteq(2\upi)^{-{\mathsf{n}}}\,\overline{\ket{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}}\bra{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}}\vphantom{\widetilde{{\boldsymbol{\Psi}}}}}. (356)

This readily yields an equation for the Wigner matrix: 𝚵𝗨c=𝟎\smash{{\boldsymbol{\Xi}}\star{\boldsymbol{\mathsf{U}}}_{\text{c}}={\boldsymbol{0}}}. Let us integrate this equation over ω\smash{\omega} to make the left-hand side a smooth function of (t,𝒙,𝒌)\smash{(t,{\boldsymbol{x}},{\boldsymbol{k}})}. Let us also take the trace of the resulting equation to put it in a scalar form:

trdω𝚵𝗨c=0.\displaystyle\textstyle\operatorname{tr}\int\mathrm{d}\omega\,{\boldsymbol{\Xi}}\star{\boldsymbol{\mathsf{U}}}_{\text{c}}=0. (357)

As usual, we assume 𝚵=𝚵H+i𝚵A{\boldsymbol{\Xi}}={\boldsymbol{\Xi}}_{\text{H}}+\mathrm{i}{\boldsymbol{\Xi}}_{\text{A}} with 𝚵A=𝒪(ϵ)𝚵H=𝒪(1){\boldsymbol{\Xi}}_{\text{A}}=\mathcal{O}(\epsilon)\ll{\boldsymbol{\Xi}}_{\text{H}}=\mathcal{O}(1) for generic (𝘅,𝗸)({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}}). The integrand in (357) can be written as 𝚵𝗨c=𝚵ei^𝗑/2𝗨c\smash{{\boldsymbol{\Xi}}\star{\boldsymbol{\mathsf{U}}}_{\text{c}}={\boldsymbol{\Xi}}\mathrm{e}^{\mathrm{i}\widehat{\mathcal{L}}_{\mathsf{x}}/2}{\boldsymbol{\mathsf{U}}}_{\text{c}}}, and its expansion in the differential operator ^𝗑\smash{\widehat{\mathcal{L}}_{\mathsf{x}}} (32) contains derivatives of all orders. High-order derivatives on 𝗨c\smash{{\boldsymbol{\mathsf{U}}}_{\text{c}}} are not negligible per se, because for on-shell waves this function is delta-shaped. However, using integration by parts, one can reapply all derivatives with respect to ω\smash{\omega} to 𝚵\smash{{\boldsymbol{\Xi}}} and take the remaining derivatives (with respect to t\smash{t}, 𝒙\smash{{\boldsymbol{x}}}, and 𝒌\smash{{\boldsymbol{k}}}) outside the integral. Then it is seen that each power m\smash{m} of ^𝗑\smash{\widehat{\mathcal{L}}_{\mathsf{x}}} in the expansion of 𝚵ei^𝗑/2𝗨c\smash{{\boldsymbol{\Xi}}\mathrm{e}^{\mathrm{i}\widehat{\mathcal{L}}_{\mathsf{x}}/2}{\boldsymbol{\mathsf{U}}}_{\text{c}}} contributes 𝒪(ϵm)\smash{\mathcal{O}(\epsilon^{m})} to the integral. Let us neglect terms with m2\smash{m\geq 2} and use (353). Hence, one obtains343434McDonald & Kaufman (1985) first Taylor-expand 𝚵𝗨c\smash{{\boldsymbol{\Xi}}\star{\boldsymbol{\mathsf{U}}}_{\text{c}}} and then integrate over ω\smash{\omega}. Strictly speaking, that is incorrect (because 𝚵𝗨c\smash{{\boldsymbol{\Xi}}\star{\boldsymbol{\mathsf{U}}}_{\text{c}}} is not smooth), but the final result is the same.

0\displaystyle 0 trdω(𝚵H𝗨c+i𝚵A𝗨c+i2{𝚵H,𝗨c}𝗑)\displaystyle\approx\operatorname{tr}\int\mathrm{d}\omega\,\left({\boldsymbol{\Xi}}_{\text{H}}{\boldsymbol{\mathsf{U}}}_{\text{c}}+\mathrm{i}{\boldsymbol{\Xi}}_{\text{A}}{\boldsymbol{\mathsf{U}}}_{\text{c}}+\frac{\mathrm{i}}{2}\,\{{\boldsymbol{\Xi}}_{\text{H}},{\boldsymbol{\mathsf{U}}}_{\text{c}}\}_{\mathsf{x}}\right)
(𝜼𝚵H𝜼+i𝜼𝚵A𝜼)hc+i2trdω(𝚵H𝗑i𝗨c𝗄i𝚵H𝗄i𝗨c𝗑i).\displaystyle\approx({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}_{\text{H}}{\boldsymbol{\eta}}+\mathrm{i}{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}_{\text{A}}{\boldsymbol{\eta}})h_{\text{c}}+\frac{\mathrm{i}}{2}\,\operatorname{tr}\int\mathrm{d}\omega\left(\frac{\partial{\boldsymbol{\Xi}}_{\text{H}}}{\partial\mathsf{x}^{i}}\frac{\partial{\boldsymbol{\mathsf{U}}}_{\text{c}}}{\partial\mathsf{k}_{i}}-\frac{\partial{\boldsymbol{\Xi}}_{\text{H}}}{\partial\mathsf{k}_{i}}\frac{\partial{\boldsymbol{\mathsf{U}}}_{\text{c}}}{\partial\mathsf{x}^{i}}\right). (358)

Let us also re-express this as follows, using (295) and (338):

0\displaystyle 0 (ΛiγΛω)hci2trdωω(𝚵Ht𝗨c)+i2ttrdω𝚵Hω𝗨c\displaystyle\approx\left(\Lambda-\mathrm{i}\gamma\,\frac{\partial\Lambda}{\partial\omega}\right)h_{\text{c}}-\frac{\mathrm{i}}{2}\operatorname{tr}\int\mathrm{d}\omega\,\frac{\partial}{\partial\omega}\left(\frac{\partial{\boldsymbol{\Xi}}_{\text{H}}}{\partial t}\,{\boldsymbol{\mathsf{U}}}_{\text{c}}\right)+\frac{\mathrm{i}}{2}\frac{\partial}{\partial t}\operatorname{tr}\int\mathrm{d}\omega\,\frac{\partial{\boldsymbol{\Xi}}_{\text{H}}}{\partial\omega}\,{\boldsymbol{\mathsf{U}}}_{\text{c}}
+i2kitrdω𝚵Hxi𝗨ci2xitrdω𝚵Hki𝗨c.\displaystyle\hskip 83.11005pt+\frac{\mathrm{i}}{2}\frac{\partial}{\partial k_{i}}\operatorname{tr}\int\mathrm{d}\omega\,\frac{\partial{\boldsymbol{\Xi}}_{\text{H}}}{\partial x^{i}}\,{\boldsymbol{\mathsf{U}}}_{\text{c}}-\frac{\mathrm{i}}{2}\frac{\partial}{\partial x^{i}}\operatorname{tr}\int\mathrm{d}\omega\,\frac{\partial{\boldsymbol{\Xi}}_{\text{H}}}{\partial k_{i}}\,{\boldsymbol{\mathsf{U}}}_{\text{c}}. (359)

Clearly,

dωω(𝚵Ht𝗨c)=0.\displaystyle\int\mathrm{d}\omega\,\frac{\partial}{\partial\omega}\left(\frac{\partial{\boldsymbol{\Xi}}_{\text{H}}}{\partial t}\,{\boldsymbol{\mathsf{U}}}_{\text{c}}\right)=0. (360)

To simplify the remaining terms, we proceed as follows. As a Hermitian matrix, 𝚵H\smash{{\boldsymbol{\Xi}}_{\text{H}}} can be represented in terms of its eigenvalues Λb\smash{\Lambda_{b}} and eigenvectors 𝜼b\smash{{\boldsymbol{\eta}}_{b}} as 𝚵H=Λb𝜼b𝜼b\smash{{\boldsymbol{\Xi}}_{\text{H}}=\Lambda_{b}{\boldsymbol{\eta}}_{b}{\boldsymbol{\eta}}_{b}^{\dagger}}. For 𝗨c\smash{{\boldsymbol{\mathsf{U}}}_{\text{c}}}, let us use (353) again, where 𝜼\smash{{\boldsymbol{\eta}}} is one of the vectors 𝜼b\smash{{\boldsymbol{\eta}}_{b}}, say, 𝜼𝜼0\smash{{\boldsymbol{\eta}}\equiv{\boldsymbol{\eta}}_{0}}. (Accordingly, ΛΛ0\smash{\Lambda\equiv\Lambda_{0}}.) Then, for any {ω,xi,ki}\smash{\sqbullet\in\{\omega,x^{i},k_{i}\}}, one has

trdω𝚵H𝗨c\displaystyle\operatorname{tr}\int\mathrm{d}\omega\,\frac{\partial{\boldsymbol{\Xi}}_{\text{H}}}{\partial\sqbullet}\,{\boldsymbol{\mathsf{U}}}_{\text{c}} =Λb|𝜼b𝜼|2hc+Λb(𝜼𝜼b)(𝜼b𝜼)hc+Λb(𝜼𝜼b)(𝜼b𝜼)hc\displaystyle=\frac{\partial\Lambda_{b}}{\partial\sqbullet}\,|{\boldsymbol{\eta}}_{b}^{\dagger}{\boldsymbol{\eta}}|^{2}h_{\text{c}}+\Lambda_{b}\,\Big{(}{\boldsymbol{\eta}}^{\dagger}\frac{\partial{\boldsymbol{\eta}}_{b}}{\partial\sqbullet}\Big{)}({\boldsymbol{\eta}}_{b}^{\dagger}{\boldsymbol{\eta}})h_{\text{c}}+\Lambda_{b}\,({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\eta}}_{b})\Big{(}\frac{\partial{\boldsymbol{\eta}}_{b}^{\dagger}}{\partial\sqbullet}\,{\boldsymbol{\eta}}\Big{)}h_{\text{c}}
=Λb(δb,0)2hc+Λb(𝜼𝜼b)δb,0hc+Λbδb,0(𝜼b𝜼)hc\displaystyle=\frac{\partial\Lambda_{b}}{\partial\sqbullet}\,(\delta_{b,0})^{2}h_{\text{c}}+\Lambda_{b}\,\Big{(}{\boldsymbol{\eta}}^{\dagger}\frac{\partial{\boldsymbol{\eta}}_{b}}{\partial\sqbullet}\Big{)}\delta_{b,0}h_{\text{c}}+\Lambda_{b}\,\delta_{b,0}\Big{(}\frac{\partial{\boldsymbol{\eta}}_{b}^{\dagger}}{\partial\sqbullet}\,{\boldsymbol{\eta}}\Big{)}h_{\text{c}}
=Λhc+(𝜼𝜼+𝜼𝜼)Λhc\displaystyle=\frac{\partial\Lambda}{\partial\sqbullet}\,h_{\text{c}}+\Big{(}{\boldsymbol{\eta}}^{\dagger}\frac{\partial{\boldsymbol{\eta}}}{\partial\sqbullet}+\frac{\partial{\boldsymbol{\eta}}^{\dagger}}{\partial\sqbullet}\,{\boldsymbol{\eta}}\Big{)}\,\Lambda h_{\text{c}}
=Λhc+(𝜼𝜼)Λhc\displaystyle=\frac{\partial\Lambda}{\partial\sqbullet}\,h_{\text{c}}+\frac{\partial({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\eta}})}{\partial\sqbullet}\,\Lambda h_{\text{c}}
=Λhc,\displaystyle=\frac{\partial\Lambda}{\partial\sqbullet}\,h_{\text{c}}, (361)

where we used 𝜼b𝜼=δb,0\smash{{\boldsymbol{\eta}}_{b}^{\dagger}{\boldsymbol{\eta}}=\delta_{b,0}} and, in particular, 𝜼𝜼=1\smash{{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\eta}}=1}. Then, (359) can be written as

Λhc2i=0,\displaystyle\Lambda h_{\text{c}}-2\mathrm{i}\mathscr{E}=0, (362)

where

=2γ(Λωh)t(Λωh)ki(Λxih)+xi(Λkih).\displaystyle\mathscr{E}=2\gamma\left(\frac{\partial\Lambda}{\partial\omega}\,h\right)-\frac{\partial}{\partial t}\left(\frac{\partial\Lambda}{\partial\omega}\,h\right)-\frac{\partial}{\partial k_{i}}\left(\frac{\partial\Lambda}{\partial x^{i}}\,h\right)+\frac{\partial}{\partial x^{i}}\left(\frac{\partial\Lambda}{\partial k_{i}}\,h\right). (363)

The real part of (362) gives Λ=0\smash{\Lambda=0}, which is the dispersion relation. The imaginary part of (362) gives =0\smash{\mathscr{E}=0}. To understand this equation, let us rewrite \smash{\mathscr{E}} as

=2γJJt+ki(wxiJ)xi(wkiJ).\displaystyle\mathscr{E}=2\gamma J-\frac{\partial J}{\partial t}+\frac{\partial}{\partial k_{i}}\left(\frac{\partial w}{\partial x^{i}}\,J\right)-\frac{\partial}{\partial x^{i}}\left(\frac{\partial w}{\partial k_{i}}\,J\right). (364)

Here, we introduced

J(𝒌)h(𝒌)ωΛ(𝒌),Λ(𝒌)Λ(w(𝒌),𝒌),\displaystyle J({\boldsymbol{k}})\doteq h({\boldsymbol{k}})\,\partial_{\omega}\Lambda({\boldsymbol{k}}),\qquad\Lambda({\boldsymbol{k}})\doteq\Lambda(w({\boldsymbol{k}}),{\boldsymbol{k}}), (365)

which, according to (315), satisfy

Jtw(𝒌)\displaystyle J\partial_{t}w({\boldsymbol{k}}) =h(tΛ)(𝒌),\displaystyle=-h(\partial_{t}\Lambda)({\boldsymbol{k}}), (366a)
J𝒙w(𝒌)\displaystyle J\partial_{{\boldsymbol{x}}}w({\boldsymbol{k}}) =h(𝒙Λ)(𝒌),\displaystyle=-h(\partial_{{\boldsymbol{x}}}\Lambda)({\boldsymbol{k}}), (366b)
J𝒌w(𝒌)\displaystyle J\partial_{{\boldsymbol{k}}}w({\boldsymbol{k}}) =h(𝒌Λ)(𝒌).\displaystyle=-h(\partial_{{\boldsymbol{k}}}\Lambda)({\boldsymbol{k}}). (366c)

Note that using (355), one can also express J\smash{J} as

J=σ(14|a˘σ|2ωΛ(𝒌σ))δ(𝒌𝒌¯σ)=σσδ(𝒌𝒌¯σ),\displaystyle\textstyle J=\sum_{\sigma}\big{(}\frac{1}{4}\,|{\breve{a}}_{\sigma}|^{2}\partial_{\omega}\Lambda({\boldsymbol{k}}_{\sigma})\big{)}\delta({\boldsymbol{k}}-\overline{{\boldsymbol{k}}}_{\sigma})=\sum_{\sigma}\mathcal{I}_{\sigma}\delta({\boldsymbol{k}}-\overline{{\boldsymbol{k}}}_{\sigma}), (367)

where σ\smash{\mathcal{I}_{\sigma}} are the action densities (318) of the individual eikonal waves that comprise the total wave field (section 7.4.2). In particular, d𝒌J=σσ\smash{\int\mathrm{d}{\boldsymbol{k}}\,J=\sum_{\sigma}\mathcal{I}_{\sigma}}, which is the total action density. Therefore, the function J\smash{J} can be interpreted as the phase-space action density. In terms of J\smash{J}, the equation =0\smash{\mathscr{E}=0} can be written as

Jt+wkiJxiwxiJki=2γJ.\displaystyle\frac{\partial J}{\partial t}+\frac{\partial w}{\partial k_{i}}\,\frac{\partial J}{\partial x^{i}}-\frac{\partial w}{\partial x^{i}}\,\frac{\partial J}{\partial k_{i}}=2\gamma J. (368)

This equation, called the WKE, serves the same role in QL wave-kinetic theory as the Vlasov equation serves in plasma kinetic theory.353535The term ‘WKE’ is also used for the equation that describes nonlinear interactions of waves in statistically homogeneous media, or ‘wave–wave collisions’ (Zakharov et al., 1992). That is not what we consider here. Inhomogeneities are essential in our formulation, and the QL WKE is linear (in J\smash{J}) by definition of the QL approximation. That said, the Weyl symbol calculus that we use can facilitate derivations of wave–wave collision operators as well (Ruiz et al., 2019). Unlike the field equation used in the standard QLT (Drummond & Pines, 1962), (368) exactly conserves the action of nonresonant waves, i.e. those with γ=0\smash{\gamma=0}. Also note that (341) for eikonal waves can be deduced from (368) as a particular case by assuming the ansatz

J(t,𝒙,𝒌)=(t,𝒙)δ(𝒌𝒌¯(t,𝒙))\displaystyle J(t,{\boldsymbol{x}},{\boldsymbol{k}})=\mathcal{I}(t,{\boldsymbol{x}})\delta({\boldsymbol{k}}-\overline{{\boldsymbol{k}}}(t,{\boldsymbol{x}})) (369)

and integrating over 𝒌\smash{{\boldsymbol{k}}}. In other words, eikonal-wave theory can be understood as the ‘cold-fluid’ limit of wave-kinetic theory.

7.4.4 Function 𝖶¯s{\overline{\mathsf{W}}_{s}} in terms of J{J}

Here we explicitly calculate the function (244a) that determines the nonlinear potentials (284). Using (354), one obtains

𝖶¯s(ω,𝒌;𝒑)=|𝜶s𝜼|2(𝒌;𝒑)(h(𝒌)+h(𝒌))δ(ωw(𝒌))0,\displaystyle\overline{\mathsf{W}}_{s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})=|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}({\boldsymbol{k}};{\boldsymbol{p}})\,(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\,\delta(\omega-w({\boldsymbol{k}}))\geq 0, (370)

where (𝒌;𝒑)(w(𝒌),𝒌;𝒑)\smash{({\boldsymbol{k}};{\boldsymbol{p}})\equiv(w({\boldsymbol{k}}),{\boldsymbol{k}};{\boldsymbol{p}})}. By definition of 𝜶^s\smash{\widehat{\boldsymbol{\alpha}}_{s}}, the function t1,𝒙1|𝜶^s|t2,𝒙2\smash{\braket{t_{1},{\boldsymbol{x}}_{1}}{\widehat{\boldsymbol{\alpha}}_{s}}{t_{2},{\boldsymbol{x}}_{2}}} is real for all (t1,𝒙1)\smash{(t_{1},{\boldsymbol{x}}_{1})} and (t2,𝒙2)\smash{(t_{2},{\boldsymbol{x}}_{2})}, so 𝜶s(ω,𝒌)=𝜶s(ω,𝒌)\smash{{\boldsymbol{\alpha}}_{s}(-\omega,-{\boldsymbol{k}})={\boldsymbol{\alpha}}_{s}^{*}(\omega,{\boldsymbol{k}})} by definition of the Weyl symbol (26). Together with (298), this gives |𝜶s𝜼|2(ω,𝒌;𝒑)=|𝜶s𝜼|2(ω,𝒌;𝒑)\smash{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})}=\smash{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}(-\omega,-{\boldsymbol{k}};{\boldsymbol{p}})}, so

|𝜶s𝜼|2|𝜶s𝜼|2(𝒌;𝒑)=|𝜶s𝜼|2(𝒌;𝒑),\displaystyle|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}\equiv|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}({\boldsymbol{k}};{\boldsymbol{p}})=|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}(-{\boldsymbol{k}};{\boldsymbol{p}}), (371a)
and similarly,
|𝜼s𝜼|2|𝜼s𝜼|2(𝒌;𝒑)=|𝜼s𝜼|2(𝒌;𝒑).\displaystyle|{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}}|^{2}\equiv|{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}}|^{2}({\boldsymbol{k}};{\boldsymbol{p}})=|{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}}|^{2}(-{\boldsymbol{k}};{\boldsymbol{p}}). (371b)

This also means that 𝖶¯s(ω,𝒌;𝒑)=𝖶¯s(ω,𝒌;𝒑)\smash{\overline{\mathsf{W}}_{s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})}=\smash{\overline{\mathsf{W}}_{s}(-\omega,-{\boldsymbol{k}};{\boldsymbol{p}})}. Then finally, using (365), one can express this function through the phase-space action density:

𝖶¯s(ω,𝒌;𝒑)=|𝜶s𝜼|2(ς𝒌J(𝒌)+ς𝒌J(𝒌))δ(Λ(ω,𝒌)),\displaystyle\overline{\mathsf{W}}_{s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})=|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}(\varsigma_{{\boldsymbol{k}}}J({\boldsymbol{k}})+\varsigma_{-{\boldsymbol{k}}}J(-{\boldsymbol{k}}))\,\delta(\Lambda(\omega,{\boldsymbol{k}})), (372)
ς𝒌sgnωΛ(𝒌)=sgn(J(𝒌)/h(𝒌))=sgnJ(𝒌).\displaystyle\varsigma_{{\boldsymbol{k}}}\doteq\operatorname{sgn}\partial_{\omega}\Lambda({\boldsymbol{k}})=\operatorname{sgn}(J({\boldsymbol{k}})/h({\boldsymbol{k}}))=\operatorname{sgn}J({\boldsymbol{k}}). (373)

7.5 Conservation laws

Let us rewrite (368) together with (283) in the ‘divergence’ form:

Jt+(vgiJ)xiki(wxiJ)\displaystyle\frac{\partial J}{\partial t}+\frac{\partial(v_{\text{g}}^{i}J)}{\partial x^{i}}-\frac{\partial}{\partial k_{i}}\left(\frac{\partial w}{\partial x^{i}}\,J\right) =2γJ,\displaystyle=2\gamma J, (374)
Fst+(vsiFs)xipi(sxiFs)\displaystyle\frac{\partial F_{s}}{\partial t}+\frac{\partial(v_{s}^{i}\,F_{s})}{\partial x^{i}}-\frac{\partial}{\partial p_{i}}\left(\frac{\partial\mathcal{H}_{s}}{\partial x^{i}}\,F_{s}\right) =pi(𝖣s,ijFspj)+𝒞s.\displaystyle=\frac{\partial}{\partial p_{i}}\left(\mathsf{D}_{s,ij}\,\frac{\partial F_{s}}{\partial p_{j}}\right)+\mathcal{C}_{s}. (375)

Using (353), the diffusion matrix 𝖣s,ij\smash{\mathsf{D}_{s,ij}} can be represented as follows:

𝖣s,ij=2\upid𝒌kikj|𝜶s𝜼|2J(𝒌)ωΛ(𝒌)δ(𝒌𝒗sw(𝒌)).\displaystyle\mathsf{D}_{s,ij}=2\upi\int\mathrm{d}{\boldsymbol{k}}\,k_{i}k_{j}\,|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}\,\frac{J({\boldsymbol{k}})}{\partial_{\omega}\Lambda({\boldsymbol{k}})}\,\delta({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-w({\boldsymbol{k}})). (376)

Also, by substituting (231) into (338), one finds

γ=\upisd𝒑|𝜶s𝜼|2ωΛ(𝒌)δ(w(𝒌)𝒌𝒗s(𝒑))𝒌Fs(𝒑)𝒑.\displaystyle\gamma=\upi\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\frac{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}{\partial_{\omega}\Lambda({\boldsymbol{k}})}\,\delta(w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}({\boldsymbol{p}}))\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}. (377)

Together with (366), these yield the following notable corollaries. First of all, if 𝚵0\smash{{\boldsymbol{\Xi}}_{0}}, |𝜶s𝜼|2\smash{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}, and 𝜼s𝜼\smash{{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}}} are independent of 𝒙\smash{{\boldsymbol{x}}},363636Having 𝒙\smash{{\boldsymbol{x}}}-dependence in 𝚵0\smash{{\boldsymbol{\Xi}}_{0}}, |𝜶s𝜼|2\smash{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}, or 𝜼s𝜼\smash{{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}}} would signify interaction with external fields not treated self-consistently. Such fields could exchange momentum with the wave–plasma system, so the momentum of the latter would not be conserved. A similar argument applies to the temporal dependence of these coefficients vs. energy conservation considered below. one has for each l\smash{l} that (appendix D.1)

t(sd𝒑plFs+d𝒌klJ)\displaystyle\frac{\partial}{\partial t}\left(\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,k_{l}J\right) +xi(sd𝒑plvsiFs+d𝒌klvgiJ)\displaystyle+\frac{\partial}{\partial x^{i}}\left(\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}v_{s}^{i}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,k_{l}v_{\text{g}}^{i}J\right)
+xlsd𝒑ΔsFs=sd𝒑H0sxlFs.\displaystyle+\frac{\partial}{\partial x^{l}}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}F_{s}=-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial H_{0s}}{\partial x^{l}}\,F_{s}. (378)

This can be viewed as a momentum-conservation theorem, because at lH0s=0\smash{\partial_{l}H_{0s}=0}, one has

sd𝒙d𝒑plFs+d𝒙d𝒌klJ=const.\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{x}}\,\mathrm{d}{\boldsymbol{p}}\,p_{l}F_{s}+\int\mathrm{d}{\boldsymbol{x}}\,\mathrm{d}{\boldsymbol{k}}\,k_{l}J=\text{const}. (379)

Also, the ponderomotive force on a plasma is readily found from (378) as the sum of the terms quadratic in the wave amplitude (after Fs\smash{F_{s}} has been expressed through f¯s\smash{\overline{f}_{s}}). Similarly, if 𝚵0\smash{{\boldsymbol{\Xi}}_{0}}, |𝜶s𝜼|2\smash{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}, and 𝜼s𝜼\smash{{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}}} are independent of t\smash{t}, one has (appendix D.2)

t(sd𝒑H0sFs+d𝒌wJ)\displaystyle\frac{\partial}{\partial t}\left(\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,wJ\right) +xi(sd𝒑H0svsiFs+d𝒌wvgiJ)\displaystyle+\frac{\partial}{\partial x^{i}}\left(\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}v_{s}^{i}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,wv_{\text{g}}^{i}J\right)
+xisd𝒑ΔsvsiFs=sd𝒑H0stFs.\displaystyle+\frac{\partial}{\partial x^{i}}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}v_{s}^{i}F_{s}=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial H_{0s}}{\partial t}\,F_{s}. (380)

This can be viewed as an energy-conservation theorem, because at tH0s=0\smash{\partial_{t}H_{0s}=0}, one has

sd𝒙d𝒑H0sFs+d𝒙d𝒌wJ=const.\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{x}}\,\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s}+\int\mathrm{d}{\boldsymbol{x}}\,\mathrm{d}{\boldsymbol{k}}\,wJ=\text{const}. (381)

Related equations are also discussed in (Dodin & Fisch, 2012; Dewar, 1977).

The individual terms in (378) and (380) can be interpreted as described in table 1. The results of section 7.2.3 are reproduced as a particular case for the eikonal-wave ansatz (369).373737There is no ambiguity in the definition of the wave momentum and energy in this case (i.e. κ=1\smash{\kappa=1}), because (379) and (381) connect those with the momentum and energy of particles (OCs), which are defined unambiguously. In particular, note that electrostatic waves carry nonzero momentum density d𝒌𝒌J\smash{\int\mathrm{d}{\boldsymbol{k}}\,{\boldsymbol{k}}J} just like any other waves, even though the electrostatic field of these waves carries no momentum. The momentum is stored in the particle motion in this case (section 9.1.3), and it is pumped there via either temporal dependence (Liu & Dodin, 2015, section II.2) or spatial dependence (Ochs & Fisch, 2021b, 2022) of the wave amplitude. This shows that homogeneous-plasma models that ignore ponderomotive effects cannot adequately describe the energy–momentum transfer between waves and plasma even when resonant absorption per se occurs in a homogeneous-plasma region. The OC formalism presented here provides means to describe such processes rigorously, generally, and without cumbersome calculations.

Quantity Notation Interpretation
d𝒑𝒑Fs\smash{\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{p}}F_{s}} 𝑷s\smash{{\boldsymbol{P}}_{s}} OC momentum density
d𝒑H0sFs\smash{\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s}} s\smash{\mathcal{E}_{s}} OC energy density
d𝒑(𝒑𝒗s+Δs𝟏)Fs\smash{\int\mathrm{d}{\boldsymbol{p}}\,({\boldsymbol{p}}{\boldsymbol{v}}_{s}+\Delta_{s}{\boldsymbol{1}})F_{s}} 𝚷s\smash{{\boldsymbol{\Pi}}_{s}} OC momentum flux density
d𝒑(H0s+Δs)𝒗sFs\smash{\int\mathrm{d}{\boldsymbol{p}}\,(H_{0s}+\Delta_{s}){\boldsymbol{v}}_{s}F_{s}} 𝑸s\smash{{\boldsymbol{Q}}_{s}} OC energy flux density
d𝒌𝒌J\smash{\int\mathrm{d}{\boldsymbol{k}}\,{\boldsymbol{k}}J} 𝑷w\smash{{\boldsymbol{P}}_{\text{w}}} wave momentum density
d𝒌wJ\smash{\int\mathrm{d}{\boldsymbol{k}}\,wJ} w\smash{\mathcal{E}_{\text{w}}} wave energy density
d𝒌𝒌𝒗gJ\smash{\int\mathrm{d}{\boldsymbol{k}}\,{\boldsymbol{k}}{\boldsymbol{v}}_{\text{g}}J} 𝚷w\smash{{\boldsymbol{\Pi}}_{\text{w}}} wave momentum flux density
d𝒌w𝒗gJ\smash{\int\mathrm{d}{\boldsymbol{k}}\,w{\boldsymbol{v}}_{\text{g}}J} 𝑸w\smash{{\boldsymbol{Q}}_{\text{w}}} wave energy flux density
Table 1: Interpretation of the individual terms in (378) and (380). The wave energy–momentum is understood as the canonical (‘Minkowski’) energy–momentum, which must not be confused with the kinetic (‘Abraham’) energy–momentum (Dodin & Fisch, 2012; Dewar, 1977). Whether the terms with ΔsFs\smash{\Delta_{s}F_{s}} should be attributed to OCs or to the wave is a matter of convention, because ΔsFs\smash{\Delta_{s}F_{s}} scales linearly both with Fs\smash{F_{s}} and with J\smash{J}. In contrast, the wave energy density is defined unambiguously as sd𝒑H0sFs\smash{\mathcal{E}_{s}\doteq\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s}} and does not contain Δs\smash{\Delta_{s}}. This is because d𝒑ΔsFs\smash{\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}F_{s}} is a part of the wave energy density w\smash{\mathcal{E}_{\text{w}}} (Dodin & Fisch, 2010a). Similarly, d𝒑(𝒗sΔs)Fs\smash{\int\mathrm{d}{\boldsymbol{p}}\,(\partial_{{\boldsymbol{v}}_{s}}\Delta_{s})F_{s}} is a part of the wave momentum density (Dodin & Fisch, 2012).

7.6 Summary of section 7

In summary, we have considered plasma interaction with general broadband single-mode on-shell waves (for examples, see section 9). Assuming a general response matrix 𝚵\smash{{\boldsymbol{\Xi}}}, these waves have a dispersion function Λ(t,𝒙,ω,𝒌)\smash{\Lambda(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}})} and polarization 𝜼(t,𝒙,ω,𝒌){\boldsymbol{\eta}}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}}) determined by

𝚵H𝜼=Λ𝜼,Λ=𝜼𝚵H𝜼,\displaystyle{\boldsymbol{\Xi}}_{\text{H}}{\boldsymbol{\eta}}=\Lambda{\boldsymbol{\eta}},\qquad\Lambda={\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}_{\text{H}}{\boldsymbol{\eta}}, (382)

where the normalization 𝜼𝜼=1\smash{{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\eta}}=1} is assumed. Specifically for 𝚵\smash{{\boldsymbol{\Xi}}} given by (287), one has

Λ(t,𝒙,ω,𝒌)=𝜼𝚵0𝜼\displaystyle\Lambda(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}})={\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}_{0}{\boldsymbol{\eta}} sd𝒑𝜼s(𝒑)𝜼Fs(𝒑)\displaystyle-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}({\boldsymbol{p}}){\boldsymbol{\eta}}\,F_{s}({\boldsymbol{p}})
+sd𝒑|𝜶s𝜼|2(𝒑)ω𝒌𝒗s(𝒑)𝒌Fs(𝒑)𝒑,\displaystyle+\sum_{s}\fint\mathrm{d}{\boldsymbol{p}}\,\frac{|{\boldsymbol{\alpha}}_{s}^{{\dagger}}{\boldsymbol{\eta}}|^{2}({\boldsymbol{p}})}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}({\boldsymbol{p}})}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}, (383)

where the arguments (t,𝒙,ω,𝒌)\smash{(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}})} are omitted for brevity. (Some notation is summarized in section 6.9.) The wave frequency ω=w(t,𝒙,𝒌)\smash{\omega=w(t,{\boldsymbol{x}},{\boldsymbol{k}})} satisfies

Λ(t,𝒙,w(t,𝒙,𝒌),𝒌)=0\displaystyle\Lambda(t,{\boldsymbol{x}},w(t,{\boldsymbol{x}},{\boldsymbol{k}}),{\boldsymbol{k}})=0 (384)

and w(t,𝒙,𝒌)=w(t,𝒙,𝒌)\smash{w(t,{\boldsymbol{x}},-{\boldsymbol{k}})=-w(t,{\boldsymbol{x}},{\boldsymbol{k}})}, where ww is a real function at real arguments. The wave local linear growth rate γ\smash{\gamma}, which is assumed to be small in this section, is

γ(t,𝒙,𝒌)=(𝜼𝚵A𝜼ωΛ)(t,𝒙,w(t,𝒙,𝒌),𝒌),\displaystyle\gamma(t,{\boldsymbol{x}},{\boldsymbol{k}})=-\left(\frac{{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}_{\text{A}}{\boldsymbol{\eta}}}{\partial_{\omega}\Lambda}\right)_{(t,{\boldsymbol{x}},w(t,{\boldsymbol{x}},{\boldsymbol{k}}),{\boldsymbol{k}})}, (385)

or explicitly,

γ(t,𝒙,𝒌)=\upisd𝒑|𝜶s𝜼|2ωΛ(t,𝒙,w,𝒌)δ(w𝒌𝒗s(t,𝒙,𝒑))𝒌Fs(t,𝒙,𝒑)𝒑,\displaystyle\gamma(t,{\boldsymbol{x}},{\boldsymbol{k}})=\upi\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\frac{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}{\partial_{\omega}\Lambda(t,{\boldsymbol{x}},w,{\boldsymbol{k}})}\,\delta(w-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}}))\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}})}{\partial{\boldsymbol{p}}},

where ww(t,𝒙,𝒌)w\equiv w(t,{\boldsymbol{x}},{\boldsymbol{k}}) and |𝜶s𝜼|2|𝜶s𝜼|2(t,𝒙,w,𝒌;𝒑)\smash{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}\equiv|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}(t,{\boldsymbol{x}},w,{\boldsymbol{k}};{\boldsymbol{p}})}. The nonlinear potentials (284) are expressed through the scalar function

𝖶¯s(t,𝒙,ω,𝒌;𝒑)=|𝜶s𝜼|2(ς𝒌J(t,𝒙,𝒌)+ς𝒌J(t,𝒙,𝒌))δ(Λ(t,𝒙,ω,𝒌)),\displaystyle\overline{\mathsf{W}}_{s}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}})=|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}(\varsigma_{{\boldsymbol{k}}}J(t,{\boldsymbol{x}},{\boldsymbol{k}})+\varsigma_{-{\boldsymbol{k}}}J(t,{\boldsymbol{x}},-{\boldsymbol{k}}))\,\delta(\Lambda(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}})), (386)

where ς𝒌sgn(ωΛ(t,𝒙,ω,𝒌))\smash{\varsigma_{{\boldsymbol{k}}}\doteq\operatorname{sgn}(\partial_{\omega}\Lambda(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}}))} is evaluated at ω=w(t,𝒙,𝒌)\omega=w(t,{\boldsymbol{x}},{\boldsymbol{k}}); see also (373). The function JJ is the phase-space action density governed by the WKE:

Jtw𝒙J𝒌+w𝒌J𝒙=2γJ,\displaystyle\frac{\partial J}{\partial t}-\frac{\partial w}{\partial{\boldsymbol{x}}}\cdot\frac{\partial J}{\partial{\boldsymbol{k}}}+\frac{\partial w}{\partial{\boldsymbol{k}}}\cdot\frac{\partial J}{\partial{\boldsymbol{x}}}=2\gamma J, (387)

where 𝒌w=𝒗g\smash{\partial_{\boldsymbol{k}}w={\boldsymbol{v}}_{\text{g}}} is the group velocity. Collisional dissipation is assumed small compared to collisionless dissipation, so it is neglected in (387) but can be reintroduced by an ad hoc modification of γ\smash{\gamma} (section 6.2). Unlike the field equation used in the standard QLT, (387) exactly conserves the action of nonresonant waves, i.e. those with γ=0\smash{\gamma=0}. The WKE must be solved together with the QL equation for the OC distribution FsF_{s},

Fsts𝒙Fs𝒑+s𝒑Fs𝒙=𝒑(𝗗sFs𝒑)+𝒞s,\displaystyle\frac{\partial F_{s}}{\partial t}-\frac{\partial\mathcal{H}_{s}}{\partial{\boldsymbol{x}}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}+\frac{\partial\mathcal{H}_{s}}{\partial{\boldsymbol{p}}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{x}}}=\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left({\boldsymbol{\mathsf{D}}}_{s}\,\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}\right)+\mathcal{C}_{s}, (388)

because FsF_{s} determines the coefficients in (387) and J\smash{J} determines the coefficients in (388). When 𝚵0\smash{{\boldsymbol{\Xi}}_{0}} and |𝜶s𝜼|2\smash{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}} are independent of t\smash{t} and 𝒙\smash{{\boldsymbol{x}}}, (387) and (388) conserve the total momentum and energy of the system; specifically,

t(sPs,i+Pw,i)+j(sΠs,ij+Πw,ij)\displaystyle\textstyle\partial_{t}(\sum_{s}P_{s,i}+P_{\text{w},i})+\partial_{j}(\sum_{s}{\Pi_{s,i}}^{j}+{\Pi_{\text{w},i}}^{j}) =sd𝒑FsiH0s,\displaystyle=-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,F_{s}\partial_{i}H_{0s}, (389)
t(ss+w)+j(sQsj+Qwj)\displaystyle\textstyle\partial_{t}(\sum_{s}\mathcal{E}_{s}+\mathcal{E}_{\text{w}})+\partial_{j}(\sum_{s}Q^{j}_{s}+Q^{j}_{\text{w}}) =sd𝒑FstH0s.\displaystyle=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,F_{s}\partial_{t}H_{0s}. (390)

Here, the notation is as in table 1, or see (378) and (380) instead.

8 Thermal equilibrium

In this section, we discuss, for completeness, the properties of plasmas in thermal equilibrium.

8.1 Boltzmann–Gibbs distribution

As discussed in section 6.8, collisions conserve the density of each species, the total momentum density, and the total energy density, while the plasma total entropy density σ\smash{\sigma} either grows or remains constant. Let us search for equilibrium states in particular. At least one of the states in which σ\smash{\sigma} remains constant is the one that maximizes the entropy density at fixed d𝒑Fs\smash{\int\mathrm{d}{\boldsymbol{p}}\,F_{s}}, sd𝒑𝒑Fs\smash{\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{p}}F_{s}}, and sd𝒑sFs\smash{\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}F_{s}}. This ‘state of thermal equilibrium’ can be found as an extremizer of

σσsλs(𝒩)d𝒑Fs𝝀(𝑷)sd𝒑𝒑Fsλ()sd𝒑sFs\displaystyle\sigma^{\prime}\doteq\sigma-\sum_{s}\lambda_{s}^{(\mathcal{N})}\int\mathrm{d}{\boldsymbol{p}}\,F_{s}-{\boldsymbol{\lambda}}^{({\boldsymbol{P}})}\cdot\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{p}}F_{s}-\lambda^{(\mathcal{H})}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}F_{s} (391)

considered as a functional of all Fs\smash{F_{s}}, where λs(𝒩)\smash{\lambda_{s}^{(\mathcal{N})}}, 𝝀(𝑷)\smash{{\boldsymbol{\lambda}}^{({\boldsymbol{P}})}}, and λ()\smash{\lambda^{(\mathcal{H})}} are Lagrange multipliers. Using (280), one finds that extremizers of σ\smash{\sigma^{\prime}} satisfy

0=δσδFs=lnFs1λs(𝒩)𝝀(𝑷)𝒑λ()s,\displaystyle 0=\frac{\delta\sigma^{\prime}}{\delta F_{s}}=-\ln F_{s}-1-\lambda_{s}^{(\mathcal{N})}-{\boldsymbol{\lambda}}^{({\boldsymbol{P}})}\cdot{\boldsymbol{p}}-\lambda^{(\mathcal{H})}\mathcal{H}_{s}, (392)

whence

Fs=consts×exp(𝝀(𝑷)𝒑λ()s).\displaystyle F_{s}=\text{const}_{s}\times\exp(-{\boldsymbol{\lambda}}^{({\boldsymbol{P}})}\cdot{\boldsymbol{p}}-\lambda^{(\mathcal{H})}\mathcal{H}_{s}). (393)

The pre-exponential constant is determined by the given density of species s\smash{s}, while 𝝀(𝑷)\smash{{\boldsymbol{\lambda}}^{({\boldsymbol{P}})}} and λ()\smash{\lambda^{(\mathcal{H})}} can be expressed through the densities of the plasma momentum and energy stored in the whole distribution. Because

δ2σδFsδFs=1Fs<0,δ2σδFsδFss=0,\displaystyle\frac{\delta^{2}\sigma^{\prime}}{\delta F_{s}\delta F_{s}}=-\frac{1}{F_{s}}<0,\qquad\frac{\delta^{2}\sigma^{\prime}}{\delta F_{s}\delta F_{s^{\prime}\neq s}}=0, (394)

the matrix δ2σ/δFsδFs\smash{\delta^{2}\sigma^{\prime}/\delta F_{s}\delta F_{s^{\prime}}} is positive-definite, so the entropy is maximal (as opposed to minimal) at the extremizer (393).

The distribution (393) is known as the Boltzmann–Gibbs distribution, with T1/λ()\smash{T\doteq 1/\lambda^{(\mathcal{H})}} being the temperature (common for all species). Also, let us introduce a new, rescaled Lagrange multiplier 𝖚\smash{{\boldsymbol{\mathfrak{u}}}} via 𝝀(𝑷)=𝖚/T\smash{{\boldsymbol{\lambda}}^{({\boldsymbol{P}})}=-{\boldsymbol{\mathfrak{u}}}/T}. Then,

Fs(𝒑)=Fs(0)exp(s(𝒑)𝖚𝒑T),\displaystyle F_{s}({\boldsymbol{p}})=F_{s}^{(0)}\exp\left(-\frac{\mathcal{H}_{s}({\boldsymbol{p}})-{\boldsymbol{\mathfrak{u}}}\cdot{\boldsymbol{p}}}{T}\right), (395)

where Fs(0)\smash{F_{s}^{(0)}} is independent of 𝒑\smash{{\boldsymbol{p}}}. Correspondingly,

Fs(𝒑)𝒑=Fs(𝒑)𝒗s𝖚T,\displaystyle\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}=-F_{s}({\boldsymbol{p}})\,\frac{{\boldsymbol{v}}_{s}-{\boldsymbol{\mathfrak{u}}}}{T}, (396)

where we used (204). From (396), one obtains

δ(\displaystyle\delta( 𝒌𝒗s𝒌𝒗s)𝒌(Fs(𝒑)𝒑Fs(𝒑)Fs(𝒑)Fs(𝒑)𝒑)\displaystyle{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,{\boldsymbol{k}}\cdot\left(\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}\,F_{s^{\prime}}({\boldsymbol{p}}^{\prime})-F_{s}({\boldsymbol{p}})\,\frac{\partial F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial{\boldsymbol{p}}^{\prime}}\right)
=1Tδ(𝒌𝒗s𝒌𝒗s)(𝒌𝒗s𝒌𝒗s)Fs(𝒑)Fs(𝒑)\displaystyle=-\frac{1}{T}\,\delta({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,F_{s}({\boldsymbol{p}})F_{s^{\prime}}({\boldsymbol{p}}^{\prime})
=0,\displaystyle=0, (397)

where ss(𝒑)\smash{\mathcal{H}^{\prime}_{s^{\prime}}\doteq\mathcal{H}_{s^{\prime}}({\boldsymbol{p}}^{\prime})}. Then, (6.8) yields that the collision operator vanishes on the Boltzmann–Gibbs distribution, and thus, expectedly, (dσ/dt)coll=0\smash{(\mathrm{d}\sigma/\mathrm{d}t)_{\text{coll}}=0}. One can also show that the Boltzmann–Gibbs distribution is the only distribution (strictly speaking, a class of distributions parameterized by T\smash{T} and 𝖚\smash{{\boldsymbol{\mathfrak{u}}}}) for which the entropy density is conserved (appendix E).

The property (396) of the thermal-equilibrium state also leads to other notable results that we derive below. In doing so, we will assume the reference frame where 𝖚=0\smash{{\boldsymbol{\mathfrak{u}}}=0}, so the Boltzmann–Gibbs distribution has a better known form

Fs(𝒑)=Fs(0)exp(s(𝒑)T),Fs(𝒑)𝒑=Fs(𝒑)𝒗sT.\displaystyle F_{s}({\boldsymbol{p}})=F_{s}^{(0)}\exp\left(-\frac{\mathcal{H}_{s}({\boldsymbol{p}})}{T}\right),\qquad\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}=-F_{s}({\boldsymbol{p}})\,\frac{{\boldsymbol{v}}_{s}}{T}. (398)

(For s\smash{\mathcal{H}_{s}} isotropic in 𝒑\smash{{\boldsymbol{p}}}, this is the frame where the plasma total momentum density sd𝒑𝒑Fs\smash{\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{p}}F_{s}} is zero.) The generalizations to arbitrary 𝖚\smash{{\boldsymbol{\mathfrak{u}}}} are straightforward.

8.2 Fluctuation–dissipation theorem

Let us describe microscopic fluctuations in equilibrium plasmas in terms of 𝗦(ω,𝒌)(2\upi)𝗇𝖂(ω,𝒌)\smash{{\boldsymbol{\mathsf{S}}}(\omega,{\boldsymbol{k}})\doteq(2\upi)^{{\mathsf{n}}}\,{\boldsymbol{\mathfrak{W}}}(\omega,{\boldsymbol{k}})}, i.e.

𝗦(ω,𝒌)dτd𝒔 ~Ψ ~  (t+τ/2,𝒙+𝒔/2) ~Ψ ~  (tτ/2,𝒙𝒔/2)¯eiωτi𝒌𝒔,{\boldsymbol{\mathsf{S}}}(\omega,{\boldsymbol{k}})\doteq\int\mathrm{d}\tau\mathrm{d}{\boldsymbol{s}}\,\overline{\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}(t+\tau/2,{\boldsymbol{x}}+{\boldsymbol{s}}/2)\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}^{\dagger}(t-\tau/2,{\boldsymbol{x}}-{\boldsymbol{s}}/2)}\,\mathrm{e}^{\mathrm{i}\omega\tau-\mathrm{i}{\boldsymbol{k}}\cdot{\boldsymbol{s}}}, (399)

which can also be represented in terms of the Fourier image 𝚿̊(ω,𝒌)\smash{\mathring{{\boldsymbol{\Psi}}}(\omega,{\boldsymbol{k}})} of the microscopic field ~Ψ ~  (t,𝒙)\smash{\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}(t,{\boldsymbol{x}})}:

𝗦(ω,𝒌)=𝚿̊(ω,𝒌)𝚿̊(ω,𝒌)¯𝒯𝒱n.{\boldsymbol{\mathsf{S}}}(\omega,{\boldsymbol{k}})=\frac{\overline{\mathring{{\boldsymbol{\Psi}}}(\omega,{\boldsymbol{k}})\smash{\mathring{{\boldsymbol{\Psi}}}}^{\dagger}(\omega,{\boldsymbol{k}})}}{\mathscr{T}\mathscr{V}_{n}}. (400)

For statistically homogeneous fields that persist on time 𝒯\smash{\mathscr{T}\to\infty} within volume 𝒱n\smash{\mathscr{V}_{n}\to\infty}, the Fourier transform is formally divergent; hence the appearance of the factors 𝒯\smash{\mathscr{T}} and 𝒱n\smash{\mathscr{V}_{n}} in (400).383838To make (400) look more physical (local), one can absorb the global factors 𝒯\smash{\mathscr{T}} and 𝒱n\smash{\mathscr{V}_{n}} in the definition of the Fourier transform; cf. section 9.1.5. Also, as seen from (239), any quadratic function of the microscopic field can be expressed through 𝗦\smash{{\boldsymbol{\mathsf{S}}}} via

(𝗟^𝝍)(𝗥^𝝍)¯dω2\upid𝒌(2\upi)n(𝗟𝗦𝗥)(ω,𝒌),\displaystyle\overline{(\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}{\boldsymbol{\psi}})(\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}{\boldsymbol{\psi}})^{\dagger}}\approx\int\frac{\mathrm{d}\omega}{2\upi}\,\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,({\boldsymbol{\mathsf{L}}}{\boldsymbol{\mathsf{S}}}{\boldsymbol{\mathsf{R}}}^{\dagger})(\omega,{\boldsymbol{k}}), (401)

where 𝗟^\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}} and 𝗥^\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}} are any linear operators and 𝗟\smash{{\boldsymbol{\mathsf{L}}}} and 𝗥\smash{{\boldsymbol{\mathsf{R}}}} are their symbols.

From (238), one finds that, in general,

𝗦(ω,𝒌)=2\upisd𝒑δ(ω𝒌𝒗s)Fs(𝒑)𝚵1(ω,𝒌)(𝜶s𝜶s)(ω,𝒌;𝒑)𝚵(ω,𝒌).{\boldsymbol{\mathsf{S}}}(\omega,{\boldsymbol{k}})=2\upi\,\sum_{s^{\prime}}\int\mathrm{d}{\boldsymbol{p}}^{\prime}\,\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})F_{s^{\prime}}({\boldsymbol{p}}^{\prime}){\boldsymbol{\Xi}}^{-1}(\omega,{\boldsymbol{k}})({\boldsymbol{\alpha}}_{s^{\prime}}{\boldsymbol{\alpha}}_{s^{\prime}}^{\dagger})(\omega,{\boldsymbol{k}};{\boldsymbol{p}}^{\prime}){\boldsymbol{\Xi}}^{-{\dagger}}(\omega,{\boldsymbol{k}}). (402)

For a thermal distribution in particular, which satisfies (398), one can rewrite (231) as follows:

𝚵A(ω,𝒌)\displaystyle{\boldsymbol{\Xi}}_{\text{A}}(\omega,{\boldsymbol{k}}) \upiTsd𝒑𝜶s(ω,𝒌;𝒑)𝜶s(ω,𝒌;𝒑)δ(ω𝒌𝒗s)(𝒌𝒗s)Fs(𝒑)\displaystyle\approx\frac{\upi}{T}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{\alpha}}_{s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}){\boldsymbol{\alpha}}_{s}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})\,({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})F_{s}({\boldsymbol{p}})
=\upiωTsd𝒑𝜶s(ω,𝒌;𝒑)𝜶s(ω,𝒌;𝒑)δ(ω𝒌𝒗s)Fs(𝒑).\displaystyle=\frac{\upi\omega}{T}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{\alpha}}_{s}(\omega,{\boldsymbol{k}};{\boldsymbol{p}}){\boldsymbol{\alpha}}_{s}^{\dagger}(\omega,{\boldsymbol{k}};{\boldsymbol{p}})\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})F_{s}({\boldsymbol{p}}). (403)

By comparing this with (402), one also finds that

𝗦(ω,𝒌)=2Tω(𝚵1𝚵A𝚵)(ω,𝒌).\displaystyle{\boldsymbol{\mathsf{S}}}(\omega,{\boldsymbol{k}})=\frac{2T}{\omega}\,({\boldsymbol{\Xi}}^{-1}{\boldsymbol{\Xi}}_{\text{A}}{\boldsymbol{\Xi}}^{-{\dagger}})(\omega,{\boldsymbol{k}}). (404)

Due to (233), this leads to the fluctuation–dissipation theorem in the following form:

𝗦(ω,𝒌)=2Tω(𝚵1)A(ω,𝒌).\displaystyle{\boldsymbol{\mathsf{S}}}(\omega,{\boldsymbol{k}})=-\frac{2T}{\omega}\,({\boldsymbol{\Xi}}^{-1})_{\text{A}}(\omega,{\boldsymbol{k}}). (405)

For examples of 𝚵\smash{{\boldsymbol{\Xi}}} for specific systems, see section 9.

8.3 Kirchhoff’s law

Consider the power deposition via polarization drag:

𝔓=sd𝒑(𝒗s𝕱s)Fs(𝒑).\displaystyle\mathfrak{P}=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,({\boldsymbol{v}}_{s}\cdot{\boldsymbol{\mathfrak{F}}}_{s})F_{s}({\boldsymbol{p}}). (406)

Using (265a) for 𝕱s\smash{{\boldsymbol{\mathfrak{F}}}_{s}}, (403) for 𝚵A\smash{{\boldsymbol{\Xi}}_{\text{A}}}, and (405) for 𝗦\smash{{\boldsymbol{\mathsf{S}}}}, this can also be expressed as follows:

𝔓\displaystyle\mathfrak{P} sd𝒌(2\upi)nd𝒑(𝒌𝒗s)(𝜶s(𝚵1)A𝜶s)(𝒌𝒗s,𝒌;𝒑)Fs(𝒑)\displaystyle\approx\sum_{s}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}\,({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})\,({\boldsymbol{\alpha}}_{s}^{\dagger}({\boldsymbol{\Xi}}^{-1})_{\text{A}}{\boldsymbol{\alpha}}_{s})({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}})F_{s}({\boldsymbol{p}})
=sdω2\upid𝒌(2\upi)nd𝒑 2\upiωδ(ω𝒌𝒗s)(𝜶s(𝚵1)A𝜶s)(ω,𝒌;𝒑)Fs(𝒑)\displaystyle=\sum_{s}\int\frac{\mathrm{d}\omega}{2\upi}\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}\,2\upi\omega\,\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})({\boldsymbol{\alpha}}_{s}^{\dagger}({\boldsymbol{\Xi}}^{-1})_{\text{A}}{\boldsymbol{\alpha}}_{s})(\omega,{\boldsymbol{k}};{\boldsymbol{p}})F_{s}({\boldsymbol{p}})
=2Tdω2\upid𝒌(2\upi)ntr((𝚵1)A\upiωTsd𝒑δ(ω𝒌𝒗s)(𝜶s𝜶s)(ω,𝒌;𝒑)Fs(𝒑))\displaystyle=2T\int\frac{\mathrm{d}\omega}{2\upi}\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\operatorname{tr}\bigg{(}({\boldsymbol{\Xi}}^{-1})_{\text{A}}\frac{\upi\omega}{T}\,\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})({\boldsymbol{\alpha}}_{s}{\boldsymbol{\alpha}}_{s}^{\dagger})(\omega,{\boldsymbol{k}};{\boldsymbol{p}})F_{s}({\boldsymbol{p}})\bigg{)}
=dω2\upid𝒌(2\upi)nωtr(𝗦𝚵A)(ω,𝒌).\displaystyle=-\int\frac{\mathrm{d}\omega}{2\upi}\,\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\omega\operatorname{tr}({\boldsymbol{\mathsf{S}}}{\boldsymbol{\Xi}}_{\text{A}})(\omega,{\boldsymbol{k}}). (407)

Thus, the spectral density of the power deposition via polarization drag is given by

𝔓ω,𝒌=ωtr(𝗦𝚵A),\displaystyle\mathfrak{P}_{\omega,{\boldsymbol{k}}}=-\omega\operatorname{tr}({\boldsymbol{\mathsf{S}}}{\boldsymbol{\Xi}}_{\text{A}}), (408)

which is a restatement of Kirchhoff’s law (Krall & Trivelpiece, 1973, section 11.4). For examples of 𝚵\smash{{\boldsymbol{\Xi}}} for specific systems, see section 9.

8.4 Equipartition theorem

As flows from section 7.5, the energy of on-shell waves of a field 𝚿~\smash{{\boldsymbol{\widetilde{\Psi}}}} in a homogeneous n\smash{n}-dimensional plasma of a given volume 𝒱n\smash{\mathscr{V}_{n}} can be written as

𝒱nw\displaystyle\mathscr{V}_{n}\mathcal{E}_{\text{w}} =d𝒌𝒱nw(𝒌)J(𝒌)\displaystyle=\int\mathrm{d}{\boldsymbol{k}}\,\mathscr{V}_{n}w({\boldsymbol{k}})J({\boldsymbol{k}})
=(2\upi)n𝒱nd𝒌(2\upi)n0dωωωΛ(ω,𝒌)h(𝒌)δ(ωw(𝒌))\displaystyle=(2\upi)^{n}\int\frac{\mathscr{V}_{n}\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\int_{0}^{\infty}\mathrm{d}\omega\,\omega\,\partial_{\omega}\Lambda(\omega,{\boldsymbol{k}})\,h({\boldsymbol{k}})\,\delta(\omega-w({\boldsymbol{k}}))
=(2\upi)n𝒌0dωωωΛ(ω,𝒌)(𝜼𝗨𝜼)(ω,𝒌).\displaystyle=(2\upi)^{n}\sum_{{\boldsymbol{k}}}\int^{\infty}_{0}\mathrm{d}\omega\,\omega\,\partial_{\omega}\Lambda(\omega,{\boldsymbol{k}})\,({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\mathsf{U}}}{\boldsymbol{\eta}})(\omega,{\boldsymbol{k}}). (409)

To apply this to microscopic fluctuations, one can replace 𝗨\smash{{\boldsymbol{\mathsf{U}}}} with 𝖂\smash{{\boldsymbol{\mathfrak{W}}}} and substitute 𝖂=(2\upi)(n+1)𝗦\smash{{\boldsymbol{\mathfrak{W}}}=(2\upi)^{-(n+1)}{\boldsymbol{\mathsf{S}}}}. Then, the total energy of a mode with given wavevector 𝒌\smash{{\boldsymbol{k}}} and polarization 𝜼\smash{{\boldsymbol{\eta}}} can be expressed as

𝒌,𝜼=12\upi0dωω(ωΛ)𝜼𝗦𝜼,\displaystyle\mathcal{E}_{{\boldsymbol{k}},{\boldsymbol{\eta}}}=\frac{1}{2\upi}\int^{\infty}_{0}\mathrm{d}\omega\,\omega\,(\partial_{\omega}\Lambda)\,{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\mathsf{S}}}{\boldsymbol{\eta}}, (410)

where the arguments (ω,𝒌)\smash{(\omega,{\boldsymbol{k}})} are omitted for brevity. For thermal equilibrium, one can substitute (405) for 𝗦\smash{{\boldsymbol{\mathsf{S}}}}; then,

𝒌,𝜼=T\upiim0dω(ωΛ)𝜼𝚵1𝜼.\displaystyle\mathcal{E}_{{\boldsymbol{k}},{\boldsymbol{\eta}}}=-\frac{T}{\upi}\,\operatorname{im}\int^{\infty}_{0}\mathrm{d}\omega\,(\partial_{\omega}\Lambda)\,{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}^{-1}{\boldsymbol{\eta}}. (411)

The integrand peaks at ω=w(𝒌)\smash{\omega=w({\boldsymbol{k}})}, where the mode eigenvalue Λ\smash{\Lambda} is small. Due to damping, the actual zero of Λ\smash{\Lambda} is slightly below the real axis in the complex-frequency space. Then, at infinitesimally small damping, 𝜼𝚵1𝜼\smash{{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}^{-1}{\boldsymbol{\eta}}} can be approximated near ω=w(𝒌)\smash{\omega=w({\boldsymbol{k}})} as

𝜼𝚵1𝜼1Λ1ωΛ(ω,𝒌)(pv1ωw(𝒌)i\upiδ(ωw(𝒌))).\displaystyle{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}^{-1}{\boldsymbol{\eta}}\approx\frac{1}{\Lambda}\approx\frac{1}{\partial_{\omega}\Lambda(\omega,{\boldsymbol{k}})}\left(\operatorname{pv}\frac{1}{\omega-w({\boldsymbol{k}})}-\mathrm{i}\upi\delta(\omega-w({\boldsymbol{k}}))\right). (412)

This leads to the well-known equipartition theorem:

𝒌=T.\displaystyle\mathcal{E}_{{\boldsymbol{k}}}=T. (413)

Note that according to (413), the sum 𝒱nw=𝒌,𝜼𝒌,𝜼\smash{\mathscr{V}_{n}\mathcal{E}_{\text{w}}=\sum_{{\boldsymbol{k}},{\boldsymbol{\eta}}}\mathcal{E}_{{\boldsymbol{k}},{\boldsymbol{\eta}}}} is divergent. This indicates that not all modes can be classical and on-shell (weakly damped) simultaneously.

8.5 Summary of section 8

In thermal equilibrium, when all species have Boltzmann–Gibbs distributions with common temperature T\smash{T}, the collision operator vanishes, the entropy is conserved, and the spectrum of microscopic fluctuations (399) satisfies the fluctuation–dissipation theorem:

𝗦(ω,𝒌)=2Tω(𝚵1)A(ω,𝒌),\displaystyle{\boldsymbol{\mathsf{S}}}(\omega,{\boldsymbol{k}})=-\frac{2T}{\omega}\,({\boldsymbol{\Xi}}^{-1})_{\text{A}}(\omega,{\boldsymbol{k}}), (414)

where 𝚵\smash{{\boldsymbol{\Xi}}} is the dispersion matrix (287) and A\smash{{}_{\text{A}}} denotes the anti-Hermitian part (or the imaginary part in case of scalar fields). From here, it is shown that the spectral density of the power deposition via polarization drag is given by 𝔓ω,𝒌=ωtr(𝗦𝚵A)\smash{\mathfrak{P}_{\omega,{\boldsymbol{k}}}=-\omega\operatorname{tr}({\boldsymbol{\mathsf{S}}}{\boldsymbol{\Xi}}_{\text{A}})}, which is a restatement of Kirchhoff’s law. For on-shell waves, (414) reduces to the equipartition theorem, which says that the energy per mode equals T\smash{T}. Applications to specific systems are discussed in section 9.

9 Examples

In this section, we show how to apply our general formulation to nonrelativistic electrostatic interactions (section 9.1), relativistic electromagnetic interactions (section 9.2), Newtonian gravity (section 9.3), and relativistic gravity, including gravitational waves (section 9.4).

9.1 Nonrelativistic electrostatic interactions

9.1.1 Main equations

Let us show how our general formulation reproduces (and generalizes) the well-known results for electrostatic turbulence in nonmagnetized nonrelativistic plasma. In this case,

Hs=p22ms+esφ¯+esφ~,\displaystyle H_{s}=\frac{p^{2}}{2m_{s}}+e_{s}\overline{\varphi}+e_{s}\widetilde{\varphi}, (415)

where ese_{s} is the electric charge, φ\varphi is the electrostatic potential, and φ¯\smash{\overline{\varphi}} and φ~\smash{\widetilde{\varphi}} are its average and oscillating parts, respectively. Then, H0s=H¯s=p2/(2ms)+esφ¯\smash{H_{0s}=\overline{H}_{s}=p^{2}/(2m_{s})+e_{s}\overline{\varphi}}, H~s=esφ~\smash{\widetilde{H}_{s}=e_{s}\widetilde{\varphi}}, 𝜶^s=es\smash{\widehat{\boldsymbol{\alpha}}_{s}=e_{s}}, and 𝑳^s=𝑹^s=𝟎^\smash{\widehat{\boldsymbol{L}}_{s}=\widehat{\boldsymbol{R}}_{s}=\widehat{\boldsymbol{0}}}, so s=𝟎\smash{{\boldsymbol{\wp}}_{s}={\boldsymbol{0}}}. The matrix (285) is a scalar (Wigner function) given by

𝖴(t,𝒙,ω,𝒌)=dτ2\upid𝒔(2\upi)nφ¯~(t+τ/2,𝒙+𝒔/2)φ¯~(tτ/2,𝒙𝒔/2)¯eiωτi𝒌𝒔.\mathsf{U}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}})=\int\frac{\mathrm{d}\tau}{2\upi}\,\frac{\mathrm{d}{\boldsymbol{s}}}{(2\upi)^{n}}\,\,\overline{{\underline{\widetilde{\varphi}}}(t+\tau/2,{\boldsymbol{x}}+{\boldsymbol{s}}/2)\,{\underline{\widetilde{\varphi}}}(t-\tau/2,{\boldsymbol{x}}-{\boldsymbol{s}}/2)}\,\mathrm{e}^{\mathrm{i}\omega\tau-\mathrm{i}{\boldsymbol{k}}\cdot{\boldsymbol{s}}}. (416)

(Underlining denotes the macroscopic part, ndim𝒙\smash{n\doteq\dim{\boldsymbol{x}}}, and the arguments (t,𝒙)\smash{(t,{\boldsymbol{x}})} will be omitted from now on.) Correspondingly,

𝗗s\displaystyle{\boldsymbol{\mathsf{D}}}_{s} es2d𝒌\upi𝒌𝒌𝖴(𝒌𝒗s,𝒌),\displaystyle\approx e_{s}^{2}\int\mathrm{d}{\boldsymbol{k}}\,\upi\,{\boldsymbol{k}}{\boldsymbol{k}}\mathsf{U}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}}), (417)
𝚯s\displaystyle{\boldsymbol{\Uptheta}}_{s} =es2ϑdωd𝒌𝒌𝒌𝖴(ω,𝒌)ω𝒌𝒗s+ϑ|ϑ=0,\displaystyle=e_{s}^{2}\,\frac{\partial}{\partial\vartheta}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\left.\frac{{\boldsymbol{k}}{\boldsymbol{k}}\mathsf{U}(\omega,{\boldsymbol{k}})}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\right|_{\vartheta=0}, (418)
Δs=Φs\displaystyle\Delta_{s}=\Phi_{s} =es22𝒑dωd𝒌𝒌𝖴(ω,𝒌)ω𝒌𝒗s,\displaystyle=\frac{e_{s}^{2}}{2}\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\frac{{\boldsymbol{k}}\mathsf{U}(\omega,{\boldsymbol{k}})}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}, (419)

and also

s=p22ms+esφ¯+Δs,𝒗s=𝒑ms+Δs𝒑.\displaystyle\mathcal{H}_{s}=\frac{p^{2}}{2m_{s}}+e_{s}\overline{\varphi}+\Delta_{s},\qquad{\boldsymbol{v}}_{s}=\frac{{\boldsymbol{p}}}{m_{s}}+\frac{\partial\Delta_{s}}{\partial{\boldsymbol{p}}}. (420)

The Lagrangian density of a free electrostatic field is

𝔏0=18\upiδij(iφ~)(jφ~)=xi(18\upiδijφ~(jφ~))+12φ~(δijij4\upi)φ~.\displaystyle\mathfrak{L}_{0}=\frac{1}{8\upi}\,\delta^{ij}(\partial_{i}\widetilde{\varphi})(\partial_{j}\widetilde{\varphi})=\frac{\partial}{\partial x^{i}}\left(\frac{1}{8\upi}\,\delta^{ij}\widetilde{\varphi}(\partial_{j}\widetilde{\varphi})\right)+\frac{1}{2}\,\widetilde{\varphi}\left(-\frac{\delta^{ij}\partial_{i}\partial_{j}}{4\upi}\right)\,\widetilde{\varphi}. (421)

The first term on the right-hand side does not contribute to the field action S0\smash{S_{0}} and thus can be ignored. The second term is of the form (208) with M=1\smash{M=1}, 𝗴=1\smash{{\boldsymbol{\mathsf{g}}}=1} (section 2.1.2), and 𝚵^0=k^2/(4\upi)\smash{\widehat{\boldsymbol{\Xi}}_{0}=\widehat{k}^{2}/(4\upi)}, so 𝚵0(ω,𝒌)=k2/(4\upi)\smash{{\boldsymbol{\Xi}}_{0}(\omega,{\boldsymbol{k}})=k^{2}/(4\upi)}, where k2𝒌2δijkikj\smash{k^{2}\equiv{\boldsymbol{k}}^{2}\equiv\delta^{ij}k_{i}k_{j}}. Then, (227) gives

𝚵(ω,𝒌)=Ξ(ω,𝒌)=k2ϵ(ω,𝒌)4\upi,\displaystyle{\boldsymbol{\Xi}}(\omega,{\boldsymbol{k}})=\Xi(\omega,{\boldsymbol{k}})=\frac{k^{2}\epsilon_{\parallel}(\omega,{\boldsymbol{k}})}{4\upi}, (422)

where the arguments t\smash{t} and 𝒙\smash{{\boldsymbol{x}}} are omitted for brevity and ϵ\smash{\epsilon_{\parallel}} is the parallel permittivity:

ϵ(ω,𝒌)=1+s4\upies2k2d𝒑𝒌ω𝒌𝒗s+i0Fs𝒑.\displaystyle\epsilon_{\parallel}(\omega,{\boldsymbol{k}})=1+\sum_{s}\frac{4\upi e_{s}^{2}}{k^{2}}\int\mathrm{d}{\boldsymbol{p}}\,\frac{{\boldsymbol{k}}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\mathrm{i}0}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}. (423)

9.1.2 Collisions and fluctuations

By (402), the spectrum of microscopic oscillations of φ~\smash{\widetilde{\varphi}} is a scalar given by

𝖲(ω,𝒌)=2\upis(4\upiesk2|ϵ(ω,𝒌)|)2d𝒑δ(ω𝒌𝒗s)Fs(𝒑),\displaystyle\mathsf{S}(\omega,{\boldsymbol{k}})=2\upi\sum_{s}\left(\frac{4\upi e_{s}}{k^{2}|\epsilon_{\parallel}(\omega,{\boldsymbol{k}})|}\right)^{2}\int\mathrm{d}{\boldsymbol{p}}\,\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})F_{s}({\boldsymbol{p}}), (424)

where we substituted n=3n=3 for three-dimensional plasma. For thermal equilibrium, (405) leads to the well-known formula (Lifshitz & Pitaevskii, 1981, section 51)

𝖲(ω,𝒌)=2Tωim(1Ξ(ω,𝒌))=8\upiTωk2im(1ϵ(ω,𝒌))=8\upiTωk2imϵ(ω,𝒌)|ϵ(ω,𝒌)|2.\displaystyle\mathsf{S}(\omega,{\boldsymbol{k}})=-\frac{2T}{\omega}\operatorname{im}\left(\frac{1}{\Xi(\omega,{\boldsymbol{k}})}\right)=-\frac{8\upi T}{\omega k^{2}}\operatorname{im}\left(\frac{1}{\epsilon_{\parallel}(\omega,{\boldsymbol{k}})}\right)=\frac{8\upi T}{\omega k^{2}}\frac{\operatorname{im}\epsilon_{\parallel}(\omega,{\boldsymbol{k}})}{|\epsilon_{\parallel}(\omega,{\boldsymbol{k}})|^{2}}. (425)

The spectrum 𝖲ρ\smash{\mathsf{S}_{\rho}} of charge-density fluctuations is found using Poisson’s equation ~ρ ~  =k^2 ~φ ~  /4\upi\smash{\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{\rho}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{\rho}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{\rho}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{\rho}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}=\widehat{k}^{2}\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{\varphi}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{\varphi}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{\varphi}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{\varphi}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}/4\upi}, whence 𝖲ρ(k2/4\upi)2𝖲\smash{\mathsf{S}_{\rho}\approx(k^{2}/4\upi)^{2}\mathsf{S}}. Fluctuations of other fields are found similarly. Also, (243) leads to

|𝒳ss(ω,𝒌;𝒑,𝒑)|2=(4\upiesesk2|ϵ(ω,𝒌)|)2.\displaystyle|\mathcal{X}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}=\left(\frac{4\upi e_{s}e_{s^{\prime}}}{k^{2}|\epsilon_{\parallel}(\omega,{\boldsymbol{k}})|}\right)^{2}. (426)

Then, (6.8) yields the standard Balescu–Lenard collision operator:

𝒞s=𝒑sd𝒌(2\upi)3d𝒑\displaystyle\mathcal{C}_{s}=\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\sum_{s^{\prime}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{3}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\, \upi𝒌𝒌|ϵ(𝒌𝒗s,𝒌)|2(4\upiesesk2)2δ(𝒌𝒗s𝒌𝒗s)\displaystyle\frac{\upi{\boldsymbol{k}}{\boldsymbol{k}}}{|\epsilon_{\parallel}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}})|^{2}}\,\left(\frac{4\upi e_{s}e_{s^{\prime}}}{k^{2}}\right)^{2}\delta({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})
(Fs(𝒑)𝒑Fs(𝒑)Fs(𝒑)Fs(𝒑)𝒑).\displaystyle\cdot\left(\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}\,F_{s^{\prime}}({\boldsymbol{p}}^{\prime})-F_{s}({\boldsymbol{p}})\,\frac{\partial F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial{\boldsymbol{p}}^{\prime}}\right). (427)

(As a reminder, the distribution functions are normalized such that d𝒑Fs\int\mathrm{d}{\boldsymbol{p}}\,F_{s} is the local average density of species s\smash{s} (199).)

9.1.3 On-shell waves

For on-shell waves, (354) gives 𝖴(ω,𝒌)=(h(𝒌)+h(𝒌))δ(ωw(𝒌))\smash{\mathsf{U}(\omega,{\boldsymbol{k}})=(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\delta(\omega-w({\boldsymbol{k}}))}, where w(𝒌)\smash{w({\boldsymbol{k}})} is determined by the dispersion relation

ϵH(w(𝒌),𝒌)=0,\displaystyle\epsilon_{\parallel\text{H}}(w({\boldsymbol{k}}),{\boldsymbol{k}})=0, (428)

and ϵHreϵ\smash{\epsilon_{\parallel\text{H}}\equiv\operatorname{re}\epsilon_{\parallel}} is given by

ϵH(ω,𝒌)=1+s4\upies2k2d𝒑𝒌ω𝒌𝒗sFs𝒑.\displaystyle\epsilon_{\parallel\text{H}}(\omega,{\boldsymbol{k}})=1+\sum_{s}\frac{4\upi e_{s}^{2}}{k^{2}}\fint\mathrm{d}{\boldsymbol{p}}\,\frac{{\boldsymbol{k}}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}. (429)

The phase-space density of the wave action, defined in (365), is

J(𝒌)=h(𝒌)ΞH(𝒌)ω=h(𝒌)k24\upiϵH(w(𝒌),𝒌)ω,\displaystyle J({\boldsymbol{k}})=h({\boldsymbol{k}})\,\frac{\partial\Xi_{\text{H}}({\boldsymbol{k}})}{\partial\omega}=h({\boldsymbol{k}})\,\frac{k^{2}}{4\upi}\frac{\partial\epsilon_{\parallel\text{H}}(w({\boldsymbol{k}}),{\boldsymbol{k}})}{\partial\omega}, (430)

and the dressing function (418) is given by

𝚯s\displaystyle{\boldsymbol{\Uptheta}}_{s} =es2ϑd𝒌(h(𝒌)+h(𝒌))𝒌𝒌w(𝒌)𝒌𝒗s+ϑ|ϑ=0\displaystyle=e_{s}^{2}\,\frac{\partial}{\partial\vartheta}\fint\mathrm{d}{\boldsymbol{k}}\,(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\left.\frac{{\boldsymbol{k}}{\boldsymbol{k}}}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\right|_{\vartheta=0}
=2es2ϑd𝒌h(𝒌)𝒌𝒌w(𝒌)𝒌𝒗s+ϑ|ϑ=0.\displaystyle=2e_{s}^{2}\,\frac{\partial}{\partial\vartheta}\fint\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\left.\frac{{\boldsymbol{k}}{\boldsymbol{k}}}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\right|_{\vartheta=0}. (431)

Using these, one obtains (appendix F.1.1)

sd𝒑𝒑Fs+d𝒌𝒌J=sd𝒑𝒑f¯s,\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{p}}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,{\boldsymbol{k}}J=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{p}}\overline{f}_{s}, (432)

so the conserved quantity (379) is the average momentum of the plasma (while the electrostatic field carries no momentum, naturally). Also (appendix F.1.2),

sd𝒑H0sFs+d𝒌wJ=sd𝒑H0sf¯s+18\upi𝑬~𝑬~¯,\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,wJ=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}\overline{f}_{s}+\frac{1}{8\upi}\,\overline{\smash{\widetilde{{\boldsymbol{E}}}}^{\dagger}\widetilde{{\boldsymbol{E}}}}, (433)

so, expectedly, the conserved quantity (381) is the average particle energy plus the energy of the electrostatic field. In combination with our equations for Fs\smash{F_{s}} and J\smash{J} (section 7.6), these results can be considered as a generalization and concise restatement of the OC QLT by Dewar (1973), which is rigorously reproduced from our general formulation as a particular case.

9.1.4 Eikonal waves

As a particular case, let us consider an eikonal wave

φ¯~re(eiθφ˘),ω¯tθ,𝒌¯𝒙θ,\displaystyle\underline{\widetilde{\varphi}}\approx\operatorname{re}(\mathrm{e}^{\mathrm{i}\theta}{\breve{\varphi}}),\qquad\overline{\omega}\doteq-\partial_{t}\theta,\qquad\overline{{\boldsymbol{k}}}\doteq\partial_{{\boldsymbol{x}}}\theta, (434)

which may or may not be on-shell. As seen from section 7.4.1,

𝖴|φ˘|24±δ(ω±ω¯)δ(𝒌±𝒌¯).\displaystyle\mathsf{U}\approx\frac{|{\breve{\varphi}}|^{2}}{4}\,\sum_{\pm}\delta(\omega\pm\overline{\omega})\,\delta({\boldsymbol{k}}\pm\overline{{\boldsymbol{k}}}). (435)

For nonresonant particles, the dressing function is well defined is found as follows:

𝚯s\displaystyle{\boldsymbol{\Uptheta}}_{s} dωd𝒌es2𝒌𝒌|φ˘|24(ω𝒌𝒗s)2±δ(ω±ω¯)δ(𝒌±𝒌¯)\displaystyle\approx-\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\frac{e_{s}^{2}{\boldsymbol{k}}{\boldsymbol{k}}|{\breve{\varphi}}|^{2}}{4(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})^{2}}\,\sum_{\pm}\delta(\omega\pm\overline{\omega})\,\delta({\boldsymbol{k}}\pm\overline{{\boldsymbol{k}}})
=es2𝒌¯𝒌¯|φ˘|22(ω¯𝒌¯𝒗s)2.\displaystyle=-\frac{e_{s}^{2}\overline{{\boldsymbol{k}}}\,\overline{{\boldsymbol{k}}}|{\breve{\varphi}}|^{2}}{2(\overline{\omega}-\overline{{\boldsymbol{k}}}\cdot{\boldsymbol{v}}_{s})^{2}}. (436)

Similarly, the ponderomotive energy for nonresonant particles is

Δs\displaystyle\Delta_{s} es2|φ˘|28ms𝒗sdωd𝒌𝒌ω𝒌𝒗s±δ(ω±ω¯)δ(𝒌±𝒌¯)\displaystyle\approx\frac{e_{s}^{2}|{\breve{\varphi}}|^{2}}{8m_{s}}\frac{\partial}{\partial{\boldsymbol{v}}_{s}}\cdot\int\mathrm{d}{\omega}\,\mathrm{d}{\boldsymbol{k}}\,\frac{{\boldsymbol{k}}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}\sum_{\pm}\delta(\omega\pm\overline{\omega})\,\delta({\boldsymbol{k}}\pm\overline{{\boldsymbol{k}}})
=es2|φ˘|28msdωd𝒌k2(ω𝒌𝒗s)2±δ(ω±ω¯)δ(𝒌±𝒌¯)\displaystyle=\frac{e_{s}^{2}|{\breve{\varphi}}|^{2}}{8m_{s}}\int\mathrm{d}{\omega}\,\mathrm{d}{\boldsymbol{k}}\,\frac{k^{2}}{(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})^{2}}\sum_{\pm}\delta(\omega\pm\overline{\omega})\,\delta({\boldsymbol{k}}\pm\overline{{\boldsymbol{k}}})
=es2k¯2|φ˘|24ms(ω¯𝒌¯𝒗s)2,\displaystyle=\frac{e_{s}^{2}\overline{k}^{2}|{\breve{\varphi}}|^{2}}{4m_{s}(\overline{\omega}-\overline{{\boldsymbol{k}}}\cdot{\boldsymbol{v}}_{s})^{2}}, (437)

in agreement with (Dewar, 1972; Cary & Kaufman, 1977). One can also express these functions in terms of the electric-field envelope 𝑬˘i𝒌¯φ˘\smash{{\breve{{\boldsymbol{E}}}}\approx-\mathrm{i}\overline{{\boldsymbol{k}}}{\breve{\varphi}}}:

𝚯ses2𝑬˘𝑬˘2(ω¯𝒌¯𝒗s)2,Δses2|𝑬˘|24ms(ω¯𝒌¯𝒗s)2.\displaystyle{\boldsymbol{\Uptheta}}_{s}\approx-\frac{e_{s}^{2}{\breve{{\boldsymbol{E}}}}\smash{{\breve{{\boldsymbol{E}}}}}^{\dagger}}{2(\overline{\omega}-\overline{{\boldsymbol{k}}}\cdot{\boldsymbol{v}}_{s})^{2}},\qquad\Delta_{s}\approx\frac{e_{s}^{2}|\smash{{\breve{{\boldsymbol{E}}}}}|^{2}}{4m_{s}(\overline{\omega}-\overline{{\boldsymbol{k}}}\cdot{\boldsymbol{v}}_{s})^{2}}. (438)

For on-shell in particular, one can use (430) together with h(𝒌)=14|φ˘|2δ(𝒌𝒌¯)\smash{h({\boldsymbol{k}})=\frac{1}{4}|{\breve{\varphi}}|^{2}\delta({\boldsymbol{k}}-\overline{{\boldsymbol{k}}})} (cf. (355)) to obtain the well-known expression for the wave action density d𝒌J\smash{\mathcal{I}\doteq\int\mathrm{d}{\boldsymbol{k}}\,J}:

=|𝑬˘|216\upiϵH(ω,𝒌¯)ω|ω=w(𝒌¯).\displaystyle\mathcal{I}=\frac{|\smash{{\breve{{\boldsymbol{E}}}}}|^{2}}{16\upi}\frac{\partial\epsilon_{\parallel\text{H}}(\omega,\overline{{\boldsymbol{k}}})}{\partial\omega}\,\Big{|}_{\omega=w(\overline{{\boldsymbol{k}}})}. (439)

For non-too-hot plasma, one has ϵH(ω,𝒌)1ωp2/ω2\smash{\epsilon_{\parallel\text{H}}(\omega,{\boldsymbol{k}})\approx 1-\omega_{p}^{2}/\omega^{2}}, where ωps4\upi𝒩ses2/ms\smash{\omega_{p}\doteq\sum_{s}4\upi\mathcal{N}_{s}e_{s}^{2}/m_{s}} is the plasma frequency. The corresponding waves are Langmuir waves. Their dispersion relation is w(𝒌¯)±ωp\smash{w(\overline{{\boldsymbol{k}}})\approx\pm\omega_{p}}, so ±|𝑬˘|2/(8\upiωp)\smash{\mathcal{I}\approx\pm|{\breve{{\boldsymbol{E}}}}|^{2}/(8\upi\omega_{p})} (and accordingly, the wave energy density is w=w0\smash{\mathcal{E}_{\text{w}}=w\mathcal{I}\geq 0} for either sign). Remember, though, that this expression is only approximate. Using it instead of (439) can result in violation of the exact conservation laws of QLT. Conservation of the Langmuir-wave action in non-stationary plasmas beyond the cold-plasma approximation is also discussed in (Dodin et al., 2009; Dodin & Fisch, 2010b; Schmit et al., 2010).

9.1.5 Homogeneous plasma

In homogeneous nn-dimensional plasma of a given volume 𝒱n\mathscr{V}_{n}, the Wigner function (416) has the form 𝖴=𝒰(t,𝒌)δ(ωw(t,𝒌))\smash{\mathsf{U}=\mathcal{U}(t,{\boldsymbol{k}})\delta(\omega-w(t,{\boldsymbol{k}}))}. The function 𝒰\mathcal{U} is readily found using (207):

𝒰(t,𝒌)=1𝒱nd𝒙dωU=1𝒱n|φ~¯̊(t,𝒌)|2¯=1𝒱n|φ~¯̊(t,𝒌)|2.\displaystyle\mathcal{U}(t,{\boldsymbol{k}})=\frac{1}{\mathscr{V}_{n}}\int\mathrm{d}{\boldsymbol{x}}\,\mathrm{d}\omega\,U=\frac{1}{\mathscr{V}_{n}}\,|\overline{\mathring{\underline{\widetilde{\varphi}}}(t,{\boldsymbol{k}})|^{2}}=\frac{1}{\mathscr{V}_{n}}\,|\mathring{\underline{\widetilde{\varphi}}}(t,{\boldsymbol{k}})|^{2}. (440)

Then,

𝗗s\upies2𝒱nd𝒌𝒌𝒌|φ~¯̊(t,𝒌)|2δ(w(t,𝒌)𝒌𝒗s).\displaystyle{\boldsymbol{\mathsf{D}}}_{s}\approx\frac{\upi e_{s}^{2}}{\mathscr{V}_{n}}\int\mathrm{d}{\boldsymbol{k}}\,{\boldsymbol{k}}{\boldsymbol{k}}\,|\mathring{\underline{\widetilde{\varphi}}}(t,{\boldsymbol{k}})|^{2}\,\delta(w(t,{\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}). (441)

This coincides with the well-known formula for the QL-diffusion coefficient in homogeneous electrostatic plasma.393939See, for example, equation (16.17) in (Stix, 1992). The extra mass factor appears there because QL diffusion is considered in the velocity space instead of the momentum space. The functions 𝚯s\smash{{\boldsymbol{\Uptheta}}_{s}} and Δs\smash{\Delta_{s}} are also important in homogeneous turbulence in that they ensure the proper energy–momentum conservation; for example, see (Stix, 1992, section 16.3) and (Liu & Dodin, 2015, section II.2). These functions can be expressed through φ~¯̊\smash{\mathring{\underline{\widetilde{\varphi}}}} too. However, they have a simpler representation in terms of the Wigner function 𝖴\smash{\mathsf{U}}, as in (418) and (419), respectively. This is because 𝖴\smash{\mathsf{U}} is a local property of the field, which makes it more fundamental than the amplitudes of global Fourier harmonics commonly used in the literature.

9.2 Relativistic electromagnetic interactions

9.2.1 Main equations

Let us extend the above results to relativistic electromagnetic interactions. In this case,

Hs=ms2c4+(𝒑ces𝑨)2+esφ,\displaystyle H_{s}=\sqrt{m_{s}^{2}c^{4}+({\boldsymbol{p}}c-e_{s}{\boldsymbol{A}})^{2}}+e_{s}\varphi, (442)

where c\smash{c} is the speed of light and 𝑨\smash{{\boldsymbol{A}}} is the vector potential. Let us adopt the Weyl gauge for the oscillating part of the electromagnetic field (φ~=0\smash{\widetilde{\varphi}=0}) and Taylor-expand Hs\smash{H_{s}} to the second order in 𝑨~\smash{\widetilde{{\boldsymbol{A}}}}. This leads to

HsH0ses𝜷s𝑨~+es22c2𝑨~𝝁s1𝑨~,\displaystyle\displaystyle H_{s}\approx H_{0s}-e_{s}{\boldsymbol{\beta}}_{s}^{\dagger}\widetilde{{\boldsymbol{A}}}+\frac{e_{s}^{2}}{2c^{2}}\,\smash{\widetilde{{\boldsymbol{A}}}}^{\dagger}{\boldsymbol{\mu}}_{s}^{-1}\widetilde{{\boldsymbol{A}}}, (443)
H0s=ms2c4+(𝒑ces𝑨¯)2+esφ¯\displaystyle\displaystyle H_{0s}=\sqrt{m_{s}^{2}c^{4}+({\boldsymbol{p}}c-e_{s}\overline{{\boldsymbol{A}}})^{2}}+e_{s}\overline{\varphi} (444)

(although plasma is assumed nonmagnetized, a weak average magnetic field 𝑩¯=×𝑨¯\smash{\overline{{\boldsymbol{B}}}=\nabla\times\overline{{\boldsymbol{A}}}} is allowed, so 𝑨¯\smash{\overline{{\boldsymbol{A}}}} can be order-one and thus generally must be retained), where

𝜷s=1mscγs(𝒑esc𝑨¯),𝝁s1=𝟏𝜷s𝜷smsγs,\displaystyle{\boldsymbol{\beta}}_{s}=\frac{1}{m_{s}c\gamma_{s}}\left({\boldsymbol{p}}-\frac{e_{s}}{c}\,\overline{{\boldsymbol{A}}}\right),\qquad{\boldsymbol{\mu}}_{s}^{-1}=\frac{{\boldsymbol{1}}-{\boldsymbol{\beta}}_{s}{\boldsymbol{\beta}}_{s}^{\dagger}}{m_{s}\gamma_{s}}, (445)

and γs(1βs2)1/2\smash{\gamma_{s}\doteq(1-\beta_{s}^{2})^{-1/2}}. In the equations presented below, 𝜷s=𝒗s/c\smash{{\boldsymbol{\beta}}_{s}={\boldsymbol{v}}_{s}/c} (where 𝒗s\smash{{\boldsymbol{v}}_{s}} is the OC velocity) is a sufficiently accurate approximation. Also, 𝝁s(𝒑𝒑2H0s)1\smash{{\boldsymbol{\mu}}_{s}\equiv(\partial^{2}_{{\boldsymbol{p}}{\boldsymbol{p}}}H_{0s})^{-1}} can be interpreted as the relativistic-mass tensor.

Let us choose the field 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}} of our general theory to be the oscillating electric field 𝑬~=iω^𝑨~/c\smash{\widetilde{{\boldsymbol{E}}}=\mathrm{i}\widehat{\omega}\widetilde{{\boldsymbol{A}}}/c}; then (cf. (209)),

𝜶^s=ies𝒗sω^1,𝑳^s=es2ω^1,𝑹^s=𝝁s1ω^1.\displaystyle\widehat{\boldsymbol{\alpha}}_{s}=\mathrm{i}e_{s}{\boldsymbol{v}}_{s}\widehat{\omega}^{-1},\qquad\widehat{\boldsymbol{L}}_{s}=e_{s}^{2}\widehat{\omega}^{-1},\qquad\widehat{\boldsymbol{R}}_{s}={\boldsymbol{\mu}}_{s}^{-1}\widehat{\omega}^{-1}. (446)

(Other ways to identify 𝑳^s\smash{\widehat{\boldsymbol{L}}_{s}} and 𝑹^s\smash{\widehat{\boldsymbol{R}}_{s}} are also possible and lead to the same results.) Then,

𝜶s=ies𝒗sω,s=es2ω2𝝁s1.\displaystyle{\boldsymbol{\alpha}}_{s}=\frac{\mathrm{i}e_{s}{\boldsymbol{v}}_{s}}{\omega},\qquad{\boldsymbol{\wp}}_{s}=\frac{e_{s}^{2}}{\omega^{2}}\,{\boldsymbol{\mu}}_{s}^{-1}. (447)

The average Wigner matrix of 𝑬~\smash{\widetilde{{\boldsymbol{E}}}} is

𝗨(t,𝒙,ω,𝒌)=dτ2\upid𝒔(2\upi)3𝑬¯~(t+τ/2,𝒙+𝒔/2)𝑬¯~(tτ/2,𝒙𝒔/2)¯eiωτi𝒌𝒔{\boldsymbol{\mathsf{U}}}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}})=\int\frac{\mathrm{d}\tau}{2\upi}\,\frac{\mathrm{d}{\boldsymbol{s}}}{(2\upi)^{3}}\,\,\overline{\underline{\widetilde{{\boldsymbol{E}}}}(t+\tau/2,{\boldsymbol{x}}+{\boldsymbol{s}}/2)\,\smash{\underline{\widetilde{{\boldsymbol{E}}}}}^{\dagger}(t-\tau/2,{\boldsymbol{x}}-{\boldsymbol{s}}/2)}\,\mathrm{e}^{\mathrm{i}\omega\tau-\mathrm{i}{\boldsymbol{k}}\cdot{\boldsymbol{s}}} (448)

(the arguments t\smash{t} and 𝒙\smash{{\boldsymbol{x}}} are henceforth omitted), and the nonlinear potentials are

𝗗s\displaystyle{\boldsymbol{\mathsf{D}}}_{s} =\upies2d𝒌𝒌𝒌𝒗s𝗨(𝒌𝒗s,𝒌)𝒗s(𝒌𝒗s)2,\displaystyle=\upi e_{s}^{2}\int\mathrm{d}{\boldsymbol{k}}\,{\boldsymbol{k}}{\boldsymbol{k}}\,\frac{{\boldsymbol{v}}_{s}^{\dagger}{\boldsymbol{\mathsf{U}}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}}){\boldsymbol{v}}_{s}}{({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})^{2}}, (449)
𝚯s\displaystyle{\boldsymbol{\Uptheta}}_{s} =es2ϑdωd𝒌𝒌𝒌ω2(𝒗s𝗨𝒗s)ω𝒌𝒗s+ϑ|ϑ=0,\displaystyle=e_{s}^{2}\,\frac{\partial}{\partial\vartheta}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\left.\frac{{\boldsymbol{k}}{\boldsymbol{k}}}{\omega^{2}}\frac{({\boldsymbol{v}}_{s}^{\dagger}{\boldsymbol{\mathsf{U}}}{\boldsymbol{v}}_{s})}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\right|_{\vartheta=0}, (450)
Δs\displaystyle\Delta_{s} =es22𝒑dωd𝒌𝒌ω2(𝒗s𝗨𝒗s)ω𝒌𝒗s+es22dωd𝒌tr(𝗨𝝁s1)ω2.\displaystyle=\frac{e_{s}^{2}}{2}\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\frac{{\boldsymbol{k}}}{\omega^{2}}\frac{({\boldsymbol{v}}_{s}^{\dagger}{\boldsymbol{\mathsf{U}}}{\boldsymbol{v}}_{s})}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}+\frac{e_{s}^{2}}{2}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\frac{\operatorname{tr}({\boldsymbol{\mathsf{U}}}{\boldsymbol{\mu}}_{s}^{-1})}{\omega^{2}}. (451)

When plasma is nonrelativistic and the field is electrostatic (so 𝗨=𝒌𝒌𝖴φ\smash{{\boldsymbol{\mathsf{U}}}={\boldsymbol{k}}{\boldsymbol{k}}^{\dagger}\mathsf{U}_{\varphi}}, where 𝖴φ\smash{\mathsf{U}_{\varphi}} is scalar), (449) gives the same 𝗗s\smash{{\boldsymbol{\mathsf{D}}}_{s}} as (417) and (451) gives the same Δs\smash{\Delta_{s}} as (419). For 𝚯s\smash{{\boldsymbol{\Uptheta}}_{s}}, the equivalence between (450) and (418) should not be expected because 𝚯s\smash{{\boldsymbol{\Uptheta}}_{s}} is a part of a distribution function, which is not gauge-invariant. (Canonical momenta in the Weyl gauge are different from those in the electrostatic gauge.) But it is precisely the dressing function (450) that leads to the correct expressions for the momentum and energy stored in the OC distribution (section 9.2.3).

The Lagrangian density of a free electromagnetic field is

𝔏0=𝑬~𝑬~𝑩~𝑩~8\upi.\displaystyle\mathfrak{L}_{0}=\frac{\smash{\widetilde{{\boldsymbol{E}}}}^{\dagger}\widetilde{{\boldsymbol{E}}}-\smash{\widetilde{{\boldsymbol{B}}}}^{\dagger}\widetilde{{\boldsymbol{B}}}}{8\upi}. (452)

From Faraday’s law, one has 𝑩~=ω^1c(𝒌^×𝑬)\smash{\widetilde{{\boldsymbol{B}}}=\widehat{\omega}^{-1}c(\widehat{\boldsymbol{k}}\times{\boldsymbol{E}})}.404040Here, the oscillating field 𝚿~=𝑬~\smash{\widetilde{{\boldsymbol{\Psi}}}=\widetilde{{\boldsymbol{E}}}} has the same dimension as 𝒙\smash{{\boldsymbol{x}}}, so the standard vector notation (including the dot product and the cross product) is naturally extended to 𝚿~\smash{\widetilde{{\boldsymbol{\Psi}}}}. Then, 𝑩~𝑩~/c2\smash{-\smash{\widetilde{{\boldsymbol{B}}}}^{\dagger}\widetilde{{\boldsymbol{B}}}/c^{2}} can be represented as follows (up to a divergence, which is insignificant):

(𝑬~×ω^1𝒌^)(ω^1𝒌^×𝑬~)\displaystyle(\widetilde{{\boldsymbol{E}}}\times\widehat{\omega}^{-1}\widehat{\boldsymbol{k}})\cdot(\widehat{\omega}^{-1}\widehat{\boldsymbol{k}}\times\widetilde{{\boldsymbol{E}}}) =𝑬~ω^2(𝒌^×(𝒌^×𝑬~))\displaystyle=\widetilde{{\boldsymbol{E}}}\cdot\widehat{\omega}^{-2}(\widehat{\boldsymbol{k}}\times(\widehat{\boldsymbol{k}}\times\widetilde{{\boldsymbol{E}}}))
=𝑬~ω^2(𝒌^(𝒌^𝑬~)𝑬~k^2)\displaystyle=\smash{\widetilde{{\boldsymbol{E}}}}^{\dagger}\widehat{\omega}^{-2}(\widehat{\boldsymbol{k}}(\widehat{\boldsymbol{k}}\cdot\widetilde{{\boldsymbol{E}}})-\widetilde{{\boldsymbol{E}}}\widehat{k}^{2})
=𝑬~ω^2(𝒌^𝒌^𝟏k^2)𝑬~.\displaystyle=\smash{\widetilde{{\boldsymbol{E}}}}^{\dagger}\widehat{\omega}^{-2}(\widehat{\boldsymbol{k}}\smash{\widehat{\boldsymbol{k}}}^{\dagger}-{\boldsymbol{1}}\widehat{k}^{2})\widetilde{{\boldsymbol{E}}}. (453)

Then, the vacuum dispersion operator can be written as (cf. (208))

𝚵^0(ω,𝒌)=14\upi(𝟏+c2ω^2(𝒌^𝒌^𝟏k^2)).\displaystyle\widehat{\boldsymbol{\Xi}}_{0}(\omega,{\boldsymbol{k}})=\frac{1}{4\upi}\,\left({\boldsymbol{1}}+c^{2}\widehat{\omega}^{-2}\big{(}\widehat{\boldsymbol{k}}\smash{\widehat{\boldsymbol{k}}}^{\dagger}-{\boldsymbol{1}}\widehat{k}^{2}\big{)}\right). (454)

The total dispersion matrix is readily found to be

𝚵(ω,𝒌)=14\upi(ϵ(ω,𝒌)+c2ω2(𝒌𝒌𝟏k2)),\displaystyle{\boldsymbol{\Xi}}(\omega,{\boldsymbol{k}})=\frac{1}{4\upi}\left({\boldsymbol{\epsilon}}(\omega,{\boldsymbol{k}})+\frac{c^{2}}{\omega^{2}}\,({\boldsymbol{k}}\smash{{\boldsymbol{k}}}^{\dagger}-{\boldsymbol{1}}k^{2})\right), (455)

where ϵ\smash{{\boldsymbol{\epsilon}}} (not to be confused with the small parameter ϵ\smash{\epsilon} that we introduced earlier) is the dielectric tensor:

ϵ(ω,𝒌)=𝟏𝖜pω2+s4\upies2ω2d𝒑𝒗s𝒗sω𝒌𝒗s+i0𝒌Fs𝒑.\displaystyle{\boldsymbol{\epsilon}}(\omega,{\boldsymbol{k}})={\boldsymbol{1}}-\frac{{\boldsymbol{\mathfrak{w}}}_{p}}{\omega^{2}}+\sum_{s}\frac{4\upi e_{s}^{2}}{\omega^{2}}\int\mathrm{d}{\boldsymbol{p}}\,\frac{{\boldsymbol{v}}_{s}{\boldsymbol{v}}_{s}^{\dagger}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\mathrm{i}0}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}. (456)

Here, 𝖜p\smash{{\boldsymbol{\mathfrak{w}}}_{p}} is the squared relativistic plasma frequency, which is a matrix, because the ‘masses’ 𝝁s\smash{{\boldsymbol{\mu}}_{s}} are matrices:

𝖜ps4\upies2d𝒑Fs𝝁s1.\displaystyle{\boldsymbol{\mathfrak{w}}}_{p}\doteq\sum_{s}4\upi e_{s}^{2}\int\mathrm{d}{\boldsymbol{p}}\,F_{s}{\boldsymbol{\mu}}_{s}^{-1}. (457)

9.2.2 Collisions and fluctuations

By (402), the spectrum of microscopic oscillations of 𝑬~\smash{\widetilde{{\boldsymbol{E}}}} is a matrix given by

𝗦(ω,𝒌)=2\upis(4πesω)2d𝒑δ(ω𝒌𝒗s)Fs(𝒑)ϵ1(ω,𝒌)𝒗s𝒗sϵ(ω,𝒌).{\boldsymbol{\mathsf{S}}}(\omega,{\boldsymbol{k}})=2\upi\,\sum_{s}\left(\frac{4\pi e_{s}}{\omega}\right)^{2}\int\mathrm{d}{\boldsymbol{p}}\,\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})F_{s}({\boldsymbol{p}}){\boldsymbol{\epsilon}}^{-1}(\omega,{\boldsymbol{k}})\,{\boldsymbol{v}}_{s}{\boldsymbol{v}}_{s}^{\dagger}\,{\boldsymbol{\epsilon}}^{-{\dagger}}(\omega,{\boldsymbol{k}}). (458)

In the electrostatic limit, one can replace ϵ1\smash{{\boldsymbol{\epsilon}}^{-1}} with ϵ1𝒌𝒌/k2\smash{\epsilon_{\parallel}^{-1}{\boldsymbol{k}}{\boldsymbol{k}}^{\dagger}/k^{2}}, where ϵ\smash{\epsilon_{\parallel}} is the relativistic generalization of (423); then (458) leads to (424) as a particular case. For thermal equilibrium, one can also use (405) and the following form of ϵ1\smash{{\boldsymbol{\epsilon}}^{-1}} for isotropic plasma:

ϵ1=1ϵ(𝟏𝒌𝒌k2)+1ϵ𝒌𝒌k2,{\boldsymbol{\epsilon}}^{-1}=\frac{1}{\epsilon_{\perp}}\bigg{(}{\boldsymbol{1}}-\frac{{\boldsymbol{k}}{\boldsymbol{k}}^{\dagger}}{k^{2}}\bigg{)}+\frac{1}{\epsilon_{\parallel}}\frac{{\boldsymbol{k}}{\boldsymbol{k}}^{\dagger}}{k^{2}}, (459)

where ϵ\smash{\epsilon_{\perp}} is the (scalar) transverse permittivity. Also, (243) leads to

|𝒳ss(ω,𝒌;𝒑,𝒑)|2(4\upiesesω2)2|𝒗sϵ1(ω,𝒌)𝒗s|2.\displaystyle|\mathcal{X}_{ss^{\prime}}(\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})|^{2}\approx\left(\frac{4\upi e_{s}e_{s^{\prime}}}{\omega^{2}}\right)^{2}|{\boldsymbol{v}}_{s}^{\dagger}{\boldsymbol{\epsilon}}^{-1}(\omega,{\boldsymbol{k}}){\boldsymbol{v}}^{\prime}_{s^{\prime}}|^{2}. (460)

Then the collision operator (6.8) is obtained in the form

𝒞s=𝒑s2es2es2\displaystyle\mathcal{C}_{s}=\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\sum_{s^{\prime}}2e_{s}^{2}e_{s^{\prime}}^{2}\int d𝒌(2\upi)3d𝒑|𝒗sϵ1(𝒌𝒗s,𝒌)𝒗s|2(𝒌𝒗s)4δ(𝒌𝒗s𝒌𝒗s)\displaystyle\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{3}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,\frac{|{\boldsymbol{v}}_{s}^{\dagger}{\boldsymbol{\epsilon}}^{-1}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}}){\boldsymbol{v}}^{\prime}_{s^{\prime}}|^{2}}{({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})^{4}}\,\delta({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})
×𝒌𝒌(Fs(𝒑)𝒑Fs(𝒑)Fs(𝒑)Fs(𝒑)𝒑),\displaystyle\times{\boldsymbol{k}}{\boldsymbol{k}}\cdot\left(\frac{\partial F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}\,F_{s^{\prime}}({\boldsymbol{p}}^{\prime})-F_{s}({\boldsymbol{p}})\,\frac{\partial F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial{\boldsymbol{p}}^{\prime}}\right), (461)

which is in agreement with (Hizanidis et al., 1983; Silin, 1961). Replacing ϵ1\smash{{\boldsymbol{\epsilon}}^{-1}} with ϵ1𝒌𝒌/k2\smash{\epsilon_{\parallel}^{-1}{\boldsymbol{k}}{\boldsymbol{k}}^{\dagger}/k^{2}} leads to the standard Balescu–Lenard operator (427) as a particular case.

9.2.3 On-shell waves

Electromagnetic on-shell waves satisfy

(ϵH(w(𝒌),𝒌)+c2w(𝒌)2(𝒌𝒌𝟏k2))𝑬˘=0,\displaystyle\left({\boldsymbol{\epsilon}}_{\text{H}}(w({\boldsymbol{k}}),{\boldsymbol{k}})+\frac{c^{2}}{w({\boldsymbol{k}})^{2}}\,({\boldsymbol{k}}\smash{{\boldsymbol{k}}}^{\dagger}-{\boldsymbol{1}}k^{2})\right){\breve{{\boldsymbol{E}}}}=0, (462)

where 𝑬˘\smash{{\breve{{\boldsymbol{E}}}}} is the complex envelope vector parallel to the polarization vector 𝜼\smash{{\boldsymbol{\eta}}}; also,

ϵH(ω,𝒌)=𝟏𝖜pω2+s4\upies2ω2d𝒑𝒗s𝒗sω𝒌𝒗s𝒌Fs𝒑.\displaystyle{\boldsymbol{\epsilon}}_{\text{H}}(\omega,{\boldsymbol{k}})={\boldsymbol{1}}-\frac{{\boldsymbol{\mathfrak{w}}}_{p}}{\omega^{2}}+\sum_{s}\frac{4\upi e_{s}^{2}}{\omega^{2}}\fint\mathrm{d}{\boldsymbol{p}}\,\frac{{\boldsymbol{v}}_{s}{\boldsymbol{v}}_{s}^{\dagger}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}. (463)

This yields (see (453))

𝑬˘ϵH(w(𝒌),𝒌)𝑬˘=c2w(𝒌)2𝑬˘(𝒌𝒌𝟏k2)𝑬˘=𝑩˘𝑩˘.\displaystyle\smash{{\breve{{\boldsymbol{E}}}}}^{\dagger}{\boldsymbol{\epsilon}}_{\text{H}}(w({\boldsymbol{k}}),{\boldsymbol{k}}){\breve{{\boldsymbol{E}}}}=-\frac{c^{2}}{w({\boldsymbol{k}})^{2}}\,\smash{{\breve{{\boldsymbol{E}}}}}^{\dagger}({\boldsymbol{k}}\smash{{\boldsymbol{k}}}^{\dagger}-{\boldsymbol{1}}k^{2}){\breve{{\boldsymbol{E}}}}=\smash{{\breve{{\boldsymbol{B}}}}}^{\dagger}{\breve{{\boldsymbol{B}}}}. (464)

Then, the phase-space density of the wave action (365) can be cast in the form

J(𝒌)=h(𝒌)𝜼𝚵H(ω,𝒌)ω𝜼|ω=w(𝒌¯)=h(𝒌)4\upiω2𝜼(ω2ϵH(ω,𝒌))ω𝜼|ω=w(𝒌¯)\displaystyle J({\boldsymbol{k}})=h({\boldsymbol{k}}){\boldsymbol{\eta}}^{\dagger}\,\frac{\partial{\boldsymbol{\Xi}}_{\text{H}}(\omega,{\boldsymbol{k}})}{\partial\omega}\,{\boldsymbol{\eta}}\,\Big{|}_{\omega=w(\overline{{\boldsymbol{k}}})}=\frac{h({\boldsymbol{k}})}{4\upi\omega^{2}}\,{\boldsymbol{\eta}}^{\dagger}\,\frac{\partial(\omega^{2}{\boldsymbol{\epsilon}}_{\text{H}}(\omega,{\boldsymbol{k}}))}{\partial\omega}\,{\boldsymbol{\eta}}\,\Big{|}_{\omega=w(\overline{{\boldsymbol{k}}})} (465)

(cf. (Dodin et al., 2019)), and the dressing function (450) is given by

𝚯s\displaystyle{\boldsymbol{\Uptheta}}_{s} =es2ϑd𝒌(h(𝒌)+h(𝒌))𝒌𝒌w2(𝒌)(𝒗s𝜼𝜼𝒗s)w(𝒌)𝒌𝒗s+ϑ|ϑ=0\displaystyle=e_{s}^{2}\,\frac{\partial}{\partial\vartheta}\fint\mathrm{d}{\boldsymbol{k}}\,(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\left.\frac{{\boldsymbol{k}}{\boldsymbol{k}}}{w^{2}({\boldsymbol{k}})}\frac{({\boldsymbol{v}}_{s}^{\dagger}{\boldsymbol{\eta}}{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{v}}_{s})}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\right|_{\vartheta=0}
=2es2ϑd𝒌h(𝒌)𝒌𝒌w2(𝒌)(𝜼𝒗s𝒗s𝜼)w(𝒌)𝒌𝒗s+ϑ|ϑ=0.\displaystyle=2e_{s}^{2}\,\frac{\partial}{\partial\vartheta}\fint\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\left.\frac{{\boldsymbol{k}}{\boldsymbol{k}}}{w^{2}({\boldsymbol{k}})}\frac{({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{v}}_{s}{\boldsymbol{v}}_{s}^{\dagger}{\boldsymbol{\eta}})}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\right|_{\vartheta=0}. (466)

Using these, one obtains (appendix F.2.1)

sd𝒑𝒑Fs+d𝒌𝒌J=𝓟(kin)+𝑬~×𝑩~¯4\upic,\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{p}}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,{\boldsymbol{k}}J={\boldsymbol{\mathcal{P}}}^{(\text{kin})}+\frac{\overline{\widetilde{{\boldsymbol{E}}}\times\widetilde{{\boldsymbol{B}}}}}{4\upi c}, (467)

where 𝓟(kin)\smash{{\boldsymbol{\mathcal{P}}}^{(\text{kin})}} is the average density of the plasma kinetic (up to 𝑨¯\smash{\overline{{\boldsymbol{A}}}}) momentum,

𝓟(kin)sd𝒑(𝒑es𝑨~/c)fs¯=sd𝒑𝒑fs(kin)¯,\displaystyle{\boldsymbol{\mathcal{P}}}^{(\text{kin})}\doteq\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\overline{({\boldsymbol{p}}-e_{s}\widetilde{{\boldsymbol{A}}}/c)f_{s}}=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{p}}\,\overline{f_{s}^{\text{(kin)}}}, (468)

the functions fs(kin)(𝒑)fs(𝒑+es𝑨~/c)\smash{f_{s}^{\text{(kin)}}({\boldsymbol{p}})\doteq f_{s}({\boldsymbol{p}}+e_{s}\widetilde{{\boldsymbol{A}}}/c)} are the distributions of kinetic (up to 𝑨¯\smash{\overline{{\boldsymbol{A}}}}) momenta, and the second term in (467) is the well-known average momentum of electromagnetic field. Similarly (appendix F.2.2),

sd𝒑H0sFs+d𝒌wJ=𝒦(kin)+18\upi(𝑬~𝑬~+𝑩~𝑩~)¯,\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,wJ=\mathcal{K}^{(\text{kin})}+\frac{1}{8\upi}\,\overline{\big{(}\smash{\widetilde{{\boldsymbol{E}}}}^{\dagger}\widetilde{{\boldsymbol{E}}}+\smash{\widetilde{{\boldsymbol{B}}}}^{\dagger}\widetilde{{\boldsymbol{B}}}\big{)}}, (469)

where 𝒦(kin)\smash{\mathcal{K}^{(\text{kin})}} is given by

𝒦(kin)sd𝒑H0sfs(kin)¯.\displaystyle\mathcal{K}^{(\text{kin})}\doteq\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}\,\overline{f_{s}^{(\text{kin})}}. (470)

In other words, the total momentum and energy of the system in the OC–wave representation are the same as those in the original particle–field variables.

9.2.4 Eikonal waves

As a particular case, let us consider an eikonal wave

𝑬¯~re(eiθ𝑬˘),ω¯tθ,𝒌¯𝒙θ,\displaystyle\underline{\widetilde{{\boldsymbol{E}}}}\approx\operatorname{re}(\mathrm{e}^{\mathrm{i}\theta}{\breve{{\boldsymbol{E}}}}),\qquad\overline{\omega}\doteq-\partial_{t}\theta,\qquad\overline{{\boldsymbol{k}}}\doteq\partial_{{\boldsymbol{x}}}\theta, (471)

which may or may not be on-shell. Then, (450) and (451) lead to (cf. section 9.1.4)

𝚯s=𝒌¯𝒌¯ω¯2es2|𝒗s𝑬˘|22(ω¯𝒌¯𝒗s)2,\displaystyle\displaystyle{\boldsymbol{\Uptheta}}_{s}=-\frac{\overline{{\boldsymbol{k}}}\,\overline{{\boldsymbol{k}}}}{\overline{\omega}^{2}}\frac{e_{s}^{2}|{\boldsymbol{v}}_{s}^{\dagger}{\breve{{\boldsymbol{E}}}}|^{2}}{2(\overline{\omega}-\overline{{\boldsymbol{k}}}\cdot{\boldsymbol{v}}_{s})^{2}}, (472)
Δs=es2𝑬˘𝝁s1𝑬˘4ω¯2+es2𝒌¯4ω¯2𝒑(|𝒗s𝑬˘|2ω¯𝒌¯𝒗s).\displaystyle\displaystyle\Delta_{s}=\frac{e_{s}^{2}\smash{{\breve{{\boldsymbol{E}}}}}^{\dagger}{\boldsymbol{\mu}}_{s}^{-1}{\breve{{\boldsymbol{E}}}}}{4\overline{\omega}^{2}}+\frac{e_{s}^{2}\overline{{\boldsymbol{k}}}}{4\overline{\omega}^{2}}\cdot\frac{\partial}{\partial{\boldsymbol{p}}}\bigg{(}\frac{|{\boldsymbol{v}}_{s}^{\dagger}{\breve{{\boldsymbol{E}}}}|^{2}}{\overline{\omega}-\overline{{\boldsymbol{k}}}\cdot{\boldsymbol{v}}_{s}}\bigg{)}. (473)

For on-shell waves in particular, one also obtains the action density in the form

\displaystyle\mathcal{I} =116\upiω2𝑬˘(ω2ϵH(ω,𝒌))ω𝑬˘|ω=w(𝒌¯)\displaystyle=\frac{1}{16\upi\omega^{2}}\,\smash{{\breve{{\boldsymbol{E}}}}}^{\dagger}\,\frac{\partial(\omega^{2}{\boldsymbol{\epsilon}}_{\text{H}}(\omega,{\boldsymbol{k}}))}{\partial\omega}\,{\breve{{\boldsymbol{E}}}}\,\Big{|}_{\omega=w(\overline{{\boldsymbol{k}}})}
=116\upiω(𝑬˘(ωϵH(ω,𝒌))ω𝑬˘+𝑩˘𝑩˘)|ω=w(𝒌¯),\displaystyle=\frac{1}{16\upi\omega}\left(\smash{{\breve{{\boldsymbol{E}}}}}^{\dagger}\,\frac{\partial(\omega{\boldsymbol{\epsilon}}_{\text{H}}(\omega,{\boldsymbol{k}}))}{\partial\omega}\,{\breve{{\boldsymbol{E}}}}+\smash{{\breve{{\boldsymbol{B}}}}}^{\dagger}{\breve{{\boldsymbol{B}}}}\right)\Big{|}_{\omega=w(\overline{{\boldsymbol{k}}})}, (474)

where we used (464).

9.3 Newtonian gravity

For Newtonian interactions governed by a gravitostatic potential φg\smash{\varphi_{g}}, one has

Hs=p22ms+msφ¯g+msφ~g,𝔏0=(φ~g)28\upi𝖦,\displaystyle H_{s}=\frac{p^{2}}{2m_{s}}+m_{s}\overline{\varphi}_{g}+m_{s}\widetilde{\varphi}_{g},\qquad\mathfrak{L}_{0}=-\frac{(\nabla\tilde{\varphi}_{g})^{2}}{8\upi\mathsf{G}}, (475)

where 𝖦\smash{\mathsf{G}} is the gravitational constant. This system is identical to that considered in section 9.1 for nonrelativistic electrostatic interactions up to coefficients. Specifically, es\smash{e_{s}} are replaced with ms\smash{m_{s}}, a factor 𝖦1\smash{-\mathsf{G}^{-1}} appears in 𝚵0\smash{{\boldsymbol{\Xi}}_{0}}, and the dispersion matrix becomes

𝚵(ω,𝒌)=Ξ(ω,𝒌)=k2ϵg(ω,𝒌)4\upi𝖦.\displaystyle{\boldsymbol{\Xi}}(\omega,{\boldsymbol{k}})=\Xi(\omega,{\boldsymbol{k}})=-\frac{k^{2}\epsilon_{g}(\omega,{\boldsymbol{k}})}{4\upi\mathsf{G}}. (476)

Thus, ϵ\smash{\epsilon_{\parallel}} is replaced with ϵg/𝖦\smash{-\epsilon_{g}/\mathsf{G}}, where ϵg\smash{\epsilon_{g}} is the gravitostatic permittivity given by

ϵg(ω,𝒌)=1s4\upi𝖦ms2k2d𝒑𝒌ω𝒌𝒗s+i0Fs𝒑.\displaystyle\epsilon_{g}(\omega,{\boldsymbol{k}})=1-\sum_{s}\frac{4\upi\mathsf{G}m_{s}^{2}}{k^{2}}\int\mathrm{d}{\boldsymbol{p}}\,\frac{{\boldsymbol{k}}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\mathrm{i}0}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}. (477)

This readily yields, for example, kinetic theory of the Jeans instability (Trigger et al., 2004), whose dispersion relation is given by ϵg(ω,𝒌)=0\smash{\epsilon_{g}(\omega,{\boldsymbol{k}})=0} (modulo the usual analytic continuation of the permittivity to modes with imω<0\smash{\operatorname{im}\omega<0}).

9.4 Relativistic gravity

9.4.1 Main equations

The dynamics of a relativistic neutral particle with mass m\smash{m} in a spacetime metric gαβ\smash{g_{\alpha\beta}} with signature (+++)\smash{(-+++)} is governed by a covariant Hamiltonian (see, for example, (Garg & Dodin, 2020))

𝖧(𝘅,𝗽)=12m(m2+gαβ(𝘅)pαpβ)𝖧(𝒈,𝗽).\displaystyle\mathsf{H}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{p}}})=\frac{1}{2m}\left(m^{2}+g^{\alpha\beta}({\boldsymbol{\mathsf{x}}})p_{\alpha}p_{\beta}\right)\equiv\mathsf{H}({\boldsymbol{g}},{\boldsymbol{\mathsf{p}}}). (478)

Here, 𝘅(x0,𝒙)\smash{{\boldsymbol{\mathsf{x}}}\equiv(x^{0},{\boldsymbol{x}})}, and x0=t\smash{x^{0}=t}, as usual. Also, 𝗽(p0,𝒑)\smash{{\boldsymbol{\mathsf{p}}}\equiv(p_{0},{\boldsymbol{p}})} is the index-free notation for the four-momentum pα\smash{p_{\alpha}}, gαβ\smash{g^{\alpha\beta}} is the inverse metric, 𝒈\smash{{\boldsymbol{g}}} is the index-free notation for gαβ\smash{g^{\alpha\beta}}, the units are such that c=1\smash{c=1}, and the species index is omitted.414141This section uses notation different from that used in the rest of the paper. In particular, the Greek indices span from 0 to 3, and the standard rules of index manipulations apply. The corresponding Hamilton’s equations, with τ\smash{\tau} the proper time, are

dxαdτ=𝖧pα,dpαdτ=𝖧xα.\displaystyle\frac{\mathrm{d}x^{\alpha}}{\mathrm{d}\tau}=\frac{\partial\mathsf{H}}{\partial p_{\alpha}},\qquad\frac{\mathrm{d}p_{\alpha}}{\mathrm{d}\tau}=-\frac{\partial\mathsf{H}}{\partial x^{\alpha}}. (479)

This dynamics is constrained to the shell p0=P0(t,𝒙,𝒑)\smash{p_{0}=P_{0}(t,{\boldsymbol{x}},{\boldsymbol{p}})}, where P0\smash{P_{0}} is the (negative) solution of

𝖧(𝒈,P0(t,𝒙,𝒑),𝒑)=0.\displaystyle\mathsf{H}({\boldsymbol{g}},P_{0}(t,{\boldsymbol{x}},{\boldsymbol{p}}),{\boldsymbol{p}})=0. (480)

This means that the particle distribution in the (𝘅,𝗽)\smash{({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{p}}})} space is delta-shaped and thus does not satisfy (119). Hence, we will consider particles in the six-dimensional space (𝒙,𝒑)\smash{({\boldsymbol{x}},{\boldsymbol{p}})} instead. The corresponding dynamics is governed by the Hamiltonian

H=P0(t,𝒙,𝒑).\displaystyle H=-P_{0}(t,{\boldsymbol{x}},{\boldsymbol{p}}). (481)

This is seen from the fact that

H=P0=𝖧/𝖧/p0,\displaystyle\frac{\partial H}{\partial\sqbullet}=-\frac{\partial P_{0}}{\partial\sqbullet}=\frac{\partial\mathsf{H}/\partial\sqbullet}{\partial\mathsf{H}/\partial p_{0}}, (482)

where {t,𝒙,𝒑}\smash{\sqbullet\in\{t,{\boldsymbol{x}},{\boldsymbol{p}}\}}, so Hamilton’s equations obtained from (481) are equivalent to (479):

dxαdt=Hpα=𝖧/pα𝖧/p0,dpαdt=Hxα=𝖧/xα𝖧/p0.\displaystyle\frac{\mathrm{d}x^{\alpha}}{\mathrm{d}t}=\frac{\partial H}{\partial p_{\alpha}}=\frac{\partial\mathsf{H}/\partial p_{\alpha}}{\partial\mathsf{H}/\partial p_{0}},\qquad\frac{\mathrm{d}p_{\alpha}}{\mathrm{d}t}=-\frac{\partial H}{\partial x^{\alpha}}=\frac{\partial\mathsf{H}/\partial x^{\alpha}}{\partial\mathsf{H}/\partial p_{0}}. (483)

Let us decompose the metric into the average part and oscillations, gαβ=g¯αβ+g~αβ\smash{g_{\alpha\beta}=\overline{g}_{\alpha\beta}+\widetilde{g}_{\alpha\beta}}, and approximate the inverse metric to the second order in 𝒈~\smash{\widetilde{{\boldsymbol{g}}}}:

gαβg¯αβg~αβ+g~αγg¯γδg~δβ,\displaystyle g^{\alpha\beta}\approx\overline{g}^{\alpha\beta}-\widetilde{g}^{\alpha\beta}+\widetilde{g}^{\alpha\gamma}\overline{g}_{\gamma\delta}\widetilde{g}^{\delta\beta}, (484)

where the indices of 𝒈~\smash{\widetilde{{\boldsymbol{g}}}} are manipulated using the background metric g¯αβ\smash{\overline{g}_{\alpha\beta}}. This gives

𝖧=12m(m2+g¯αβpαpβg~αβpαpβ+g~αβg¯βγg~γδpαpδ).\displaystyle\mathsf{H}=\frac{1}{2m}\left(m^{2}+\overline{g}^{\alpha\beta}p_{\alpha}p_{\beta}-\widetilde{g}^{\alpha\beta}p_{\alpha}p_{\beta}+\widetilde{g}^{\alpha\beta}\overline{g}_{\beta\gamma}\widetilde{g}^{\gamma\delta}p_{\alpha}p_{\delta}\right). (485)

The Hamiltonian (481) is expanded in 𝒈~\smash{\widetilde{{\boldsymbol{g}}}} as follows:

H(𝒈,𝗽)P0P0g~αβg~αβ122P0g~αβg~γδg~αβg~γδ,\displaystyle H({\boldsymbol{g}},{\boldsymbol{\mathsf{p}}})\approx-P_{0}-\frac{\partial P_{0}}{\partial\widetilde{g}^{\alpha\beta}}\,\widetilde{g}^{\alpha\beta}-\frac{1}{2}\,\frac{\partial^{2}P_{0}}{\partial\widetilde{g}^{\alpha\beta}\partial\widetilde{g}^{\gamma\delta}}\,\widetilde{g}^{\alpha\beta}\widetilde{g}^{\gamma\delta}, (486)

where P0\smash{P_{0}} and the derivatives on the right-hand side are evaluated on (𝒈¯,𝗽)\smash{(\overline{{\boldsymbol{g}}},{\boldsymbol{\mathsf{p}}})}. To calculate these derivatives, let us differentiate (480) and use (485) for 𝖧\smash{\mathsf{H}}. This gives

0=𝖧g~αβ+𝖧p0P0g~αβ=12m(pαpβ+2P0P0g~αβ),\displaystyle 0=\frac{\partial\mathsf{H}}{\partial\widetilde{g}^{\alpha\beta}}+\frac{\partial\mathsf{H}}{\partial p_{0}}\frac{\partial P_{0}}{\partial\widetilde{g}^{\alpha\beta}}=\frac{1}{2m}\left(-p_{\alpha}p_{\beta}+2P^{0}\frac{\partial P_{0}}{\partial\widetilde{g}^{\alpha\beta}}\right), (487)

where the derivatives with respect to the oscillating metric are taken at fixed pα\smash{p_{\alpha}} and at 𝒈~0\smash{\widetilde{{\boldsymbol{g}}}\to 0}, and P0P0(𝒈,𝗽)=g¯0αpα\smash{P^{0}\equiv P^{0}({\boldsymbol{g}},{\boldsymbol{\mathsf{p}}})=\overline{g}^{0\alpha}p_{\alpha}}; thus,

P0g~αβ=pαpβ2P0.\displaystyle\frac{\partial P_{0}}{\partial\widetilde{g}^{\alpha\beta}}=\frac{p_{\alpha}p_{\beta}}{2P^{0}}. (488)

Similarly, differentiating (480) twice gives

0\displaystyle 0 =2𝖧g~αβg~γδ+𝖧p02P0g~αβg~γδ+P0g~αβp0𝖧g~γδ+p0(𝖧g~αβ+𝖧p0P0g~αβ)P0gγδ\displaystyle=\frac{\partial^{2}\mathsf{H}}{\partial\widetilde{g}^{\alpha\beta}\partial\widetilde{g}^{\gamma\delta}}+\frac{\partial\mathsf{H}}{\partial p_{0}}\frac{\partial^{2}P_{0}}{\partial\widetilde{g}^{\alpha\beta}\partial\widetilde{g}^{\gamma\delta}}+\frac{\partial P_{0}}{\partial\widetilde{g}^{\alpha\beta}}\frac{\partial}{\partial p_{0}}\frac{\partial\mathsf{H}}{\partial\widetilde{g}^{\gamma\delta}}+\frac{\partial}{\partial p_{0}}\left(\frac{\partial\mathsf{H}}{\partial\widetilde{g}^{\alpha\beta}}+\frac{\partial\mathsf{H}}{\partial p_{0}}\frac{\partial P_{0}}{\partial\widetilde{g}^{\alpha\beta}}\right)\frac{\partial P_{0}}{\partial g^{\gamma\delta}}
=2𝖧g~αβg~γδ+𝖧p02P0g~αβg~γδ+P0g~αβp0𝖧g~γδ+P0g~γδp0𝖧g~αβ+2𝖧p0p0P0g~αβP0g~γδ\displaystyle=\frac{\partial^{2}\mathsf{H}}{\partial\widetilde{g}^{\alpha\beta}\partial\widetilde{g}^{\gamma\delta}}+\frac{\partial\mathsf{H}}{\partial p_{0}}\frac{\partial^{2}P_{0}}{\partial\widetilde{g}^{\alpha\beta}\partial\widetilde{g}^{\gamma\delta}}+\frac{\partial P_{0}}{\partial\widetilde{g}^{\alpha\beta}}\frac{\partial}{\partial p_{0}}\frac{\partial\mathsf{H}}{\partial\widetilde{g}^{\gamma\delta}}+\frac{\partial P_{0}}{\partial\widetilde{g}^{\gamma\delta}}\frac{\partial}{\partial p_{0}}\frac{\partial\mathsf{H}}{\partial\widetilde{g}^{\alpha\beta}}+\frac{\partial^{2}\mathsf{H}}{\partial p_{0}\partial p_{0}}\frac{\partial P_{0}}{\partial\widetilde{g}^{\alpha\beta}}\frac{\partial P_{0}}{\partial\widetilde{g}^{\gamma\delta}}
=12m(g¯βγpαpδ+g¯δαpγpβ+2P02P0g~αβg~γδ12P0(pαpβpγpδ)p0+g¯00pαpβpγpδ2(P0)2),\displaystyle=\frac{1}{2m}\left(\overline{g}_{\beta\gamma}p_{\alpha}p_{\delta}+\overline{g}_{\delta\alpha}p_{\gamma}p_{\beta}+2P^{0}\frac{\partial^{2}P_{0}}{\partial\widetilde{g}^{\alpha\beta}\partial\widetilde{g}^{\gamma\delta}}-\frac{1}{2P^{0}}\frac{\partial(p_{\alpha}p_{\beta}p_{\gamma}p_{\delta})}{\partial p_{0}}+\overline{g}^{00}\,\frac{p_{\alpha}p_{\beta}p_{\gamma}p_{\delta}}{2(P^{0})^{2}}\right),

whence

2P0g~αβg~γδ=12P0(g¯βγpαpδ+g¯δαpγpβ)+14(P0)2(pαpβpγpδ)p0g¯00pαpβpγpδ4(P0)3.\displaystyle\frac{\partial^{2}P_{0}}{\partial\widetilde{g}^{\alpha\beta}\partial\widetilde{g}^{\gamma\delta}}=-\frac{1}{2P^{0}}\,(\overline{g}_{\beta\gamma}p_{\alpha}p_{\delta}+\overline{g}_{\delta\alpha}p_{\gamma}p_{\beta})+\frac{1}{4(P^{0})^{2}}\frac{\partial(p_{\alpha}p_{\beta}p_{\gamma}p_{\delta})}{\partial p_{0}}-\overline{g}^{00}\,\frac{p_{\alpha}p_{\beta}p_{\gamma}p_{\delta}}{4(P^{0})^{3}}.

Then, (486) yields

HH0+ααβg~αβ+12g~αβαβg~γδγδ,\displaystyle H\approx H_{0}+\alpha_{\alpha\beta}\,\widetilde{g}^{\alpha\beta}+\frac{1}{2}\,\widetilde{g}_{\alpha\beta}\wp^{\alpha\beta}{}_{\gamma\delta}\widetilde{g}^{\gamma\delta}, (489)

where we introduced H0=P0\smash{H_{0}=-P_{0}}, ααβ=pαpβ/(2P0)\smash{\alpha^{\alpha\beta}=p^{\alpha}p^{\beta}/(2P^{0})}, and

αβ=γδδγβpαpδ+δδαpβpγ2P014(P0)2(pαpβpγpδ)p0+g¯00pαpβpγpδ4(P0)3.\displaystyle\wp^{\alpha\beta}{}_{\gamma\delta}=\frac{\delta^{\beta}_{\gamma}p^{\alpha}p_{\delta}+\delta_{\delta}^{\alpha}p^{\beta}p_{\gamma}}{2P^{0}}-\frac{1}{4(P^{0})^{2}}\frac{\partial(p^{\alpha}p^{\beta}p_{\gamma}p_{\delta})}{\partial p_{0}}+\overline{g}^{00}\,\frac{p^{\alpha}p^{\beta}p_{\gamma}p_{\delta}}{4(P^{0})^{3}}. (490)

9.4.2 Nonlinear potentials

Let us treat 𝒈~\smash{\widetilde{{\boldsymbol{g}}}} as a 16-dimensional vector (Garg & Dodin, 2021b), so ααβ\smash{\alpha_{\alpha\beta}} serves as 𝜶\smash{{\boldsymbol{\alpha}}^{\dagger}} and αβγδ\smash{\wp^{\alpha\beta}{}_{\gamma\delta}} serves as \smash{{\boldsymbol{\wp}}}. (Because these operators happen to be local in the 𝘅\smash{{\boldsymbol{\mathsf{x}}}} representation, here we do not distinguish them from their symbols.) Let us also introduce

𝔈pαpβpγpδ𝖴αβγδ\displaystyle\mathfrak{E}\doteq p_{\alpha}p_{\beta}p_{\gamma}p_{\delta}\mathsf{U}^{\alpha\beta\gamma\delta} (491)

and notice that vix˙i=pi/p0\smash{v^{i}\approx\dot{x}^{i}=p^{i}/p^{0}} (see (483)), so ω𝒌𝒗=kρpρ/P0\smash{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}=-k_{\rho}p^{\rho}/P^{0}} and δ(ω𝒌𝒗)=P0δ(kρpρ)\smash{\delta(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}})}=\smash{P^{0}\delta(k^{\rho}p_{\rho})}. Then, one finds from (284) that (appendix B.8)

𝗗\displaystyle{\boldsymbol{\mathsf{D}}} =\upi4P0d𝗸𝒌𝒌𝔈δ(kρpρ),\displaystyle=\frac{\upi}{4P^{0}}\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,{\boldsymbol{k}}{\boldsymbol{k}}\mathfrak{E}\,\delta(k^{\rho}p_{\rho}), (492)
𝚯\displaystyle{\boldsymbol{\Uptheta}} =14P0ϑd𝗸𝒌𝒌𝔈ϑP0kρpρ|ϑ=0,\displaystyle=\frac{1}{4P^{0}}\frac{\partial}{\partial\vartheta}\fint\mathrm{d}{\boldsymbol{\mathsf{k}}}\left.\frac{{\boldsymbol{k}}{\boldsymbol{k}}\mathfrak{E}}{\vartheta P^{0}-k^{\rho}p_{\rho}}\right|_{\vartheta=0}, (493)
Δ\displaystyle\Delta =pαpβ2P0d𝗸𝖴αγγβ18P0pλd𝗸kλ𝔈kρpρ.\displaystyle=\frac{p_{\alpha}p_{\beta}}{2P^{0}}\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\mathsf{U}^{\alpha}{}_{\gamma}{}^{\gamma\beta}-\frac{1}{8P^{0}}\frac{\partial}{\partial p_{\lambda}}\fint\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\frac{k_{\lambda}\mathfrak{E}}{k^{\rho}p_{\rho}}. (494)

Equation (494) (where one takes p0=P0\smash{p_{0}=P_{0}} after the differentiation) is in agreement with the result that was obtained for quasimonochromatic waves in (Garg & Dodin, 2020). The derivation of the dispersion matrix 𝚵\smash{{\boldsymbol{\Xi}}} for relativistic gravitational interactions in matter is cumbersome, so it is not presented here, but see (Garg & Dodin, 2022). The collision integral and fluctuations for relativistic gravitational interactions are straightforward to obtain from the general formulas presented in sections 6.9 and 8. This can be used to describe QL interactions of gravitational waves, including not only the usual vacuum modes424242Vacuum gravitational waves satisfy ω2=|𝒌|2\smash{\omega^{2}=|{\boldsymbol{k}}|^{2}}. Hence, satisfying the resonance condition kρpρ=0\smash{k^{\rho}p_{\rho}=0} requires |𝒌𝒗|=|𝒌|\smash{|{\boldsymbol{k}}\cdot{\boldsymbol{v}}|=|{\boldsymbol{k}}|}, which requires particle speeds not smaller than the speed of light (remember that c=1\smash{c=1} in our units). For massive particles, this cannot be satisfied, so 𝗗\smash{{\boldsymbol{\mathsf{D}}}} vanishes for vacuum gravitational waves. However, such waves can still produce adiabatic ponderomotive effects determined by Δ\smash{\Delta} (Garg & Dodin, 2020). but also waves coupled with matter, for example, the relativistic Jeans mode.

Also notice that the OC Hamiltonian =H0+Δ\smash{\mathcal{H}=H_{0}+\Delta} can be put in a covariant form as follows. Like in the original system (section 9.4.1), \smash{\mathcal{H}} determines the ponderomotively modified shell p0=𝒫0(t,𝒙,𝒑)\smash{p_{0}=\mathcal{P}_{0}(t,{\boldsymbol{x}},{\boldsymbol{p}})} via =𝒫0\smash{\mathcal{H}=-\mathcal{P}_{0}}. On one hand, the covariant OC Hamiltonian \smash{\mathcal{H}^{\prime}} vanishes on this shell,434343The covariant Hamiltonian is the dispersion function of particles as quantum waves in the semiclassical limit (Garg & Dodin, 2020). so it can be Taylor-expanded as follows:

(p0𝒫0)p0|p=𝒫0(p0P0+Δ)λ,λ𝖧p0|p=P0.\displaystyle\mathcal{H}^{\prime}\approx(p_{0}-\mathcal{P}_{0})\,\frac{\partial\mathcal{H}}{\partial p_{0}}\bigg{|}_{p=\mathcal{P}_{0}}\approx(p_{0}-P_{0}+\Delta)\lambda,\qquad\lambda\doteq\frac{\partial\mathsf{H}}{\partial p_{0}}\bigg{|}_{p=P_{0}}. (495)

On the other hand, it can also be represented as =𝖧(𝒈¯,𝗽)+Δ\smash{\mathcal{H}^{\prime}=\mathsf{H}(\overline{{\boldsymbol{g}}},{\boldsymbol{\mathsf{p}}})+\Delta^{\prime}} (here Δ\smash{\Delta^{\prime}} is the ponderomotive term yet to be found) and Taylor-expanded around the unperturbed shell p0=P0(t,𝒙,𝒑)\smash{p_{0}=P_{0}(t,{\boldsymbol{x}},{\boldsymbol{p}})} as

Δ+(p0P0)λ=(p0P0+Δ/λ)λ.\displaystyle\mathcal{H}^{\prime}\approx\Delta^{\prime}+(p_{0}-P_{0})\lambda=(p_{0}-P_{0}+\Delta^{\prime}/\lambda)\lambda. (496)

By comparing (495) with (496), one finds that Δ=λΔ\smash{\Delta^{\prime}=\lambda\Delta}. Because λ=P0/m\smash{\lambda=P^{0}/m}, this leads to the following covariant Hamiltonian for OCs:

=12m(m2+geffαβpαpβ14pλd𝗸kλ𝔈kρpρ),\displaystyle\displaystyle\mathcal{H}^{\prime}=\frac{1}{2m}\left(m^{2}+g_{\rm eff}^{\alpha\beta}p_{\alpha}p_{\beta}-\frac{1}{4}\frac{\partial}{\partial p_{\lambda}}\fint\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\frac{k_{\lambda}\mathfrak{E}}{k^{\rho}p_{\rho}}\right), (497)
geffαβg¯αβ+d𝗸𝖴α.γγβ\displaystyle\displaystyle g_{\rm eff}^{\alpha\beta}\doteq\overline{g}^{\alpha\beta}+\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\mathsf{U}^{\alpha}{}_{\gamma}{}^{\gamma\beta}. (498)

9.4.3 Gauge invariance

As shown in (Garg & Dodin, 2021a, b) adiabatic QL interactions via gravitational waves (i.e. those determined by 𝚯\smash{{\boldsymbol{\Uptheta}}} and Δ\smash{\Delta}) can be formulated in a form invariant with respect to gauge transformations

g~αβg~αβ=g~αβ+(αξ~β),\displaystyle\widetilde{g}^{\alpha\beta}\to\widetilde{g}^{\prime\alpha\beta}=\widetilde{g}^{\alpha\beta}+\nabla^{(\alpha}\widetilde{\xi}^{\beta)}, (499)

where \smash{\nabla} is the covariant derivative associated with the background metric 𝒈¯\smash{\overline{{\boldsymbol{g}}}}, ξ~μ\smash{\widetilde{\xi}^{\mu}} is an arbitrary vector field, and ψ(αηβ)(ψαηβ+ψβηα)/2\smash{\psi^{(\alpha}\eta^{\beta)}}\equiv\smash{(\psi^{\alpha}\eta^{\beta}+\psi^{\beta}\eta^{\alpha})/2}. Let us show that this also extends to resonant interactions. Recall that within the assumed accuracy the nonlinear potentials are supposed to be calculated only to the zeroth order in the geometrical-optics parameter. Then, the modification of the average Wigner matrix of the metric oscillations under the transformation (499) can be written as

𝖴αβγδ𝖴αβγδ\displaystyle\mathsf{U}^{\prime\alpha\beta\gamma\delta}-\mathsf{U}^{\alpha\beta\gamma\delta} =symb𝗑(ik^(α|ξ~β)g~γδ|¯i|g~αβξ~(γ|¯k^δ)+k^(α|ξ~β)ξ~(γ|¯k^δ))\displaystyle=\text{symb}_{\mathsf{x}}\Big{(}\mathrm{i}\widehat{k}^{(\alpha}\,\overline{\ket{\widetilde{\xi}^{\beta)}}\bra{\widetilde{g}^{\gamma\delta}}}-\mathrm{i}\overline{\ket{\widetilde{g}^{\alpha\beta}}\bra{\widetilde{\xi}^{(\gamma}}}\,\widehat{k}^{\delta)}+\widehat{k}^{(\alpha}\overline{\ket{\widetilde{\xi}^{\beta)}}\bra{\widetilde{\xi}^{(\gamma}}}\,\widehat{k}^{\delta)}\Big{)}
=ik(α𝒲β)γδik(δ𝒲γ)αβ+k(α𝖶ξ~β)(γkδ),\displaystyle=\mathrm{i}k^{(\alpha}\mathcal{W}^{\beta)\gamma\delta}-\mathrm{i}k^{(\delta}\mathcal{W}^{\gamma)\alpha\beta*}+k^{(\alpha}\mathsf{W}_{\widetilde{\xi}}^{\beta)(\gamma}k^{\delta)}, (500)

where 𝒲βγδsymb𝗑|ξ~βg~γδ|¯\smash{\mathcal{W}^{\beta\gamma\delta}\doteq\overline{\text{symb}_{\mathsf{x}}\ket{\widetilde{\xi}^{\beta}}\bra{\widetilde{g}^{\gamma\delta}}}} and 𝖶ξ~βγ\smash{\mathsf{W}_{\widetilde{\xi}}^{\beta\gamma}} is the average Wigner matrix of ξ~α\smash{\widetilde{\xi}^{\alpha}}. The corresponding change of 𝔈\smash{\mathfrak{E}} is

𝔈𝔈=(kρpρ)(ipβpγpδ𝒲βγδipαpβpγ𝒲γαβ+kλpλpβpγ𝖶ξ~βγ)(kρpρ)A.\displaystyle\mathfrak{E}^{\prime}-\mathfrak{E}=(k^{\rho}p_{\rho})\Big{(}\mathrm{i}p_{\beta}p_{\gamma}p_{\delta}\mathcal{W}^{\beta\gamma\delta}-\mathrm{i}p_{\alpha}p_{\beta}p_{\gamma}\mathcal{W}^{\gamma\alpha\beta*}+k^{\lambda}p_{\lambda}p_{\beta}p_{\gamma}\mathsf{W}_{\widetilde{\xi}}^{\beta\gamma}\Big{)}\equiv(k^{\rho}p_{\rho})A.

Then, the difference in the diffusion coefficients (492) is

𝗗𝗗=\upi4P0d𝗸𝒌𝒌δ(kρpρ)(kρpρ)A=0,\displaystyle{\boldsymbol{\mathsf{D}}}^{\prime}-{\boldsymbol{\mathsf{D}}}=\frac{\upi}{4P^{0}}\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,{\boldsymbol{k}}{\boldsymbol{k}}\,\delta(k^{\rho}p_{\rho})\,(k^{\rho}p_{\rho})\,A=0, (501)

because δ(kρpρ)(kρpρ)=0\smash{\delta(k^{\rho}p_{\rho})\,(k^{\rho}p_{\rho})=0}. In particular, this rules out QL diffusion via ‘coordinate waves’.

9.4.4 Lorenz gauge and effective metric

Let us consider gravitational waves in the Lorenz gauge, αg~αβ=0\smash{\nabla_{\alpha}\widetilde{g}^{\alpha\beta}=0}. In this case,

kα𝖴αβγδ=kβ𝖴αβγδ=kγ𝖴αβγδ=kδ𝖴αβγδ=0,\displaystyle k_{\alpha}\mathsf{U}^{\alpha\beta\gamma\delta}=k_{\beta}\mathsf{U}^{\alpha\beta\gamma\delta}=k_{\gamma}\mathsf{U}^{\alpha\beta\gamma\delta}=k_{\delta}\mathsf{U}^{\alpha\beta\gamma\delta}=0, (502)

and thus kλ𝔈/pλ=0\smash{k_{\lambda}\partial\mathfrak{E}/\partial p^{\lambda}=0}. Then,

pλd𝗸kλ𝔈kρpρ=ϑd𝗸(kλkλ)𝔈kρpρ+ϑ|ϑ=0.\displaystyle\frac{\partial}{\partial p_{\lambda}}\fint\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\frac{k_{\lambda}\mathfrak{E}}{k^{\rho}p_{\rho}}=\frac{\partial}{\partial\vartheta}\fint\mathrm{d}{\boldsymbol{\mathsf{k}}}\left.\frac{(k_{\lambda}k^{\lambda})\mathfrak{E}}{k^{\rho}p_{\rho}+\vartheta}\right|_{\vartheta=0}. (503)

This simplifies the expression (494) for Δ\smash{\Delta} and (497) for \smash{\mathcal{H}^{\prime}}. Furthermore, if the waves are not significantly affected by matter, so the vacuum dispersion kλkλ=0\smash{k_{\lambda}k^{\lambda}=0} can be assumed, the term (503) vanishes completely. Then, (497) becomes

=12m(m2+geffαβpαpβ)\displaystyle\mathcal{H}^{\prime}=\frac{1}{2m}\left(m^{2}+g_{\rm eff}^{\alpha\beta}p_{\alpha}p_{\beta}\right) (504)

and QL diffusion disappears, because particles cannot resonate with waves. This shows that the only average QL effect of vacuum gravitational waves on particles is the modification of the spacetime metric by d𝗸𝖴αγ=γβ𝒪(ε2)\smash{\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\mathsf{U}_{\alpha\gamma}{}^{\gamma}{}_{\beta}=\mathcal{O}(\varepsilon^{2})}. For quasimonochromatic waves, this effect is discussed in further detail in (Garg & Dodin, 2020).

10 Summary

In summary, we have presented quasilinear theory for classical plasma interacting with inhomogeneous turbulence in the presence of background fields. Because we use the Weyl symbol calculus, global-mode decomposition is not invoked, so our formulation is local and avoids the usual issues with complex-frequency modes. Also, the particle Hamiltonian is kept general, so the results are equally applicable to relativistic, electromagnetic, and even non-electromagnetic (for example, gravitational) interactions. Because our approach is not bounded by the limitations of variational analysis either, effects caused by collisional and collisionless dissipation are also included naturally.

Our main results are summarized in sections 5.6, 6.9, 7.6, 8.5 and are as follows. Starting from the Klimontovich equation, we derive a Fokker–Planck model for the dressed oscillation-center distribution. This model captures quasilinear diffusion, interaction with the background fields, and ponderomotive effects simultaneously. The local diffusion coefficient is manifestly positive-semidefinite. Waves are allowed to be off-shell (not constrained by a dispersion relation), and a collision integral of the Balescu–Lenard type emerges in a form that is not restricted to any particular Hamiltonian. This operator conserves particles, momentum, and energy, and it also satisfies the H\smash{H}-theorem, as usual. As a spin-off, a general expression for the spectrum (average Wigner matrix) of microscopic fluctuations of the interaction field is derived. For on-shell waves, which satisfy a quasilinear wave-kinetic equation, our theory conserves the momentum and energy of the wave–plasma system. Dewar’s oscillation-center quasilinear theory of electrostatic turbulence (Dewar, 1973) is proven formally as a particular case and given a concise formulation. Also discussed as examples are relativistic electromagnetic and gravitational interactions, and quasilinear theory for gravitational waves is proposed.

Aside from having the aesthetic appeal of a rigorous local theory, our formulation can help, for example, better understand and model quasilinear plasma heating and current drive. First of all, it systematically accounts for the wave-driven evolution of the nonresonant-particle distribution and for the ponderomotive effects caused by plasma inhomogeneity in both time and space. As discussed above (section 7.5), this is generally important for adequately calculating the energy–momentum transfer between waves and plasma even when resonant absorption per se occurs in a homogeneous-plasma region. Second, our formulation provides general formulas that equally hold in any canonical variables and for any Hamiltonians that satisfy our basic assumptions (section 3.1). Therefore, our results can be applied to various plasma models immediately. This eliminates the need for ad hoc calculations, which can be especially cumbersome beyond the homogeneous-plasma approximation. Discussing specific models of applied interest, however exciting, is beyond the scope of this paper and is left to future work.

Funding

This work was supported by the U.S. DOE through Contract DE-AC02-09CH11466. It is also based upon the work supported by National Science Foundation under the grant No. PHY 1903130.

Declaration of interests

The author reports no conflict of interest.

Appendix A Average Wigner matrices

A.1 Positive semidefinitness

As known from (Cartwright, 1976), the average Wigner function of any scalar field on the real axis is non-negative if the averaging is done over a sufficiently large phase-space volume. Here, we extend this theorem to average Wigner matrices of vector fields in a multi-dimensional space, 𝝍(𝘅)\smash{{\boldsymbol{\psi}}({\boldsymbol{\mathsf{x}}})}, and show that such matrices are positive-semidefinite.

For any given function h(𝒛)h(𝘅,𝗸)\smash{h({\boldsymbol{z}})\equiv h({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})}, we define its local phase-space average as the following convolution integral:444444This ensures that 𝒛h¯=𝒛h¯\smash{\partial_{\boldsymbol{z}}\overline{h}=\overline{\partial_{\boldsymbol{z}}h}}, as readily seen from (505) using integration by parts.

h¯(𝒛)d𝒛𝒢(𝒛𝒛)h(𝒛)d𝘅d𝗸𝒢(𝘅𝘅,𝗸𝗸)h(𝘅,𝗸)\displaystyle\overline{h}({\boldsymbol{z}})\doteq\int\mathrm{d}{\boldsymbol{z}}^{\prime}\,\mathcal{G}({\boldsymbol{z}}-{\boldsymbol{z}}^{\prime})\,h({\boldsymbol{z}}^{\prime})\equiv\int\mathrm{d}{\boldsymbol{\mathsf{x}}}^{\prime}\,\mathrm{d}{\boldsymbol{\mathsf{k}}}^{\prime}\,\mathcal{G}({\boldsymbol{\mathsf{x}}}-{\boldsymbol{\mathsf{x}}}^{\prime},{\boldsymbol{\mathsf{k}}}-{\boldsymbol{\mathsf{k}}}^{\prime})\,h({\boldsymbol{\mathsf{x}}}^{\prime},{\boldsymbol{\mathsf{k}}}^{\prime}) (505)

with a Gaussian window function

𝒢(𝘅,𝗸)1(2\upiσ𝗑σ𝗄)𝗇exp(|𝘅|22σ𝗑2|𝗸|22σ𝗄2)\displaystyle\mathcal{G}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\doteq\frac{1}{(2\upi\sigma_{\mathsf{x}}\sigma_{\mathsf{k}})^{\mathsf{n}}}\,\exp\left(-\frac{|{\boldsymbol{\mathsf{x}}}|^{2}}{2\sigma_{\mathsf{x}}^{2}}-\frac{|{\boldsymbol{\mathsf{k}}}|^{2}}{2\sigma_{\mathsf{k}}^{2}}\right) (506)

and positive constants σ𝗑\smash{\sigma_{\mathsf{x}}} and σ𝗄\smash{\sigma_{\mathsf{k}}} yet to be specified. Unlike in section 2.1.1, the following notation will be assumed for the ‘scalar product’ for variables with upper, lower, and mixed indices:

𝘅𝘅′′δij𝗑i𝗑′′j,𝗸𝗸′′δij𝗄i𝗄j′′,𝗸𝘅𝗄i𝗑i.\displaystyle{\boldsymbol{\mathsf{x}}}^{\prime}\cdot{\boldsymbol{\mathsf{x}}}^{\prime\prime}\doteq\delta_{ij}\mathsf{x}^{\prime i}\mathsf{x}^{\prime\prime j},\qquad{\boldsymbol{\mathsf{k}}}^{\prime}\cdot{\boldsymbol{\mathsf{k}}}^{\prime\prime}\doteq\delta^{ij}\mathsf{k}^{\prime}_{i}\mathsf{k}^{\prime\prime}_{j},\qquad{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{x}}}\doteq\mathsf{k}_{i}\mathsf{x}^{i}. (507)

(The Latin indices in this appendix range from 0 to n\smash{n}, δij\smash{\delta_{ij}} and δij\smash{\delta^{ij}} are unit matrices, and summation over repeating indices is assumed.) In particular, note that |𝘅|2𝘅𝘅0\smash{|{\boldsymbol{\mathsf{x}}}|^{2}\doteq{\boldsymbol{\mathsf{x}}}\cdot{\boldsymbol{\mathsf{x}}}\geq 0} and must not be confused with the squared spacetime interval, which can have either sign. Likewise, |𝗸|2𝗸𝗸0\smash{|{\boldsymbol{\mathsf{k}}}|^{2}\doteq{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{k}}}\geq 0} must not be confused with 𝗄i𝗄i=ω2+𝒌2\smash{\mathsf{k}_{i}\mathsf{k}^{i}=-\omega^{2}+{\boldsymbol{k}}^{2}}.

The average Wigner matrix of any given vector field 𝝍\smash{{\boldsymbol{\psi}}} is given by

𝗪¯𝝍(𝘅,𝗸)=1(2\upi)𝗇1(2\upiσ𝗑σ𝗄)𝗇\displaystyle\overline{{\boldsymbol{\mathsf{W}}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})=\frac{1}{(2\upi)^{\mathsf{n}}}\frac{1}{(2\upi\sigma_{\mathsf{x}}\sigma_{\mathsf{k}})^{\mathsf{n}}}\int d𝘀d𝘅d𝗸𝝍(𝘅+𝘀/2)𝝍(𝘅𝘀/2)\displaystyle\mathrm{d}{\boldsymbol{\mathsf{s}}}\,\mathrm{d}{\boldsymbol{\mathsf{x}}}^{\prime}\,\mathrm{d}{\boldsymbol{\mathsf{k}}}^{\prime}\,{\boldsymbol{\psi}}({\boldsymbol{\mathsf{x}}}^{\prime}+{\boldsymbol{\mathsf{s}}}/2){\boldsymbol{\psi}}^{\dagger}({\boldsymbol{\mathsf{x}}}^{\prime}-{\boldsymbol{\mathsf{s}}}/2)
×exp(|𝘅𝘅|22σ𝗑2|𝗸𝗸|22σ𝗄2i𝗸𝘀).\displaystyle\times\exp\left(-\frac{|{\boldsymbol{\mathsf{x}}}-{\boldsymbol{\mathsf{x}}}^{\prime}|^{2}}{2\sigma_{\mathsf{x}}^{2}}-\frac{|{\boldsymbol{\mathsf{k}}}-{\boldsymbol{\mathsf{k}}}^{\prime}|^{2}}{2\sigma_{\mathsf{k}}^{2}}-\mathrm{i}{\boldsymbol{\mathsf{k}}}^{\prime}\cdot{\boldsymbol{\mathsf{s}}}\right). (508)

The integral over 𝗸\smash{{\boldsymbol{\mathsf{k}}}^{\prime}} can be readily taken:

d𝗸exp(|𝗸𝗸|22σ𝗄2i𝗸𝘀)=(2\upi)𝗇/2σ𝗄𝗇exp(σ𝗄2|𝘀|22i𝗸𝘀).\displaystyle\int\mathrm{d}{\boldsymbol{\mathsf{k}}}^{\prime}\,\exp\left(-\frac{|{\boldsymbol{\mathsf{k}}}-{\boldsymbol{\mathsf{k}}}^{\prime}|^{2}}{2\sigma_{\mathsf{k}}^{2}}-\mathrm{i}{\boldsymbol{\mathsf{k}}}^{\prime}\cdot{\boldsymbol{\mathsf{s}}}\right)=(2\upi)^{{\mathsf{n}}/2}\sigma_{\mathsf{k}}^{\mathsf{n}}\exp\left(-\frac{\sigma_{\mathsf{k}}^{2}|{\boldsymbol{\mathsf{s}}}|^{2}}{2}-\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{s}}}\right). (509)

Then, using the variables 𝘅1,2𝘅±𝘀/2\smash{{\boldsymbol{\mathsf{x}}}_{1,2}\doteq{\boldsymbol{\mathsf{x}}}^{\prime}\pm{\boldsymbol{\mathsf{s}}}/2}, one can rewrite (508) as follows:

𝗪¯𝝍(𝘅,𝗸)=1(2\upi)3𝗇/2σ𝗑𝗇d𝘅1d𝘅2𝝍(𝘅1)𝝍(𝘅2)eϕ,\displaystyle\overline{{\boldsymbol{\mathsf{W}}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})=\frac{1}{(2\upi)^{3{\mathsf{n}}/2}\sigma_{\mathsf{x}}^{\mathsf{n}}}\int\mathrm{d}{\boldsymbol{\mathsf{x}}}_{1}\,\mathrm{d}{\boldsymbol{\mathsf{x}}}_{2}\,{\boldsymbol{\psi}}({\boldsymbol{\mathsf{x}}}_{1}){\boldsymbol{\psi}}^{\dagger}({\boldsymbol{\mathsf{x}}}_{2})\,\mathrm{e}^{-\phi}, (510)
ϕ=|𝘅(𝘅1+𝘅2)/2|22σ𝗑2+σ𝗄2|𝘅1𝘅2|22+i𝗸(𝘅1𝘅2).\displaystyle\phi=\frac{|{\boldsymbol{\mathsf{x}}}-({\boldsymbol{\mathsf{x}}}_{1}+{\boldsymbol{\mathsf{x}}}_{2})/2|^{2}}{2\sigma_{\mathsf{x}}^{2}}+\frac{\sigma_{\mathsf{k}}^{2}|{\boldsymbol{\mathsf{x}}}_{1}-{\boldsymbol{\mathsf{x}}}_{2}|^{2}}{2}+\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot({\boldsymbol{\mathsf{x}}}_{1}-{\boldsymbol{\mathsf{x}}}_{2}). (511)

The function ϕ\smash{\phi} can also be expressed as ϕ=|𝘅|2/(2σ𝗑2)+ϕ(𝘅1)+ϕ(𝘅2)λ𝘅1𝘅2\smash{\phi=|{\boldsymbol{\mathsf{x}}}|^{2}/(2\sigma_{\mathsf{x}}^{2})+\phi({\boldsymbol{\mathsf{x}}}_{1})+\phi^{*}({\boldsymbol{\mathsf{x}}}_{2})-\lambda{\boldsymbol{\mathsf{x}}}_{1}\cdot{\boldsymbol{\mathsf{x}}}_{2}}, where

ϕ(𝘆)|𝘆|22(14σ𝗑2+σ𝗄2)𝘅𝘆2σ𝗑2+i𝗸𝘆\displaystyle\phi({\boldsymbol{\mathsf{y}}})\doteq\frac{|{\boldsymbol{\mathsf{y}}}|^{2}}{2}\left(\frac{1}{4\sigma_{\mathsf{x}}^{2}}+\sigma_{\mathsf{k}}^{2}\right)-\frac{{\boldsymbol{\mathsf{x}}}\cdot{\boldsymbol{\mathsf{y}}}}{2\sigma_{\mathsf{x}}^{2}}+\mathrm{i}{\boldsymbol{\mathsf{k}}}\cdot{\boldsymbol{\mathsf{y}}} (512)

and λσ𝗄2(4σ𝗑2)1\smash{\lambda\doteq\sigma_{\mathsf{k}}^{2}-(4\sigma_{\mathsf{x}}^{2})^{-1}}. Then, using 𝝃(𝘆)𝝍(𝘆)eϕ(𝘆)\smash{{\boldsymbol{\xi}}({\boldsymbol{\mathsf{y}}})\doteq{\boldsymbol{\psi}}({\boldsymbol{\mathsf{y}}})\mathrm{e}^{-\phi({\boldsymbol{\mathsf{y}}})}}, one obtains from (510) that

𝗪¯𝝍(𝘅,𝗸)=e|𝘅|2/(2σ𝗑2)(2\upi)3𝗇/2σ𝗑𝗇d𝘅1d𝘅2𝝃(𝘅1)𝝃(𝘅2)eλ𝘅1𝘅2.\displaystyle\overline{{\boldsymbol{\mathsf{W}}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})=\frac{\mathrm{e}^{-|{\boldsymbol{\mathsf{x}}}|^{2}/(2\sigma_{\mathsf{x}}^{2})}}{(2\upi)^{3{\mathsf{n}}/2}\sigma_{\mathsf{x}}^{\mathsf{n}}}\int\mathrm{d}{\boldsymbol{\mathsf{x}}}_{1}\,\mathrm{d}{\boldsymbol{\mathsf{x}}}_{2}\,{\boldsymbol{\xi}}({\boldsymbol{\mathsf{x}}}_{1}){\boldsymbol{\xi}}^{\dagger}({\boldsymbol{\mathsf{x}}}_{2})\,\mathrm{e}^{\lambda{\boldsymbol{\mathsf{x}}}_{1}\cdot{\boldsymbol{\mathsf{x}}}_{2}}. (513)

By Taylor-expanding eλ𝘅1𝘅2\smash{\mathrm{e}^{\lambda{\boldsymbol{\mathsf{x}}}_{1}\cdot{\boldsymbol{\mathsf{x}}}_{2}}}, one obtains

𝗪¯𝝍(𝘅,𝗸)=e|𝘅|2/(2σ𝗑2)(2\upi)3𝗇/2σ𝗑𝗇m=0λmm!𝗝m,\displaystyle\overline{{\boldsymbol{\mathsf{W}}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})=\frac{\mathrm{e}^{-|{\boldsymbol{\mathsf{x}}}|^{2}/(2\sigma_{\mathsf{x}}^{2})}}{(2\upi)^{3{\mathsf{n}}/2}\sigma_{\mathsf{x}}^{\mathsf{n}}}\sum_{m=0}^{\infty}\frac{\lambda^{m}}{m!}\,{\boldsymbol{\mathsf{J}}}_{m}, (514)

where 𝗝md𝘅1d𝘅2(𝘅1𝘅2)m𝝃(𝘅1)𝝃(𝘅2)\smash{{\boldsymbol{\mathsf{J}}}_{m}\doteq\int\mathrm{d}{\boldsymbol{\mathsf{x}}}_{1}\,\mathrm{d}{\boldsymbol{\mathsf{x}}}_{2}\,({\boldsymbol{\mathsf{x}}}_{1}\cdot{\boldsymbol{\mathsf{x}}}_{2})^{m}{\boldsymbol{\xi}}({\boldsymbol{\mathsf{x}}}_{1}){\boldsymbol{\xi}}^{\dagger}({\boldsymbol{\mathsf{x}}}_{2})}. Note that

(𝘅1𝘅2)m=𝝁(m)i=1𝗇(𝗑1i𝗑2i)mi,\displaystyle({\boldsymbol{\mathsf{x}}}_{1}\cdot{\boldsymbol{\mathsf{x}}}_{2})^{m}=\sum_{{\boldsymbol{\mu}}(m)}\prod_{i=1}^{{\mathsf{n}}}(\mathsf{x}_{1}^{i}\mathsf{x}^{i}_{2})^{m_{i}}, (515)

where the summation is performed over all combinations 𝝁(m){m1,m2,,m𝗇}{\boldsymbol{\mu}}(m)\equiv\{m_{1},m_{2},\ldots,m_{\mathsf{n}}\} of integers mi0\smash{m_{i}\geq 0} such that imi=m\smash{\sum_{i}m_{i}=m}. Thus,

𝗝m=𝝁(m)𝓙𝝁𝓙𝝁,𝓙𝝁=d𝘆𝝃(𝘆)i=1𝗇(𝗒i)mi.\displaystyle{\boldsymbol{\mathsf{J}}}_{m}=\sum_{{\boldsymbol{\mu}}(m)}{\boldsymbol{\mathcal{J}}}_{{\boldsymbol{\mu}}}{\boldsymbol{\mathcal{J}}}^{\dagger}_{{\boldsymbol{\mu}}},\qquad{\boldsymbol{\mathcal{J}}}_{{\boldsymbol{\mu}}}=\int\mathrm{d}{\boldsymbol{\mathsf{y}}}\,{\boldsymbol{\xi}}({\boldsymbol{\mathsf{y}}})\prod_{i=1}^{{\mathsf{n}}}(\mathsf{y}^{i})^{m_{i}}. (516)

Because each 𝗝m\smash{{\boldsymbol{\mathsf{J}}}_{m}} is positive-semidefinite, the Wigner matrix 𝗪¯𝝍\smash{\overline{{\boldsymbol{\mathsf{W}}}}_{{\boldsymbol{\psi}}}} is positive-semidefinite when λ0\smash{\lambda\geq 0}, or equivalently, when σ𝗑σ𝗄>1/2\smash{\sigma_{\mathsf{x}}\sigma_{\mathsf{k}}>1/2}. This condition is assumed to be satisfied for the phase-space averaging of 𝗪𝝍\smash{{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}} used in the main text. Loosely, this means that the averaging is done over the phase-space volume Δ𝘅Δ𝗸(σ𝗑σ𝗄)𝗇1\smash{\Delta{\boldsymbol{\mathsf{x}}}\,\Delta{\boldsymbol{\mathsf{k}}}\sim(\sigma_{\mathsf{x}}\sigma_{\mathsf{k}})^{\mathsf{n}}\gtrsim 1}.

A.2 Invariant limit for eikonal fields

For eikonal fields (308), one has

𝝍(𝘅+𝘀/2)𝝍(𝘅𝘀/2)(𝑨(𝘅)ei𝗸¯(𝘅)𝘀+c.c.)+(𝑩(𝘅)e2iθ(𝘅)+c.c.).\displaystyle{\boldsymbol{\psi}}({\boldsymbol{\mathsf{x}}}+{\boldsymbol{\mathsf{s}}}/2){\boldsymbol{\psi}}^{\dagger}({\boldsymbol{\mathsf{x}}}-{\boldsymbol{\mathsf{s}}}/2)\approx\big{(}{\boldsymbol{A}}({\boldsymbol{\mathsf{x}}})\,\mathrm{e}^{\mathrm{i}\overline{{\boldsymbol{\mathsf{k}}}}({\boldsymbol{\mathsf{x}}})\cdot{\boldsymbol{\mathsf{s}}}}+\text{c.c.}\big{)}+\big{(}{\boldsymbol{B}}({\boldsymbol{\mathsf{x}}})\,\mathrm{e}^{2\mathrm{i}\theta({\boldsymbol{\mathsf{x}}})}+\text{c.c.}\big{)}. (517)

Here, 𝑨𝜼𝜼|a˘|2/4\smash{{\boldsymbol{A}}\doteq{\boldsymbol{\eta}}{\boldsymbol{\eta}}^{\dagger}|{\breve{a}}|^{2}/4}, 𝑩𝜼𝜼a˘2/4\smash{{\boldsymbol{B}}\doteq{\boldsymbol{\eta}}{\boldsymbol{\eta}}^{\intercal}{\breve{a}}^{2}/4}, ‘c.c.’ stands for complex conjugate, we used the linear approximation θ(𝘅±𝘀/2)θ(𝘅)±𝗸¯(𝘅)𝘀/2\smash{\theta({\boldsymbol{\mathsf{x}}}\pm{\boldsymbol{\mathsf{s}}}/2)\approx\theta({\boldsymbol{\mathsf{x}}})\pm\overline{{\boldsymbol{\mathsf{k}}}}({\boldsymbol{\mathsf{x}}})\cdot{\boldsymbol{\mathsf{s}}}/2}, with 𝗸¯(ω¯,𝒌¯)\smash{\overline{{\boldsymbol{\mathsf{k}}}}\equiv(-\overline{\omega},\overline{{\boldsymbol{k}}})}. Then,

𝗪𝝍(𝘅,𝗸)𝑨(𝘅)δ(𝗸𝗸¯(𝘅))+𝑨(𝘅)δ(𝗸+𝗸¯(𝘅))+2re(𝑩(𝘅)e2iθ(𝘅)δ(𝗸)).\displaystyle{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\approx{\boldsymbol{A}}({\boldsymbol{\mathsf{x}}})\,\delta({\boldsymbol{\mathsf{k}}}-\overline{{\boldsymbol{\mathsf{k}}}}({\boldsymbol{\mathsf{x}}}))+{\boldsymbol{A}}^{*}({\boldsymbol{\mathsf{x}}})\,\delta({\boldsymbol{\mathsf{k}}}+\overline{{\boldsymbol{\mathsf{k}}}}({\boldsymbol{\mathsf{x}}}))+2\operatorname{re}\big{(}{\boldsymbol{B}}({\boldsymbol{\mathsf{x}}})\mathrm{e}^{2\mathrm{i}\theta({\boldsymbol{\mathsf{x}}})}\delta({\boldsymbol{\mathsf{k}}})\big{)}. (518)

Let us adopt σ𝗑lc\smash{\sigma_{\mathsf{x}}\ll l_{c}}, where lc\smash{l_{c}} is the least characteristic scale of a˘\smash{{\breve{a}}}, 𝜼\smash{{\boldsymbol{\eta}}}, and 𝗸¯\smash{\overline{{\boldsymbol{\mathsf{k}}}}}. Then,

𝗪¯𝝍(𝘅,𝗸)𝑨(𝘅)𝒢𝗄(𝗸𝗸¯)+𝑨(𝘅)𝒢𝗄(𝗸+𝗸¯)+2re(𝑩(𝘅)𝒢𝗄(𝗸)ζe2iθ(𝘅)).\displaystyle\overline{{\boldsymbol{\mathsf{W}}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\approx{\boldsymbol{A}}({\boldsymbol{\mathsf{x}}})\,\mathcal{G}_{\mathsf{k}}({\boldsymbol{\mathsf{k}}}-\overline{{\boldsymbol{\mathsf{k}}}})+{\boldsymbol{A}}^{*}({\boldsymbol{\mathsf{x}}})\,\mathcal{G}_{\mathsf{k}}({\boldsymbol{\mathsf{k}}}+\overline{{\boldsymbol{\mathsf{k}}}})+2\operatorname{re}\big{(}{\boldsymbol{B}}({\boldsymbol{\mathsf{x}}})\mathcal{G}_{\mathsf{k}}({\boldsymbol{\mathsf{k}}})\zeta\mathrm{e}^{2\mathrm{i}\theta({\boldsymbol{\mathsf{x}}})}\big{)}. (519)

Here, 𝒢𝗄(𝗸)\smash{\mathcal{G}_{\mathsf{k}}({\boldsymbol{\mathsf{k}}})} are normalized Gaussians that can be replaced with delta functions if σ𝗄\smash{\sigma_{\mathsf{k}}} is small compared to any scale of interest in the 𝗸\smash{{\boldsymbol{\mathsf{k}}}} space:

𝒢𝗄(𝗸)1(2\upiσ𝗄)𝗇exp(|𝗸|22σ𝗄2)δ(𝗸).\displaystyle\mathcal{G}_{\mathsf{k}}({\boldsymbol{\mathsf{k}}})\doteq\frac{1}{(\sqrt{2\upi}\sigma_{\mathsf{k}})^{\mathsf{n}}}\,\exp\left(-\frac{|{\boldsymbol{\mathsf{k}}}|^{2}}{2\sigma_{\mathsf{k}}^{2}}\right)\to\delta({\boldsymbol{\mathsf{k}}}). (520)

Also, the function

ζ1(2\upiσ𝗑)𝗇d𝘅exp(|𝘅𝘅|22σ𝗑2+2i𝗸¯(𝘅)(𝘅𝘅))=e2|𝗸¯(𝘅)|2σ𝗑2\displaystyle\zeta\approx\frac{1}{(\sqrt{2\upi}\sigma_{\mathsf{x}})^{\mathsf{n}}}\int\mathrm{d}{\boldsymbol{\mathsf{x}}}^{\prime}\,\exp\left(-\frac{|{\boldsymbol{\mathsf{x}}}^{\prime}-{\boldsymbol{\mathsf{x}}}|^{2}}{2\sigma_{\mathsf{x}}^{2}}+2\mathrm{i}\overline{{\boldsymbol{\mathsf{k}}}}({\boldsymbol{\mathsf{x}}})\cdot({\boldsymbol{\mathsf{x}}}^{\prime}-{\boldsymbol{\mathsf{x}}})\right)=\mathrm{e}^{-2|\overline{{\boldsymbol{\mathsf{k}}}}({\boldsymbol{\mathsf{x}}})|^{2}\sigma_{\mathsf{x}}^{2}} (521)

can be made exponentially small by adopting σ𝗑|𝗸¯|1\smash{\sigma_{\mathsf{x}}\gg|\overline{{\boldsymbol{\mathsf{k}}}}|^{-1}}.454545Even though σ𝗑\smash{\sigma_{\mathsf{x}}} has been assumed small compared to lc\smash{l_{c}}, the smallness of the geometrical-optics parameter ϵ(|𝗸¯|lc)11\smash{\epsilon\doteq(|\overline{{\boldsymbol{\mathsf{k}}}}|l_{c})^{-1}\ll 1} allows choosing σ𝗑\smash{\sigma_{\mathsf{x}}} in the interval |𝗸¯|1σ𝗑lc\smash{|\overline{{\boldsymbol{\mathsf{k}}}}|^{-1}\ll\sigma_{\mathsf{x}}\ll l_{c}}. In this limit, the average Wigner matrix of an eikonal field is independent of σ𝗑\smash{\sigma_{\mathsf{x}}} and σ𝗄\smash{\sigma_{\mathsf{k}}}:

𝗪¯𝝍(𝘅,𝗸)𝑨(𝘅)δ(𝗸𝗸¯(𝘅))+𝑨(𝘅)δ(𝗸+𝗸¯(𝘅)).\displaystyle\overline{{\boldsymbol{\mathsf{W}}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\approx{\boldsymbol{A}}({\boldsymbol{\mathsf{x}}})\,\delta({\boldsymbol{\mathsf{k}}}-\overline{{\boldsymbol{\mathsf{k}}}}({\boldsymbol{\mathsf{x}}}))+{\boldsymbol{A}}^{*}({\boldsymbol{\mathsf{x}}})\,\delta({\boldsymbol{\mathsf{k}}}+\overline{{\boldsymbol{\mathsf{k}}}}({\boldsymbol{\mathsf{x}}})). (522)

This 𝗪¯𝝍\smash{\overline{{\boldsymbol{\mathsf{W}}}}_{{\boldsymbol{\psi}}}} is also Hermitian and positive-semidefinite (in agreement with the general theory from section A.1), because so are 𝑨\smash{{\boldsymbol{A}}} and 𝑨\smash{{\boldsymbol{A}}^{*}}. The same properties pertain to the Wigner matrix of an ensemble of randomly phased eikonal fields, because it equals the sum of the Wigner matrices of the individual components (see also section 7.4).

Appendix B Auxiliary proofs

B.1 Proof of (53)

Like in the case of (45), one finds that

(𝗟^𝝍(𝘅))i(𝗥^𝝍(𝘅))j\displaystyle(\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}{\boldsymbol{\psi}}({\boldsymbol{\mathsf{x}}}))^{i}(\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}{\boldsymbol{\psi}}({\boldsymbol{\mathsf{x}}}))^{*j} =𝘅|𝖫^ii|ψiψj|(𝖱^j)j|𝘅\displaystyle=\braket{{\boldsymbol{\mathsf{x}}}}{\widehat{\mathsf{L}}^{i}{}_{i^{\prime}}}{\psi^{i^{\prime}}}\braket{\psi^{j^{\prime}}}{(\widehat{\mathsf{R}}^{j}{}_{j^{\prime}})^{\dagger}}{{\boldsymbol{\mathsf{x}}}}
=(2\upi)𝗇𝘅|𝖫^i𝖶^𝝍iji(𝖱^)jj|𝘅\displaystyle=(2\upi)^{\mathsf{n}}\braket{{\boldsymbol{\mathsf{x}}}}{\widehat{\mathsf{L}}^{i}{}_{i^{\prime}}\widehat{\mathsf{W}}_{{\boldsymbol{\psi}}}^{i^{\prime}j^{\prime}}(\widehat{\mathsf{R}}^{\dagger})_{j^{\prime}}{}^{j}}{{\boldsymbol{\mathsf{x}}}}
=(2\upi)𝗇𝘅|(𝗟^𝗪^ψ𝗥^)ij|𝘅\displaystyle=(2\upi)^{\mathsf{n}}\braket{{\boldsymbol{\mathsf{x}}}}{(\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}\widehat{\boldsymbol{{\boldsymbol{\mathsf{W}}}}}_{\psi}\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}}^{\dagger})^{ij}}{{\boldsymbol{\mathsf{x}}}}
=d𝗸(𝗟(𝘅,𝗸)𝗪𝝍(𝘅,𝗸)𝗥(𝘅,𝗸))ij.\displaystyle=\textstyle\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\big{(}{\boldsymbol{\mathsf{L}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\star{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\star{\boldsymbol{\mathsf{R}}}^{\dagger}({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})\big{)}^{ij}. (523)

This proves (53a). At ϵ0\smash{\epsilon\to 0}, when \smash{\star} becomes the usual product, (523) gives

(𝗟^𝝍)(𝗥^𝝍)=d𝗸𝗟𝗪𝝍𝗥,\displaystyle(\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}{\boldsymbol{\psi}})(\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}{\boldsymbol{\psi}})^{\dagger}=\textstyle\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,{\boldsymbol{\mathsf{L}}}{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}{\boldsymbol{\mathsf{R}}}^{\dagger}, (524)

and in particular, taking the trace of (524) yields

(𝗥^𝝍)(𝗟^𝝍)=tr((𝗟^𝝍)(𝗥^𝝍))=d𝗸tr(𝗟𝗪𝝍𝗥)=d𝗸tr(𝗪𝝍𝗤).\displaystyle\textstyle(\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}{\boldsymbol{\psi}})^{\dagger}(\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}{\boldsymbol{\psi}})=\operatorname{tr}\big{(}(\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}{\boldsymbol{\psi}})(\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}{\boldsymbol{\psi}})^{\dagger}\big{)}=\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\operatorname{tr}({\boldsymbol{\mathsf{L}}}{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}{\boldsymbol{\mathsf{R}}}^{\dagger})=\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\operatorname{tr}({\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}{\boldsymbol{\mathsf{Q}}}). (525)

Here, 𝗤𝗥𝗟\smash{{\boldsymbol{\mathsf{Q}}}\doteq{\boldsymbol{\mathsf{R}}}^{\dagger}{\boldsymbol{\mathsf{L}}}}, and we used that tr(𝗔𝗕)=tr(𝗕𝗔)\smash{\operatorname{tr}({\boldsymbol{\mathsf{A}}}{\boldsymbol{\mathsf{B}}})=\operatorname{tr}({\boldsymbol{\mathsf{B}}}{\boldsymbol{\mathsf{A}}})} for any matrices 𝗔\smash{{\boldsymbol{\mathsf{A}}}} and 𝗕\smash{{\boldsymbol{\mathsf{B}}}}.

For real fields, one can also replace the integrand with

tr(𝗪𝝍𝗤)=tr(𝗤𝗪𝝍)=tr(𝗤𝗪𝝍)=tr(𝗪𝝍𝗤),\displaystyle\operatorname{tr}\big{(}{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}^{*}{\boldsymbol{\mathsf{Q}}}^{*}\big{)}=\operatorname{tr}\big{(}{\boldsymbol{\mathsf{Q}}}^{\dagger}{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}^{\dagger}\big{)}=\operatorname{tr}\big{(}{\boldsymbol{\mathsf{Q}}}^{\dagger}{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}\big{)}=\operatorname{tr}\big{(}{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}{\boldsymbol{\mathsf{Q}}}^{\dagger}\big{)}, (526)

where we used tr𝗔=tr𝗔\smash{\operatorname{tr}{\boldsymbol{\mathsf{A}}}^{\intercal}=\operatorname{tr}{\boldsymbol{\mathsf{A}}}}, (𝗔𝗕)=𝗕𝗔\smash{({\boldsymbol{\mathsf{A}}}{\boldsymbol{\mathsf{B}}})^{\intercal}={\boldsymbol{\mathsf{B}}}^{\intercal}{\boldsymbol{\mathsf{A}}}^{\intercal}}, 𝗪𝝍=𝗪𝝍\smash{{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}^{\dagger}={\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}}, and, again, tr(𝗔𝗕)=tr(𝗕𝗔)\smash{\operatorname{tr}({\boldsymbol{\mathsf{A}}}{\boldsymbol{\mathsf{B}}})=\operatorname{tr}({\boldsymbol{\mathsf{B}}}{\boldsymbol{\mathsf{A}}})}, respectively. In summary then,

(𝗥^𝝍)(𝗟^𝝍)=d𝗸tr(𝗪𝝍𝗤)=d𝗸tr(𝗪𝝍𝗤),\displaystyle\textstyle(\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}{\boldsymbol{\psi}})^{\dagger}(\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}{\boldsymbol{\psi}})=\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\operatorname{tr}({\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}{\boldsymbol{\mathsf{Q}}})=\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\operatorname{tr}\big{(}{\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}{\boldsymbol{\mathsf{Q}}}^{\dagger}\big{)}, (527)

so the anti-Hermitian part of 𝗤\smash{{\boldsymbol{\mathsf{Q}}}} does not contribute to the integrals. Thus,

(𝗥^𝝍)(𝗟^𝝍)=d𝗸tr(𝗪𝝍(𝗥𝗟)H).\displaystyle\textstyle(\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}{\boldsymbol{\psi}})^{\dagger}(\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}{\boldsymbol{\psi}})=\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\operatorname{tr}({\boldsymbol{\mathsf{W}}}_{{\boldsymbol{\psi}}}({\boldsymbol{\mathsf{R}}}^{\dagger}{\boldsymbol{\mathsf{L}}})_{\text{H}}). (528)

Because 𝗟^\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{L}}}}}} and 𝗥^\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{R}}}}}} are arbitrary, they can as well be swapped; then one obtains (53c).

B.2 Proof of (150)

Suppose the dominant term in 𝝁{\boldsymbol{\mu}} in (138) has the form 𝝁hτh{\boldsymbol{\mu}}_{h}\tau^{h}, where hh is a natural number and μh=𝒪(ϵ2)\smash{\mu_{h}=\mathcal{O}(\epsilon^{2})}. (Here hh is a power index, so no summation over hh is assumed.) Let us Taylor-expand 𝒥[A,G]\mathcal{J}[A,G] in 𝝁h\smash{{\boldsymbol{\mu}}_{h}}:

𝒥[A,G]\displaystyle\mathcal{J}[A,G] 𝒥[A,G0]𝒪(ϵ2)\displaystyle-\mathcal{J}[A,G_{0}]-\mathcal{O}(\epsilon^{2})
𝝁h𝝁h(d𝑲A(𝑿,𝑲)limν0+0dτeντ+iΩτ+i𝑲𝝁hτh)𝝁h=𝟎\displaystyle\approx{\boldsymbol{\mu}}_{h}\cdot\frac{\partial}{\partial{\boldsymbol{\mu}}_{h}}\left(\int\mathrm{d}{\boldsymbol{K}}\,A({\boldsymbol{X}},{\boldsymbol{K}})\lim_{\nu\to 0+}\int_{0}^{\infty}\mathrm{d}\tau\,\mathrm{e}^{-\nu\tau+\mathrm{i}\Omega\tau+\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{\mu}}_{h}\tau^{h}}\right)_{{\boldsymbol{\mu}}_{h}={\boldsymbol{0}}}
𝝁hd𝑲A(𝑿,𝑲)limν0+𝝁h(0dτeντ+iΩτ+i𝑲𝝁hτh)𝝁h=𝟎\displaystyle\approx{\boldsymbol{\mu}}_{h}\cdot\int\mathrm{d}{\boldsymbol{K}}\,A({\boldsymbol{X}},{\boldsymbol{K}})\lim_{\nu\to 0+}\frac{\partial}{\partial{\boldsymbol{\mu}}_{h}}\left(\int_{0}^{\infty}\mathrm{d}\tau\,\mathrm{e}^{-\nu\tau+\mathrm{i}\Omega\tau+\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{\mu}}_{h}\tau^{h}}\right)_{{\boldsymbol{\mu}}_{h}={\boldsymbol{0}}}
i𝝁hd𝑲𝑲A(𝑿,𝑲)limν0+0dττheντ+iΩτ\displaystyle\approx\mathrm{i}{\boldsymbol{\mu}}_{h}\cdot\int\mathrm{d}{\boldsymbol{K}}\,{\boldsymbol{K}}A({\boldsymbol{X}},{\boldsymbol{K}})\lim_{\nu\to 0+}\int_{0}^{\infty}\mathrm{d}\tau\,\tau^{h}\mathrm{e}^{-\nu\tau+\mathrm{i}\Omega\tau}
i1h𝝁hd𝑲𝑲A(𝑿,𝑲)dhG0(Ω(𝑿,𝑲))dΩh\displaystyle\approx\mathrm{i}^{1-h}{\boldsymbol{\mu}}_{h}\cdot\int\mathrm{d}{\boldsymbol{K}}\,{\boldsymbol{K}}A({\boldsymbol{X}},{\boldsymbol{K}})\,\frac{\mathrm{d}^{h}G_{0}(\Omega({\boldsymbol{X}},{\boldsymbol{K}}))}{\mathrm{d}\Omega^{h}}
i1h𝝁hðhΩhd𝑲𝑲A(𝑿,𝑲)G0(Ω(𝑿,𝑲)).\displaystyle\approx\mathrm{i}^{1-h}{\boldsymbol{\mu}}_{h}\cdot\frac{\eth^{h}}{\partial\Omega^{h}}\int\mathrm{d}{\boldsymbol{K}}\,{\boldsymbol{K}}A({\boldsymbol{X}},{\boldsymbol{K}})G_{0}(\Omega({\boldsymbol{X}},{\boldsymbol{K}})). (529)

Provided that AA is sufficiently smooth and well behaved, the overall coefficient here is 𝒪(1)\mathcal{O}(1), so 𝒥[A,G]𝒥[A,G0]=𝒪(μh)+𝒪(ϵ2)\smash{\mathcal{J}[A,G]-\mathcal{J}[A,G_{0}]=\mathcal{O}(\mu_{h})+\mathcal{O}(\epsilon^{2})}. Because μh=𝒪(ϵ2)\mu_{h}=\mathcal{O}(\epsilon^{2}), this proves (150).

B.3 Proof of (164)

Here, we show that

symb𝗑(\displaystyle\text{symb}_{\mathsf{x}}( u^αG^u^β)\displaystyle\widehat{u}^{\alpha}\widehat{G}\widehat{u}^{\beta})
=1(2\upi)Nd𝑺ei𝑲𝑺𝑿+𝑺/2|u^αG^u^β|𝑿𝑺/2\displaystyle=\frac{1}{(2\upi)^{N}}\int\mathrm{d}{\boldsymbol{S}}\,\mathrm{e}^{-\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}}\braket{{\boldsymbol{X}}+{\boldsymbol{S}}/2}{\widehat{u}^{\alpha}\widehat{G}\widehat{u}^{\beta}}{{\boldsymbol{X}}-{\boldsymbol{S}}/2}
=1(2\upi)Nd𝑿d𝑲′′d𝑲d𝑺d𝑺W𝒖αβ(𝑿,𝑲)G(𝑿,𝑲′′)ei𝑲𝑺+i(𝑲+𝑲′′)𝑺\displaystyle=\frac{1}{(2\upi)^{N}}\int\mathrm{d}{\boldsymbol{X}}^{\prime}\,\mathrm{d}{\boldsymbol{K}}^{\prime\prime}\,\mathrm{d}{\boldsymbol{K}}^{\prime}\,\mathrm{d}{\boldsymbol{S}}\,\mathrm{d}{\boldsymbol{S}}^{\prime}\,{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}}^{\prime},{\boldsymbol{K}}^{\prime})G({\boldsymbol{X}}^{\prime},{\boldsymbol{K}}^{\prime\prime})\,\mathrm{e}^{-\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}+\mathrm{i}({\boldsymbol{K}}^{\prime}+{\boldsymbol{K}}^{\prime\prime})\cdot{\boldsymbol{S}}^{\prime}}
×δ(𝑿+𝑺/2𝑿𝑺/2)δ(𝑿𝑺/2𝑿+𝑺/2)\displaystyle\hskip 56.9055pt\times\delta({\boldsymbol{X}}+{\boldsymbol{S}}/2-{\boldsymbol{X}}^{\prime}-{\boldsymbol{S}}^{\prime}/2)\delta({\boldsymbol{X}}-{\boldsymbol{S}}/2-{\boldsymbol{X}}^{\prime}+{\boldsymbol{S}}^{\prime}/2)
=1(2\upi)Nd𝑿d𝑲d𝑲′′d𝑺d𝑺W𝒖αβ(𝑿,𝑲)G(𝑿,𝑲′′)ei𝑲𝑺+i(𝑲+𝑲′′)𝑺\displaystyle=\frac{1}{(2\upi)^{N}}\int\mathrm{d}{\boldsymbol{X}}^{\prime}\,\mathrm{d}{\boldsymbol{K}}^{\prime}\,\mathrm{d}{\boldsymbol{K}}^{\prime\prime}\,\mathrm{d}{\boldsymbol{S}}\,\mathrm{d}{\boldsymbol{S}}^{\prime}\,{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}}^{\prime},{\boldsymbol{K}}^{\prime})G({\boldsymbol{X}}^{\prime},{\boldsymbol{K}}^{\prime\prime})\,\mathrm{e}^{-\mathrm{i}{\boldsymbol{K}}\cdot{\boldsymbol{S}}+\mathrm{i}({\boldsymbol{K}}^{\prime}+{\boldsymbol{K}}^{\prime\prime})\cdot{\boldsymbol{S}}^{\prime}}
×δ(𝑺𝑺)δ(𝑿𝑺/2𝑿+𝑺/2)\displaystyle\hskip 56.9055pt\times\delta({\boldsymbol{S}}-{\boldsymbol{S}}^{\prime})\delta({\boldsymbol{X}}-{\boldsymbol{S}}/2-{\boldsymbol{X}}^{\prime}+{\boldsymbol{S}}^{\prime}/2)
=1(2\upi)Nd𝑿d𝑲d𝑲′′d𝑺W𝒖αβ(𝑿,𝑲)G(𝑿,𝑲′′)ei(𝑲+𝑲′′𝑲)𝑺δ(𝑿𝑿)\displaystyle=\frac{1}{(2\upi)^{N}}\int\mathrm{d}{\boldsymbol{X}}^{\prime}\,\mathrm{d}{\boldsymbol{K}}^{\prime}\,\mathrm{d}{\boldsymbol{K}}^{\prime\prime}\,\mathrm{d}{\boldsymbol{S}}\,{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}}^{\prime},{\boldsymbol{K}}^{\prime})G({\boldsymbol{X}}^{\prime},{\boldsymbol{K}}^{\prime\prime})\,\mathrm{e}^{\mathrm{i}({\boldsymbol{K}}^{\prime}+{\boldsymbol{K}}^{\prime\prime}-{\boldsymbol{K}})\cdot{\boldsymbol{S}}}\delta({\boldsymbol{X}}-{\boldsymbol{X}}^{\prime})
=1(2\upi)Nd𝑲d𝑲′′d𝑺W𝒖αβ(𝑿,𝑲)G(𝑿,𝑲′′)ei(𝑲+𝑲′′𝑲)𝑺\displaystyle=\frac{1}{(2\upi)^{N}}\int\mathrm{d}{\boldsymbol{K}}^{\prime}\,\mathrm{d}{\boldsymbol{K}}^{\prime\prime}\,\mathrm{d}{\boldsymbol{S}}\,{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}}^{\prime})G({\boldsymbol{X}},{\boldsymbol{K}}^{\prime\prime})\,\mathrm{e}^{\mathrm{i}({\boldsymbol{K}}^{\prime}+{\boldsymbol{K}}^{\prime\prime}-{\boldsymbol{K}})\cdot{\boldsymbol{S}}}
=d𝑲W𝒖αβ(𝑿,𝑲)G(𝑿,𝑲𝑲).\displaystyle=\int\mathrm{d}{\boldsymbol{K}}^{\prime}\,{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}}^{\prime})G({\boldsymbol{X}},{\boldsymbol{K}}-{\boldsymbol{K}}^{\prime}). (530)

B.4 Proof of (170)

Using (141), (142), and (84) in application to W𝒖αβW_{{\boldsymbol{u}}}^{\alpha\beta}, one finds that

D0αβ(𝑿)\displaystyle D_{0}^{\alpha\beta}({\boldsymbol{X}}) d𝑲W¯𝒖αβ(𝑿,𝑲)G(𝑿,𝑲)\displaystyle\doteq\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}})G^{*}({\boldsymbol{X}},{\boldsymbol{K}})
=d𝑲W¯𝒖αβ(𝑿,𝑲)G(𝑿,𝑲)\displaystyle=\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},-{\boldsymbol{K}})G^{*}({\boldsymbol{X}},-{\boldsymbol{K}})
=d𝑲W¯𝒖αβ(𝑿,𝑲)G(𝑿,𝑲)\displaystyle=\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},-{\boldsymbol{K}})G({\boldsymbol{X}},{\boldsymbol{K}})
=d𝑲W¯𝒖αβ(𝑿,𝑲)G(𝑿,𝑲)\displaystyle=\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta*}({\boldsymbol{X}},{\boldsymbol{K}})G({\boldsymbol{X}},{\boldsymbol{K}})
=(D0αβ(𝑿))\displaystyle=(D_{0}^{\alpha\beta}({\boldsymbol{X}}))^{*} (531)

and also

Θαβc(𝑿)\displaystyle\Uptheta^{\alpha\beta c}({\boldsymbol{X}}) d𝑲W¯𝒖αβ(𝑿,𝑲)(G|c(𝑿,𝑲))\displaystyle\doteq-\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}})(G^{|c}({\boldsymbol{X}},{\boldsymbol{K}}))^{*}
=d𝑲W¯𝒖αβ(𝑿,𝑲)G|c(𝑿,𝑲)\displaystyle=\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}})G^{|c}({\boldsymbol{X}},-{\boldsymbol{K}})
=d𝑲W¯𝒖αβ(𝑿,𝑲)G|c(𝑿,𝑲)\displaystyle=\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},-{\boldsymbol{K}})G^{|c}({\boldsymbol{X}},{\boldsymbol{K}})
=d𝑲(W¯𝒖αβ(𝑿,𝑲))G|c(𝑿,𝑲)\displaystyle=\int\mathrm{d}{\boldsymbol{K}}\,(\overline{W}_{{\boldsymbol{u}}}^{\alpha\beta}({\boldsymbol{X}},{\boldsymbol{K}}))^{*}G^{|c}({\boldsymbol{X}},{\boldsymbol{K}})
=(Θαβc(𝑿)).\displaystyle=-(\Uptheta^{\alpha\beta c}({\boldsymbol{X}}))^{*}. (532)

B.5 Proof of (180)

Let us estimate

(1)f¯zα(JαμJβν𝒫μν(1)f¯zβ),\displaystyle\mathcal{L}^{(1)}\overline{f}\doteq\frac{\partial}{\partial z^{\alpha}}\left(J^{\alpha\mu}J^{\beta\nu}\mathcal{P}_{\mu\nu}^{(1)}\,\frac{\partial\overline{f}}{\partial z^{\beta}}\right), (533)

where 𝒫μν(1)\mathcal{P}_{\mu\nu}^{(1)} has the form

𝒫μν(1)d𝑲qνW¯(𝑿,𝑲)zμG(Ω(𝑿,𝑲)).\displaystyle\mathcal{P}_{\mu\nu}^{(1)}\doteq\int\mathrm{d}{\boldsymbol{K}}\,q_{\nu}\,\frac{\partial\overline{W}({\boldsymbol{X}},{\boldsymbol{K}})}{\partial z^{\mu}}\,G(\Omega({\boldsymbol{X}},{\boldsymbol{K}})). (534)

First, notice that

𝒫μν(1)\displaystyle\mathcal{P}_{\mu\nu}^{(1)} =zμd𝑲qνW¯Gd𝑲qνW¯Gzμ\displaystyle=\frac{\partial}{\partial z^{\mu}}\int\mathrm{d}{\boldsymbol{K}}\,q_{\nu}\overline{W}G-\int\mathrm{d}{\boldsymbol{K}}\,q_{\nu}\overline{W}\,\frac{\partial G}{\partial z^{\mu}}
=zμd𝑲qνW¯G+vλzμd𝑲qνqλW¯G\displaystyle=\frac{\partial}{\partial z^{\mu}}\int\mathrm{d}{\boldsymbol{K}}\,q_{\nu}\overline{W}G+\frac{\partial v^{\lambda}}{\partial z^{\mu}}\int\mathrm{d}{\boldsymbol{K}}\,q_{\nu}q_{\lambda}\overline{W}G^{\prime}
=zμd𝑲qνW¯G+vλzμðΩd𝑲qνqλW¯G\displaystyle=\frac{\partial}{\partial z^{\mu}}\int\mathrm{d}{\boldsymbol{K}}\,q_{\nu}\overline{W}G+\frac{\partial v^{\lambda}}{\partial z^{\mu}}\frac{\eth}{\partial\Omega}\int\mathrm{d}{\boldsymbol{K}}\,q_{\nu}q_{\lambda}\overline{W}G
𝒬ν(1)zμ+vλzμλν(1).\displaystyle\equiv\frac{\partial\mathcal{Q}_{\nu}^{(1)}}{\partial z^{\mu}}+\frac{\partial v^{\lambda}}{\partial z^{\mu}}\,\mathcal{R}_{\lambda\nu}^{(1)}. (535)

Because 𝒬ν(1)\mathcal{Q}_{\nu}^{(1)} and λν(1)\mathcal{R}_{\lambda\nu}^{(1)} are 𝒪(ε2)\mathcal{O}(\varepsilon^{2}), one has 𝒫μν(1)κμε2\smash{\mathcal{P}_{\mu\nu}^{(1)}\sim\kappa_{\mu}\varepsilon^{2}}, where κμ\kappa_{\mu} is the characteristic inverse scale along the μ\muth phase-space axis. Thus,

(1)f¯(Jαμκακμ)Jβνκβε2f¯=𝒪(ϵε2),\displaystyle\mathcal{L}^{(1)}\overline{f}\sim(J^{\alpha\mu}\kappa_{\alpha}\kappa_{\mu})J^{\beta\nu}\kappa_{\beta}\varepsilon^{2}\overline{f}=\mathcal{O}(\epsilon\varepsilon^{2}), (536)

where we used (see (69) and (86))

Jαμκακμκxκp=𝒪(ϵ).\displaystyle J^{\alpha\mu}\kappa_{\alpha}\kappa_{\mu}\sim\kappa_{x}\kappa_{p}=\mathcal{O}(\epsilon). (537)

The first part of (180) is obtained by considering im(1)f¯\operatorname{im}\mathcal{L}^{(1)}\overline{f} and using (536).

Let us also estimate

(2)f¯zα(JαμJβν𝒫μν(2)f¯zβ),\displaystyle\mathcal{L}^{(2)}\overline{f}\doteq\frac{\partial}{\partial z^{\alpha}}\left(J^{\alpha\mu}J^{\beta\nu}\mathcal{P}_{\mu\nu}^{(2)}\,\frac{\partial\overline{f}}{\partial z^{\beta}}\right), (538)

where 𝒫μν(2)\mathcal{P}_{\mu\nu}^{(2)} has the form

𝒫μν(2)d𝑲2W¯zμzνG(Ω(𝑿,𝑲)).\displaystyle\mathcal{P}_{\mu\nu}^{(2)}\doteq\int\mathrm{d}{\boldsymbol{K}}\,\frac{\partial^{2}\overline{W}}{\partial z^{\mu}\partial z^{\nu}}\,G(\Omega({\boldsymbol{X}},{\boldsymbol{K}})). (539)

First, note that

𝒫μν(2)=\displaystyle\mathcal{P}_{\mu\nu}^{(2)}= d𝑲2W¯zμzνG\displaystyle\int\mathrm{d}{\boldsymbol{K}}\,\frac{\partial^{2}\overline{W}}{\partial z^{\mu}\partial z^{\nu}}\,G
=\displaystyle= zμd𝑲W¯zνGd𝑲W¯zνGzμ\displaystyle\frac{\partial}{\partial z^{\mu}}\int\mathrm{d}{\boldsymbol{K}}\,\frac{\partial\overline{W}}{\partial z^{\nu}}\,G-\int\mathrm{d}{\boldsymbol{K}}\,\frac{\partial\overline{W}}{\partial z^{\nu}}\frac{\partial G}{\partial z^{\mu}}
=\displaystyle= 2zμzνd𝑲W¯Gzμd𝑲W¯Gzνzνd𝑲W¯Gzμ+d𝑲W¯2Gzμzν\displaystyle\frac{\partial^{2}}{\partial z^{\mu}\partial z^{\nu}}\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}G-\frac{\partial}{\partial z^{\mu}}\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}\,\frac{\partial G}{\partial z^{\nu}}-\frac{\partial}{\partial z^{\nu}}\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}\,\frac{\partial G}{\partial z^{\mu}}+\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}\,\frac{\partial^{2}G}{\partial z^{\mu}\partial z^{\nu}}
=\displaystyle= 2zμzνd𝑲W¯Gzμd𝑲W¯G(qλvλzν)\displaystyle\frac{\partial^{2}}{\partial z^{\mu}\partial z^{\nu}}\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}\,G-\frac{\partial}{\partial z^{\mu}}\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}\,G^{\prime}\left(-q_{\lambda}\,\frac{\partial v^{\lambda}}{\partial z^{\nu}}\right)
zνd𝑲W¯G(qλvλzμ)+d𝑲W¯2Gzμzν\displaystyle-\frac{\partial}{\partial z^{\nu}}\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}\,G^{\prime}\left(-q_{\lambda}\,\frac{\partial v^{\lambda}}{\partial z^{\mu}}\right)+\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}\,\frac{\partial^{2}G}{\partial z^{\mu}\partial z^{\nu}}
=\displaystyle= 2zμzνd𝑲W¯G+zμ(vλzνðΩd𝑲qλW¯G)\displaystyle\frac{\partial^{2}}{\partial z^{\mu}\partial z^{\nu}}\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}G+\frac{\partial}{\partial z^{\mu}}\left(\frac{\partial v^{\lambda}}{\partial z^{\nu}}\frac{\eth}{\partial\Omega}\int\mathrm{d}{\boldsymbol{K}}\,q_{\lambda}\overline{W}G\right)
+zν(vλzμðΩd𝑲qλW¯G)+d𝑲W¯2Gzμzν.\displaystyle+\frac{\partial}{\partial z^{\nu}}\left(\frac{\partial v^{\lambda}}{\partial z^{\mu}}\frac{\eth}{\partial\Omega}\int\mathrm{d}{\boldsymbol{K}}\,q_{\lambda}\overline{W}G\right)+\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}\frac{\partial^{2}G}{\partial z^{\mu}\partial z^{\nu}}. (540)

Next, note that

d𝑲W¯2Gzμzν\displaystyle\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}\,\frac{\partial^{2}G}{\partial z^{\mu}\partial z^{\nu}} =d𝑲W¯zμ(qλGvλzν)\displaystyle=\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}\,\frac{\partial}{\partial z^{\mu}}\left(-q_{\lambda}G^{\prime}\frac{\partial v^{\lambda}}{\partial z^{\nu}}\right)
=vλzνd𝑲qλW¯Gzμ2vλzμzνd𝑲qλW¯G\displaystyle=-\frac{\partial v^{\lambda}}{\partial z^{\nu}}\int\mathrm{d}{\boldsymbol{K}}\,q_{\lambda}\overline{W}\,\frac{\partial G^{\prime}}{\partial z^{\mu}}-\frac{\partial^{2}v^{\lambda}}{\partial z^{\mu}\partial z^{\nu}}\int\mathrm{d}{\boldsymbol{K}}\,q_{\lambda}\overline{W}G^{\prime}
=vλzνvδzμd𝑲qλqδW¯G′′2vλzμzνd𝑲qλW¯G\displaystyle=\frac{\partial v^{\lambda}}{\partial z^{\nu}}\frac{\partial v^{\delta}}{\partial z^{\mu}}\int\mathrm{d}{\boldsymbol{K}}\,q_{\lambda}q_{\delta}\overline{W}G^{\prime\prime}-\frac{\partial^{2}v^{\lambda}}{\partial z^{\mu}\partial z^{\nu}}\int\mathrm{d}{\boldsymbol{K}}\,q_{\lambda}\overline{W}G^{\prime}
=vλzνvδzμð2Ω2d𝑲qλqδW¯G2vλzμzνðΩd𝑲qλW¯G.\displaystyle=\frac{\partial v^{\lambda}}{\partial z^{\nu}}\frac{\partial v^{\delta}}{\partial z^{\mu}}\frac{\eth^{2}}{\partial\Omega^{2}}\int\mathrm{d}{\boldsymbol{K}}\,q_{\lambda}q_{\delta}\overline{W}G-\frac{\partial^{2}v^{\lambda}}{\partial z^{\mu}\partial z^{\nu}}\frac{\eth}{\partial\Omega}\int\mathrm{d}{\boldsymbol{K}}\,q_{\lambda}\overline{W}G. (541)

Assuming the notation

𝒮(2)\displaystyle\mathcal{S}^{(2)} d𝑲W¯G=𝒪(ε2),\displaystyle\doteq\int\mathrm{d}{\boldsymbol{K}}\,\overline{W}G=\mathcal{O}\left(\varepsilon^{2}\right),
𝒬λ(2)\displaystyle\mathcal{Q}_{\lambda}^{(2)} ðΩd𝑲qλW¯G=𝒪(ε2),\displaystyle\doteq\frac{\eth}{\partial\Omega}\int\mathrm{d}{\boldsymbol{K}}\,q_{\lambda}\overline{W}G=\mathcal{O}\left(\varepsilon^{2}\right),
λδ(2)\displaystyle\mathcal{R}_{\lambda\delta}^{(2)} ð2Ω2d𝑲qλqδW¯G=𝒪(ε2),\displaystyle\doteq\frac{\eth^{2}}{\partial\Omega^{2}}\int\mathrm{d}{\boldsymbol{K}}\,q_{\lambda}q_{\delta}\overline{W}G=\mathcal{O}\left(\varepsilon^{2}\right), (542)

one can then rewrite 𝒫μν(2)\mathcal{P}_{\mu\nu}^{(2)} as follows:

𝒫μν(2)\displaystyle\mathcal{P}_{\mu\nu}^{(2)} =2𝒮(2)zμzν+zμ(vλzν𝒬λ(2))+zν(vλzμ𝒬λ(2))2vλzμzν𝒬λ(2)+vλzνvδzμλδ(2)\displaystyle=\frac{\partial^{2}\mathcal{S}^{(2)}}{\partial z^{\mu}\partial z^{\nu}}+\frac{\partial}{\partial z^{\mu}}\left(\frac{\partial v^{\lambda}}{\partial z^{\nu}}\,\mathcal{Q}_{\lambda}^{(2)}\right)+\frac{\partial}{\partial z^{\nu}}\left(\frac{\partial v^{\lambda}}{\partial z^{\mu}}\,\mathcal{Q}_{\lambda}^{(2)}\right)-\frac{\partial^{2}v^{\lambda}}{\partial z^{\mu}\partial z^{\nu}}\,\mathcal{Q}_{\lambda}^{(2)}+\frac{\partial v^{\lambda}}{\partial z^{\nu}}\frac{\partial v^{\delta}}{\partial z^{\mu}}\,\mathcal{R}_{\lambda\delta}^{(2)}
=2𝒮(2)zμzν+2vλzμzν𝒬λ(2)+vλzν𝒬λ(2)zμ+vλzμ𝒬λ(2)zν+2vλzμzν𝒬λ(2)+vλzνvδzμλδ(2).\displaystyle=\frac{\partial^{2}\mathcal{S}^{(2)}}{\partial z^{\mu}\partial z^{\nu}}+\frac{\partial^{2}v^{\lambda}}{\partial z^{\mu}\partial z^{\nu}}\,\mathcal{Q}_{\lambda}^{(2)}+\frac{\partial v^{\lambda}}{\partial z^{\nu}}\frac{\partial\mathcal{Q}_{\lambda}^{(2)}}{\partial z^{\mu}}+\frac{\partial v^{\lambda}}{\partial z^{\mu}}\frac{\partial\mathcal{Q}_{\lambda}^{(2)}}{\partial z^{\nu}}+\frac{\partial^{2}v^{\lambda}}{\partial z^{\mu}\partial z^{\nu}}\,\mathcal{Q}_{\lambda}^{(2)}+\frac{\partial v^{\lambda}}{\partial z^{\nu}}\frac{\partial v^{\delta}}{\partial z^{\mu}}\,\mathcal{R}_{\lambda\delta}^{(2)}.

Each term on the right-hand side of this equation scales as ε2κμκν\varepsilon^{2}\kappa_{\mu}\kappa_{\nu}, so

(2)f¯(Jαμκακμ)(Jβνκβκν)ε2f¯ϵ2ε2f¯,\displaystyle\mathcal{L}^{(2)}\overline{f}\sim(J^{\alpha\mu}\kappa_{\alpha}\kappa_{\mu})(J^{\beta\nu}\kappa_{\beta}\kappa_{\nu})\varepsilon^{2}\overline{f}\sim\epsilon^{2}\varepsilon^{2}\overline{f}, (543)

where we again used (537). The second part of (180) is obtained by considering re(2)f¯\operatorname{re}\mathcal{L}^{(2)}\overline{f} and using (543).

B.6 Proof of (185)

Using (183) and assuming the notation dtt+vγγ\mathrm{d}_{t}\doteq\partial_{t}+v^{\gamma}\partial_{\gamma}, one finds that

α\displaystyle\partial_{\alpha} (D^αββf¯)α((𝖣αβ+ϱαβ)βf¯)\displaystyle(\widehat{D}^{\alpha\beta}\partial_{\beta}\overline{f})-\partial_{\alpha}((\mathsf{D}^{\alpha\beta}+\varrho^{\alpha\beta})\partial_{\beta}\overline{f})
=\displaystyle= α(Θαβdtβf¯+12(dtΘαβ)βf¯)\displaystyle-\partial_{\alpha}\left(\Uptheta^{\alpha\beta}\mathrm{d}_{t}\partial_{\beta}\overline{f}+\frac{1}{2}\,(\mathrm{d}_{t}\Uptheta^{\alpha\beta})\partial_{\beta}\overline{f}\right)
=\displaystyle= α(12Θαβdtβf¯+12dt(Θαββf¯))\displaystyle-\partial_{\alpha}\left(\frac{1}{2}\,\Uptheta^{\alpha\beta}\mathrm{d}_{t}\partial_{\beta}\overline{f}+\frac{1}{2}\,\mathrm{d}_{t}(\Uptheta^{\alpha\beta}\partial_{\beta}\overline{f})\right)
=\displaystyle= α(12Θαββdtf¯12Θαβ(βvγ)γf¯)α(12dt(Θαββf¯))\displaystyle-\partial_{\alpha}\left(\frac{1}{2}\,\Uptheta^{\alpha\beta}\partial_{\beta}\mathrm{d}_{t}\overline{f}-\frac{1}{2}\,\Uptheta^{\alpha\beta}(\partial_{\beta}v^{\gamma})\partial_{\gamma}\overline{f}\right)-\partial_{\alpha}\left(\frac{1}{2}\,\mathrm{d}_{t}(\Uptheta^{\alpha\beta}\partial_{\beta}\overline{f})\right)
=\displaystyle= α(12Θαββdtf¯)+α(12Θαβ(βvγ)γf¯)\displaystyle-\partial_{\alpha}\left(\frac{1}{2}\,\Uptheta^{\alpha\beta}\partial_{\beta}\mathrm{d}_{t}\overline{f}\right)+\partial_{\alpha}\left(\frac{1}{2}\,\Uptheta^{\alpha\beta}(\partial_{\beta}v^{\gamma})\partial_{\gamma}\overline{f}\right)
dt(12α(Θαββf¯))(αvγ)γ(12Θαββf¯).\displaystyle-\mathrm{d}_{t}\left(\frac{1}{2}\,\partial_{\alpha}(\Uptheta^{\alpha\beta}\partial_{\beta}\overline{f})\right)-(\partial_{\alpha}v^{\gamma})\partial_{\gamma}\left(\frac{1}{2}\,\Uptheta^{\alpha\beta}\partial_{\beta}\overline{f}\right). (544)

Because Θαβ=𝒪(ε2)\Uptheta^{\alpha\beta}=\mathcal{O}(\varepsilon^{2}) and dtf¯=𝒪(ε2)\mathrm{d}_{t}\overline{f}=\mathcal{O}(\varepsilon^{2}), the first term on the right-hand side of (544) is negligible. Also note that due to (99), the factor αvγ\smash{\partial_{\alpha}v^{\gamma}} in the last term on the right-hand side of (544) commutes with γ\smash{\partial_{\gamma}}. Hence, one obtains

α(D^αββf¯)\displaystyle\partial_{\alpha}(\widehat{D}^{\alpha\beta}\partial_{\beta}\overline{f}) α(𝖣αββf¯)+dt(12α(Θαββf¯))\displaystyle-\partial_{\alpha}(\mathsf{D}^{\alpha\beta}\partial_{\beta}\overline{f})+\mathrm{d}_{t}\left(\frac{1}{2}\,\partial_{\alpha}(\Uptheta^{\alpha\beta}\partial_{\beta}\overline{f})\right)
=α(ϱαββf¯+12Θαβ(βvγ)γf¯)γ(12Θαβ(αvγ)βf¯)\displaystyle=\partial_{\alpha}\left(\varrho^{\alpha\beta}\partial_{\beta}\overline{f}+\frac{1}{2}\,\Uptheta^{\alpha\beta}\left(\partial_{\beta}v^{\gamma}\right)\partial_{\gamma}\overline{f}\right)-\partial_{\gamma}\left(\frac{1}{2}\,\Uptheta^{\alpha\beta}(\partial_{\alpha}v^{\gamma})\partial_{\beta}\overline{f}\right)
=α(ϱαββf¯+12(Θαγ(γvβ)Θγβ(γvα))βf¯)\displaystyle=\partial_{\alpha}\left(\varrho^{\alpha\beta}\partial_{\beta}\overline{f}+\frac{1}{2}\left(\Uptheta^{\alpha\gamma}(\partial_{\gamma}v^{\beta})-\Uptheta^{\gamma\beta}(\partial_{\gamma}v^{\alpha})\right)\partial_{\beta}\overline{f}\right)
α(Uαββf¯).\displaystyle\equiv\partial_{\alpha}(U^{\alpha\beta}\partial_{\beta}\overline{f}). (545)

Next, notice that

ϱαβ\displaystyle\varrho^{\alpha\beta} =12JαμJβνd𝑲(qνW¯zμqμW¯zν)1Ω\displaystyle=-\frac{1}{2}\,J^{\alpha\mu}J^{\beta\nu}\fint\mathrm{d}{\boldsymbol{K}}\left(q_{\nu}\,\frac{\partial\overline{W}}{\partial z^{\mu}}-q_{\mu}\,\frac{\partial\overline{W}}{\partial z^{\nu}}\right)\frac{1}{\Omega}
=12JαμJβνd𝑲qνΩW¯zμ+12JαμJνβd𝑲qμΩW¯zν\displaystyle=-\frac{1}{2}\,J^{\alpha\mu}J^{\beta\nu}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\nu}}{\Omega}\frac{\partial\overline{W}}{\partial z^{\mu}}+\frac{1}{2}\,J^{\alpha\mu}J^{\nu\beta}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\mu}}{\Omega}\frac{\partial\overline{W}}{\partial z^{\nu}}
=12JαμJβν(zμd𝑲qνW¯Ωd𝑲qνW¯zμ1Ω\displaystyle=-\frac{1}{2}\,J^{\alpha\mu}J^{\beta\nu}\left(\frac{\partial}{\partial z^{\mu}}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\nu}\overline{W}}{\Omega}-\fint\mathrm{d}{\boldsymbol{K}}\,q_{\nu}\overline{W}\,\frac{\partial}{\partial z^{\mu}}\frac{1}{\Omega}\right.
zνd𝑲qμW¯Ω+d𝑲qμW¯zν1Ω)\displaystyle\hskip 71.13188pt\left.-\frac{\partial}{\partial z^{\nu}}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\mu}\overline{W}}{\Omega}+\fint\mathrm{d}{\boldsymbol{K}}\,q_{\mu}\overline{W}\frac{\partial}{\partial z^{\nu}}\frac{1}{\Omega}\right)
=12JαμJβν(zμd𝑲qνW¯Ω+vλzμðΩd𝑲qλqνW¯Ω\displaystyle=-\frac{1}{2}\,J^{\alpha\mu}J^{\beta\nu}\left(\frac{\partial}{\partial z^{\mu}}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\nu}\overline{W}}{\Omega}+\frac{\partial v^{\lambda}}{\partial z^{\mu}}\frac{\eth}{\partial\Omega}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\lambda}q_{\nu}\overline{W}}{\Omega}\right.
zνd𝑲qμW¯ΩvλzνðΩd𝑲qλqμW¯Ω).\displaystyle\hskip 71.13188pt\left.-\frac{\partial}{\partial z^{\nu}}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\mu}\overline{W}}{\Omega}-\frac{\partial v^{\lambda}}{\partial z^{\nu}}\frac{\eth}{\partial\Omega}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\lambda}q_{\mu}\overline{W}}{\Omega}\right). (546)

Assuming the notation

Qμ12d𝑲qμW¯Ω,Rμν12ðΩd𝑲qμqνW¯Ω,\displaystyle Q_{\mu}\doteq\frac{1}{2}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\mu}\overline{W}}{\Omega},\qquad R_{\mu\nu}\doteq\frac{1}{2}\frac{\eth}{\partial\Omega}\fint\mathrm{d}{\boldsymbol{K}}\,\frac{q_{\mu}q_{\nu}\overline{W}}{\Omega}, (547)

one can rewrite (546) compactly as follows:

ϱαβ=JαμJβν(νQμμQν)+JαμJβν((νvλ)Rλμ(μvλ)Rλν).\displaystyle\varrho^{\alpha\beta}=J^{\alpha\mu}J^{\beta\nu}(\partial_{\nu}Q_{\mu}-\partial_{\mu}Q_{\nu})+J^{\alpha\mu}J^{\beta\nu}((\partial_{\nu}v^{\lambda})R_{\lambda\mu}-(\partial_{\mu}v^{\lambda})R_{\lambda\nu}). (548)

Notice also that Θαβ=2JαμJβνRμν\Uptheta^{\alpha\beta}=2J^{\alpha\mu}J^{\beta\nu}R_{\mu\nu}. Hence, for Uαβ\smash{U^{\alpha\beta}} introduced in (545), one has

U\displaystyle U αβJαμJβν(νQμμQν){}^{\alpha\beta}-J^{\alpha\mu}J^{\beta\nu}(\partial_{\nu}Q_{\mu}-\partial_{\mu}Q_{\nu})
=ϱαβJαμJβν(νQμμQν)(Θγβ(γvα)Θαγ(γvβ))/2\displaystyle=\varrho^{\alpha\beta}-J^{\alpha\mu}J^{\beta\nu}(\partial_{\nu}Q_{\mu}-\partial_{\mu}Q_{\nu})-(\Uptheta^{\gamma\beta}(\partial_{\gamma}v^{\alpha})-\Uptheta^{\alpha\gamma}(\partial_{\gamma}v^{\beta}))/2
=JαμJβν((νvλ)Rλμ(μvλ)Rλν)+JαμJγν(γvβ)RμνJγμJβν(γvα)Rμν\displaystyle=J^{\alpha\mu}J^{\beta\nu}((\partial_{\nu}v^{\lambda})R_{\lambda\mu}-(\partial_{\mu}v^{\lambda})R_{\lambda\nu})+J^{\alpha\mu}J^{\gamma\nu}(\partial_{\gamma}v^{\beta})R_{\mu\nu}-J^{\gamma\mu}J^{\beta\nu}(\partial_{\gamma}v^{\alpha})R_{\mu\nu}
=JαμJβν(νvλ)RλμJαμJβν(μvλ)Rλν+JαμJγν(γvβ)RμνJγμJβν(γvα)Rμν\displaystyle=J^{\alpha\mu}J^{\beta\nu}(\partial_{\nu}v^{\lambda})R_{\lambda\mu}-J^{\alpha\mu}J^{\beta\nu}(\partial_{\mu}v^{\lambda})R_{\lambda\nu}+J^{\alpha\mu}J^{\gamma\nu}(\partial_{\gamma}v^{\beta})R_{\mu\nu}-J^{\gamma\mu}J^{\beta\nu}(\partial_{\gamma}v^{\alpha})R_{\mu\nu}
=JαμJβλ(λvν)RμνJαλJβν(λvμ)Rμν+JαμJγν(γvβ)RμνJγμJβν(γvα)Rμν\displaystyle=J^{\alpha\mu}J^{\beta\lambda}(\partial_{\lambda}v^{\nu})R_{\mu\nu}-J^{\alpha\lambda}J^{\beta\nu}(\partial_{\lambda}v^{\mu})R_{\mu\nu}+J^{\alpha\mu}J^{\gamma\nu}(\partial_{\gamma}v^{\beta})R_{\mu\nu}-J^{\gamma\mu}J^{\beta\nu}(\partial_{\gamma}v^{\alpha})R_{\mu\nu}
=(JαμJβλJνγJαλJβνJμγ+JαμJγνJβλJγμJβνJαλ)(γλ2H¯)Rμν\displaystyle=(J^{\alpha\mu}J^{\beta\lambda}J^{\nu\gamma}-J^{\alpha\lambda}J^{\beta\nu}J^{\mu\gamma}+J^{\alpha\mu}J^{\gamma\nu}J^{\beta\lambda}-J^{\gamma\mu}J^{\beta\nu}J^{\alpha\lambda})(\partial_{\gamma\lambda}^{2}\overline{H})R_{\mu\nu}
=(JαμJβλJνγJαλJβνJμγJαμJγνJλβ+JμγJβνJαλ)(γλ2H¯)Rμν\displaystyle=(J^{\alpha\mu}J^{\beta\lambda}J^{\nu\gamma}-J^{\alpha\lambda}J^{\beta\nu}J^{\mu\gamma}-J^{\alpha\mu}J^{\gamma\nu}J^{\lambda\beta}+J^{\mu\gamma}J^{\beta\nu}J^{\alpha\lambda})(\partial_{\gamma\lambda}^{2}\overline{H})R_{\mu\nu}
=0,\displaystyle=0, (549)

where we used (92) for vαv^{\alpha} and the anti-symmetry of JαβJ^{\alpha\beta}. Therefore,

Uαβ=JαμJβν(νQμμQν)=(JαμJβνJανJβμ)νQμ=Uβα,\displaystyle U^{\alpha\beta}=J^{\alpha\mu}J^{\beta\nu}(\partial_{\nu}Q_{\mu}-\partial_{\mu}Q_{\nu})=(J^{\alpha\mu}J^{\beta\nu}-J^{\alpha\nu}J^{\beta\mu})\partial_{\nu}Q_{\mu}=-U^{\beta\alpha}, (550)

and accordingly,

αUαβ\displaystyle\partial_{\alpha}U^{\alpha\beta} =(JαμJβνJανJβμ)να2Qμ=JαμJβννα2Qμ\displaystyle=(J^{\alpha\mu}J^{\beta\nu}-J^{\alpha\nu}J^{\beta\mu})\partial_{\nu\alpha}^{2}Q_{\mu}=J^{\alpha\mu}J^{\beta\nu}\partial_{\nu\alpha}^{2}Q_{\mu}
=JνμJβανα2Qμ=Jβαα(JνμνQμ)=Jαβα(JμνμQν)JαβαΦ.\displaystyle=J^{\nu\mu}J^{\beta\alpha}\partial_{\nu\alpha}^{2}Q_{\mu}=J^{\beta\alpha}\partial_{\alpha}(J^{\nu\mu}\partial_{\nu}Q_{\mu})=-J^{\alpha\beta}\partial_{\alpha}(J^{\mu\nu}\partial_{\mu}Q_{\nu})\equiv J^{\alpha\beta}\partial_{\alpha}\Phi. (551)

Here, ΦJμνμQν\smash{\Phi\doteq-J^{\mu\nu}\partial_{\mu}Q_{\nu}}, which is equivalent to (184). From (551) and the fact that Uαβαβ=0\smash{U^{\alpha\beta}\partial_{\alpha\beta}=0} due to the anti-symmetry of Uαβ\smash{U^{\alpha\beta}}, one has

α(Uαββf¯)=Jαβ(αΦ)(βf¯)={Φ,f¯}.\displaystyle\partial_{\alpha}(U^{\alpha\beta}\partial_{\beta}\overline{f})=J^{\alpha\beta}(\partial_{\alpha}\Phi)(\partial_{\beta}\overline{f})=\{\Phi,\overline{f}\}. (552)

Hence, (545) leads to (185).

B.7 Proof of (237)

The correlation function

ss(t,𝒙,τ,𝒔;𝒑,𝒑)gs(t+τ/2,𝒙+𝒔/2,𝒑)gs(tτ/2,𝒙𝒔/2,𝒑)¯\displaystyle\mathfrak{C}_{ss^{\prime}}(t,{\boldsymbol{x}},\tau,{\boldsymbol{s}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\doteq\overline{g_{s}(t+\tau/2,{\boldsymbol{x}}+{\boldsymbol{s}}/2,{\boldsymbol{p}})g_{s^{\prime}}(t-\tau/2,{\boldsymbol{x}}-{\boldsymbol{s}}/2,{\boldsymbol{p}}^{\prime})} (553)

can be readily expressed as

ss=(δssσs=σs+σsσs)\displaystyle\mathfrak{C}_{ss^{\prime}}=\bigg{(}\delta_{ss^{\prime}}\!\!\!\sum_{\sigma_{s}=\sigma^{\prime}_{s^{\prime}}}+\sum_{\sigma_{s}\neq\sigma^{\prime}_{s^{\prime}}}\bigg{)}\,\langle δ(𝒙+𝒔/2𝒙¯σs(t+τ/2))δ(𝒑𝒑¯σs(t+τ/2))\displaystyle\delta({\boldsymbol{x}}+{\boldsymbol{s}}/2-\overline{{\boldsymbol{x}}}_{\sigma_{s}}(t+\tau/2))\delta({\boldsymbol{p}}-\overline{{\boldsymbol{p}}}_{\sigma_{s}}(t+\tau/2))
×δ(𝒙𝒔/2𝒙¯σs(tτ/2))δ(𝒑𝒑¯σs(tτ/2))𝔣¯.\displaystyle\times\delta({\boldsymbol{x}}-{\boldsymbol{s}}/2-\overline{{\boldsymbol{x}}}_{\sigma^{\prime}_{s^{\prime}}}(t-\tau/2))\delta({\boldsymbol{p}}^{\prime}-\overline{{\boldsymbol{p}}}_{\sigma^{\prime}_{s^{\prime}}}(t-\tau/2))\rangle-\mathfrak{C}_{\overline{\mathfrak{f}}}.

Here, \smash{\braket{\ldots}} is another (in addition to overbar) notation for averaging used in this appendix, the dependence of 𝔣¯s\smash{\overline{\mathfrak{f}}_{s}} on (t,𝒙)\smash{(t,{\boldsymbol{x}})} is neglected, and ‘σsσs\smash{\sigma_{s}\neq\sigma^{\prime}_{s^{\prime}}}’ denotes that excluded are the terms that have s=s\smash{s^{\prime}=s} and σs=σs\smash{\sigma_{s}=\sigma^{\prime}_{s^{\prime}}} simultaneously. Aside from this, the summations over σs\smash{\sigma_{s}} are taken over all Ns1\smash{N_{s}\gg 1} particles of type s\smash{s}, and the summations over σs\smash{\sigma_{s^{\prime}}} are taken over all Ns1\smash{N_{s^{\prime}}\gg 1} particles of type s\smash{s^{\prime}}. Also,

𝔣¯𝔣¯s(t+τ/2,𝒙+𝒔/2,𝒑)𝔣¯s(tτ/2,𝒙𝒔/2,𝒑).\displaystyle\mathfrak{C}_{\overline{\mathfrak{f}}}\doteq\braket{\overline{\mathfrak{f}}_{s}(t+\tau/2,{\boldsymbol{x}}+{\boldsymbol{s}}/2,{\boldsymbol{p}})\overline{\mathfrak{f}}_{s^{\prime}}(t-\tau/2,{\boldsymbol{x}}-{\boldsymbol{s}}/2,{\boldsymbol{p}}^{\prime})}. (554)

To the leading order, pair correlations can be neglected. Then,

σsσs\displaystyle\sum_{\sigma_{s}\neq\sigma^{\prime}_{s^{\prime}}}\braket{\ldots} =σsσsδ(𝒙+𝒔/2𝒙¯σs(t+τ/2))δ(𝒑𝒑¯σs(t+τ/2))𝔣¯s(t+τ/2,𝒙+𝒔/2,𝒑)/Ns\displaystyle=\sum_{\sigma_{s}\neq\sigma^{\prime}_{s^{\prime}}}\underbrace{\braket{\delta({\boldsymbol{x}}+{\boldsymbol{s}}/2-\overline{{\boldsymbol{x}}}_{\sigma_{s}}(t+\tau/2))\delta({\boldsymbol{p}}-\overline{{\boldsymbol{p}}}_{\sigma_{s}}(t+\tau/2))}}_{\overline{\mathfrak{f}}_{s}(t+\tau/2,{\boldsymbol{x}}+{\boldsymbol{s}}/2,{\boldsymbol{p}})/N_{s}}
×δ(𝒙𝒔/2𝒙¯σs(tτ/2))δ(𝒑𝒑¯σs(tτ/2))𝔣¯s(tτ/2,𝒙𝒔/2,𝒑)/Ns\displaystyle\hphantom{\sum_{\sigma_{s}\neq\sigma^{\prime}_{s^{\prime}}}}\times\underbrace{\braket{\delta({\boldsymbol{x}}-{\boldsymbol{s}}/2-\overline{{\boldsymbol{x}}}_{\sigma^{\prime}_{s^{\prime}}}(t-\tau/2))\delta({\boldsymbol{p}}^{\prime}-\overline{{\boldsymbol{p}}}_{\sigma^{\prime}_{s^{\prime}}}(t-\tau/2))}}_{\overline{\mathfrak{f}}_{s^{\prime}}(t-\tau/2,{\boldsymbol{x}}-{\boldsymbol{s}}/2,{\boldsymbol{p}}^{\prime})/N_{s^{\prime}}}
=𝔣¯NsNsσs,σs(1δssδσsσs)=(1Ns1δss)𝔣¯𝔣¯.\displaystyle=\frac{\mathfrak{C}_{\overline{\mathfrak{f}}}}{N_{s}N_{s^{\prime}}}\sum_{\sigma_{s},\sigma^{\prime}_{s^{\prime}}}(1-\delta_{ss^{\prime}}\delta_{\sigma_{s}\sigma^{\prime}_{s^{\prime}}})=(1-N_{s}^{-1}\delta_{ss^{\prime}})\mathfrak{C}_{\overline{\mathfrak{f}}}\approx\mathfrak{C}_{\overline{\mathfrak{f}}}. (555)

Let us also use 𝒑¯σs(t+τ/2)𝒑¯σs(t)\smash{\overline{{\boldsymbol{p}}}_{\sigma_{s}}(t+\tau/2)\approx\overline{{\boldsymbol{p}}}_{\sigma_{s}}(t)}. Then,

ssδssδ(𝒑𝒑)σ=1Ns\displaystyle\mathfrak{C}_{ss^{\prime}}\approx\delta_{ss^{\prime}}\delta({\boldsymbol{p}}-{\boldsymbol{p}}^{\prime})\sum_{\sigma=1}^{N_{s}}\langle δ(𝒙+𝒔/2𝒙¯σ(t+τ/2))\displaystyle\delta({\boldsymbol{x}}+{\boldsymbol{s}}/2-\overline{{\boldsymbol{x}}}_{\sigma}(t+\tau/2))
×δ(𝒙𝒔/2𝒙¯σ(tτ/2))δ(𝒑𝒑¯σ(t)).\displaystyle\times\delta({\boldsymbol{x}}-{\boldsymbol{s}}/2-\overline{{\boldsymbol{x}}}_{\sigma}(t-\tau/2))\delta({\boldsymbol{p}}-\overline{{\boldsymbol{p}}}_{\sigma}(t))\rangle. (556)

Next, notice that

δ(𝒙+𝒔/2𝒙¯σ(t+τ/2))δ(𝒙𝒔/2𝒙¯σ(tτ/2))δ(𝒑𝒑¯σ(t))\displaystyle\braket{\delta({\boldsymbol{x}}+{\boldsymbol{s}}/2-\overline{{\boldsymbol{x}}}_{\sigma}(t+\tau/2))\delta({\boldsymbol{x}}-{\boldsymbol{s}}/2-\overline{{\boldsymbol{x}}}_{\sigma}(t-\tau/2))\delta({\boldsymbol{p}}-\overline{{\boldsymbol{p}}}_{\sigma}(t))}
=δ(𝒔+𝒙¯σ(tτ/2)𝒙¯σ(t+τ/2))δ(𝒙𝒔/2𝒙¯σ(tτ/2))δ(𝒑𝒑¯σ(t))\displaystyle=\braket{\delta({\boldsymbol{s}}+\overline{{\boldsymbol{x}}}_{\sigma}(t-\tau/2)-\overline{{\boldsymbol{x}}}_{\sigma}(t+\tau/2))\delta({\boldsymbol{x}}-{\boldsymbol{s}}/2-\overline{{\boldsymbol{x}}}_{\sigma}(t-\tau/2))\delta({\boldsymbol{p}}-\overline{{\boldsymbol{p}}}_{\sigma}(t))}
=δ(𝒔+𝒙¯σ(tτ/2)𝒙¯σ(t+τ/2))δ(𝒙(𝒙¯σ(t+τ/2)+𝒙¯σ(tτ/2))/2)δ(𝒑𝒑¯σ(t))\displaystyle=\braket{\delta({\boldsymbol{s}}+\overline{{\boldsymbol{x}}}_{\sigma}(t-\tau/2)-\overline{{\boldsymbol{x}}}_{\sigma}(t+\tau/2))\delta({\boldsymbol{x}}-(\overline{{\boldsymbol{x}}}_{\sigma}(t+\tau/2)+\overline{{\boldsymbol{x}}}_{\sigma}(t-\tau/2))/2)\delta({\boldsymbol{p}}-\overline{{\boldsymbol{p}}}_{\sigma}(t))}
δ(𝒔𝒗s(t,𝒙¯σ,𝒑¯σ)τ)δ(𝒙𝒙¯σ(t))δ(𝒑𝒑¯σ(t))\displaystyle\approx\braket{\delta({\boldsymbol{s}}-{\boldsymbol{v}}_{s}(t,\overline{{\boldsymbol{x}}}_{\sigma},\overline{{\boldsymbol{p}}}_{\sigma})\tau)\delta({\boldsymbol{x}}-\overline{{\boldsymbol{x}}}_{\sigma}(t))\delta({\boldsymbol{p}}-\overline{{\boldsymbol{p}}}_{\sigma}(t))}
δ(𝒔𝒗s(t,𝒙,𝒑)τ)δ(𝒙𝒙¯σ(t))δ(𝒑𝒑¯σ(t))\displaystyle\approx\delta({\boldsymbol{s}}-{\boldsymbol{v}}_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}})\tau)\braket{\delta({\boldsymbol{x}}-\overline{{\boldsymbol{x}}}_{\sigma}(t))\delta({\boldsymbol{p}}-\overline{{\boldsymbol{p}}}_{\sigma}(t))}
=δ(𝒔𝒗s(t,𝒙,𝒑)τ)𝔣¯s(t,𝒙,𝒑)/Ns.\displaystyle=\delta({\boldsymbol{s}}-{\boldsymbol{v}}_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}})\tau)\overline{\mathfrak{f}}_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}})/N_{s}. (557)

Hence,

ss=δssδ(𝒑𝒑)δ(𝒔𝒗s(t,𝒙,𝒑)τ)Fs(t,𝒙,𝒑),\displaystyle\mathfrak{C}_{ss^{\prime}}=\delta_{ss^{\prime}}\delta({\boldsymbol{p}}-{\boldsymbol{p}}^{\prime})\delta({\boldsymbol{s}}-{\boldsymbol{v}}_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}})\tau)F_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}}), (558)

where we used 𝔣¯sFs\smash{\overline{\mathfrak{f}}_{s}\approx F_{s}}. Therefore,

𝔊ss(t,𝒙,ω,𝒌;𝒑,𝒑)\displaystyle\mathfrak{G}_{ss^{\prime}}(t,{\boldsymbol{x}},\omega,{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime}) =dτ2\upid𝒔(2\upi)neiωτi𝒌𝒔ss(t,𝒙,τ,𝒔;𝒑,𝒑)\displaystyle=\int\frac{\mathrm{d}\tau}{2\upi}\frac{\mathrm{d}{\boldsymbol{s}}}{(2\upi)^{n}}\,\mathrm{e}^{\mathrm{i}\omega\tau-\mathrm{i}{\boldsymbol{k}}\cdot{\boldsymbol{s}}}\,\mathfrak{C}_{ss^{\prime}}(t,{\boldsymbol{x}},\tau,{\boldsymbol{s}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})
1(2\upi)nδssδ(𝒑𝒑)Fs(t,𝒙,𝒑).\displaystyle\approx\frac{1}{(2\upi)^{n}}\,\delta_{ss^{\prime}}\delta({\boldsymbol{p}}-{\boldsymbol{p}}^{\prime})F_{s}(t,{\boldsymbol{x}},{\boldsymbol{p}}). (559)

B.8 Proof of (494)

Using the symmetry 𝖴αβγδ=𝖴βαγδ=𝖴βαδγ\smash{\mathsf{U}^{\alpha\beta\gamma\delta}=\mathsf{U}^{\beta\alpha\gamma\delta}=\mathsf{U}^{\beta\alpha\delta\gamma}}, one readily obtains from (254) that

Δ=12P0d𝗸g¯βγpαpδ𝖴αβγδ+18d𝗸𝒥,\displaystyle\displaystyle\Delta=\frac{1}{2P^{0}}\int\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\overline{g}_{\beta\gamma}p_{\alpha}p_{\delta}\mathsf{U}^{\alpha\beta\gamma\delta}+\frac{1}{8}\fint\mathrm{d}{\boldsymbol{\mathsf{k}}}\,\mathcal{J}, (560)
𝒥=𝒑(𝒌𝔈P0ϖ)1(P0)2𝔈p0+g¯00𝔈(P0)3,\displaystyle\displaystyle\mathcal{J}=-\frac{\partial^{\prime}}{\partial{\boldsymbol{p}}}\cdot\left(\frac{{\boldsymbol{k}}\mathfrak{E}}{P^{0}\varpi}\right)-\frac{1}{(P^{0})^{2}}\frac{\partial\mathfrak{E}}{\partial p_{0}}+\frac{\overline{g}^{00}\mathfrak{E}}{(P^{0})^{3}}, (561)

where ϖkρpρ=P0(𝒌𝒗ω)\smash{\varpi\doteq k_{\rho}p^{\rho}=P^{0}({\boldsymbol{k}}\cdot{\boldsymbol{v}}-\omega)} and the prime in \smash{\partial^{\prime}} denotes that p0\smash{p_{0}} is considered as a function of 𝒑\smash{{\boldsymbol{p}}} at differentiation. One can also write this as follows:

𝒥=𝒌ϖ(P0𝒑)𝔈(P0)21P0𝒑(𝒌𝔈ϖ)1(P0)2𝔈p0+g¯00𝔈(P0)3.\displaystyle\mathcal{J}=\frac{{\boldsymbol{k}}}{\varpi}\cdot\left(\frac{\partial^{\prime}P_{0}}{\partial{\boldsymbol{p}}}\right)\frac{\mathfrak{E}}{(P^{0})^{2}}-\frac{1}{P^{0}}\frac{\partial^{\prime}}{\partial{\boldsymbol{p}}}\cdot\left(\frac{{\boldsymbol{k}}\mathfrak{E}}{\varpi}\right)-\frac{1}{(P^{0})^{2}}\frac{\partial\mathfrak{E}}{\partial p_{0}}+\frac{\overline{g}^{00}\mathfrak{E}}{(P^{0})^{3}}. (562)

As shown in (Garg & Dodin, 2020, appendix B), the following equality is satisfied:

𝒌ϖ(P0𝒑)=1ϖϖp0g¯00P0.\displaystyle\frac{{\boldsymbol{k}}}{\varpi}\cdot\left(\frac{\partial^{\prime}P_{0}}{\partial{\boldsymbol{p}}}\right)=\frac{1}{\varpi}\frac{\partial\varpi}{\partial p_{0}}-\frac{\overline{g}^{00}}{P^{0}}. (563)

Also notice that

𝒑(𝒌𝔈ϖ)\displaystyle\frac{\partial^{\prime}}{\partial{\boldsymbol{p}}}\cdot\left(\frac{{\boldsymbol{k}}\mathfrak{E}}{\varpi}\right) =𝒑(𝒌𝔈ϖ)+P0𝒑p0(𝒌𝔈ϖ)\displaystyle=\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left(\frac{{\boldsymbol{k}}\mathfrak{E}}{\varpi}\right)+\frac{\partial P_{0}}{\partial{\boldsymbol{p}}}\cdot\frac{\partial}{\partial p_{0}}\left(\frac{{\boldsymbol{k}}\mathfrak{E}}{\varpi}\right)
=𝒑(𝒌𝔈ϖ)𝒌𝒗p0(𝔈ϖ),\displaystyle=\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left(\frac{{\boldsymbol{k}}\mathfrak{E}}{\varpi}\right)-{\boldsymbol{k}}\cdot{\boldsymbol{v}}\,\frac{\partial}{\partial p_{0}}\left(\frac{\mathfrak{E}}{\varpi}\right), (564)

where we used Hamilton’s equation 𝒑P0=𝒑H=𝒗\smash{\partial_{{\boldsymbol{p}}}P_{0}=-\partial_{{\boldsymbol{p}}}H=-{\boldsymbol{v}}}. Therefore,

𝒥=1ϖϖp0𝔈(P0)21P0𝒑(𝒌𝔈ϖ)+𝒌𝒗P0p0(𝔈ϖ)1(P0)2𝔈p0.\displaystyle\mathcal{J}=\frac{1}{\varpi}\frac{\partial\varpi}{\partial p_{0}}\frac{\mathfrak{E}}{(P^{0})^{2}}-\frac{1}{P^{0}}\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left(\frac{{\boldsymbol{k}}\mathfrak{E}}{\varpi}\right)+\frac{{\boldsymbol{k}}\cdot{\boldsymbol{v}}}{P^{0}}\frac{\partial}{\partial p_{0}}\left(\frac{\mathfrak{E}}{\varpi}\right)-\frac{1}{(P^{0})^{2}}\frac{\partial\mathfrak{E}}{\partial p_{0}}. (565)

The first and the last terms can be merged; then, one obtains

𝒥\displaystyle\mathcal{J} =1P0𝒑(𝒌𝔈ϖ)+𝒌𝒗P0p0(𝔈ϖ)ϖ(P0)2p0(𝔈ϖ)\displaystyle=-\frac{1}{P^{0}}\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left(\frac{{\boldsymbol{k}}\mathfrak{E}}{\varpi}\right)+\frac{{\boldsymbol{k}}\cdot{\boldsymbol{v}}}{P^{0}}\frac{\partial}{\partial p_{0}}\left(\frac{\mathfrak{E}}{\varpi}\right)-\frac{\varpi}{(P^{0})^{2}}\frac{\partial}{\partial p_{0}}\left(\frac{\mathfrak{E}}{\varpi}\right)
=1P0𝒑(𝒌𝔈ϖ)+ωP0p0(𝔈ϖ)\displaystyle=-\frac{1}{P^{0}}\frac{\partial}{\partial{\boldsymbol{p}}}\cdot\left(\frac{{\boldsymbol{k}}\mathfrak{E}}{\varpi}\right)+\frac{\omega}{P^{0}}\frac{\partial}{\partial p_{0}}\left(\frac{\mathfrak{E}}{\varpi}\right)
=1P0pλ(kλ𝔈ϖ).\displaystyle=-\frac{1}{P^{0}}\frac{\partial}{\partial p_{\lambda}}\left(\frac{k_{\lambda}\mathfrak{E}}{\varpi}\right). (566)

In combination with (560), this leads to (494).

Appendix C Properties of the collision operator

Here, we prove the properties of the collision operator discussed in section 6.8. To shorten the calculations, we introduce two auxiliary functions,

𝒵ss(𝒌;𝒑,𝒑)\upiδ(𝒌𝒗s𝒌𝒗s)𝒬ss(𝒌𝒗s,𝒌;𝒑,𝒑),\displaystyle\displaystyle\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\doteq\upi\,\delta({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})\mathcal{Q}_{ss^{\prime}}({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s},{\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime}),
ss(𝒑,𝒑)Fs(𝒑)pjFs(𝒑)Fs(𝒑)Fs(𝒑)pj,\displaystyle\displaystyle\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\doteq\frac{\partial F_{s}({\boldsymbol{p}})}{\partial p_{j}}\,F_{s^{\prime}}({\boldsymbol{p}}^{\prime})-F_{s}({\boldsymbol{p}})\,\frac{\partial F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial p_{j}^{\prime}}, (567)

which have the following properties:

𝒵ss(𝒌;𝒑,𝒑)=𝒵ss(𝒌;𝒑,𝒑),ss(𝒑,𝒑)=ss(𝒑,𝒑).\displaystyle\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})=\mathcal{Z}_{s^{\prime}s}({\boldsymbol{k}};{\boldsymbol{p}}^{\prime},{\boldsymbol{p}}),\qquad\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})=-\mathcal{F}_{s^{\prime}s}({\boldsymbol{p}}^{\prime},{\boldsymbol{p}}). (568)

C.1 Momentum conservation

Momentum conservation is proven as follows. Using integration by parts, one obtains

sd𝒑pl𝒞s\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}\mathcal{C}_{s}
=s,sd𝒑plpid𝒌(2\upi)nd𝒑kikj𝒵ss(𝒌;𝒑,𝒑)ss(𝒑,𝒑)\displaystyle=\sum_{s,s^{\prime}}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}\,\frac{\partial}{\partial p_{i}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,k_{i}k_{j}\,\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})
=s,sd𝒌(2\upi)nd𝒑d𝒑klkj𝒵ss(𝒌;𝒑,𝒑)ss(𝒑,𝒑).\displaystyle=-\sum_{s,s^{\prime}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,k_{l}k_{j}\,\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime}). (569)

Now we swap the dummy variables ss\smash{s\leftrightarrow s^{\prime}} and 𝒑𝒑\smash{{\boldsymbol{p}}\leftrightarrow{\boldsymbol{p}}^{\prime}} and then apply (568):

sd𝒑pl𝒞s\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}\mathcal{C}_{s}
=s,sd𝒌(2\upi)nd𝒑d𝒑klkj𝒵ss(𝒌;𝒑,𝒑)ss(𝒑,𝒑)\displaystyle=-\sum_{s^{\prime},s}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,\mathrm{d}{\boldsymbol{p}}\,k_{l}k_{j}\,\mathcal{Z}_{s^{\prime}s}({\boldsymbol{k}};{\boldsymbol{p}}^{\prime},{\boldsymbol{p}})\mathcal{F}_{s^{\prime}s}({\boldsymbol{p}}^{\prime},{\boldsymbol{p}})
=s,sd𝒌(2\upi)nd𝒑d𝒑klkj𝒵ss(𝒌;𝒑,𝒑)ss(𝒑,𝒑).\displaystyle=\sum_{s^{\prime},s}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,\mathrm{d}{\boldsymbol{p}}\,k_{l}k_{j}\,\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime}). (570)

The expression on the right-hand side of (570) is minus that in (569). Hence, both are zero, which proves that sd𝒑pl𝒞s=0\smash{\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}\mathcal{C}_{s}=0}.

C.2 Energy conservation

Energy conservation is proven similarly, using that vsi=s/pi\smash{v_{s}^{i}=\partial\mathcal{H}_{s}/\partial p_{i}} and the fact that 𝒌𝒗s\smash{{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}} and 𝒌𝒗s\smash{{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}}} are interchangeable due to the presence of δ(𝒌𝒗s𝒌𝒗s)\smash{\delta({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}-{\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})} in 𝒵ss\smash{\mathcal{Z}_{ss^{\prime}}}:

sd𝒑s𝒞s\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\mathcal{C}_{s}
=s,sd𝒑spid𝒌(2\upi)nd𝒑kikj𝒵ss(𝒌;𝒑,𝒑)ss(𝒑,𝒑)\displaystyle=\sum_{s,s^{\prime}}\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\,\frac{\partial}{\partial p_{i}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,k_{i}k_{j}\,\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})
=s,sd𝒌(2\upi)nd𝒑d𝒑(𝒌𝒗s)kj𝒵ss(𝒌;𝒑,𝒑)ss(𝒑,𝒑)\displaystyle=-\sum_{s,s^{\prime}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})k_{j}\,\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})
=s,sd𝒌(2\upi)nd𝒑d𝒑(𝒌𝒗s)kj𝒵ss(𝒌;𝒑,𝒑)ss(𝒑,𝒑)\displaystyle=-\sum_{s^{\prime},s}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,\mathrm{d}{\boldsymbol{p}}\,({\boldsymbol{k}}\cdot{\boldsymbol{v}}^{\prime}_{s^{\prime}})k_{j}\,\mathcal{Z}_{s^{\prime}s}({\boldsymbol{k}};{\boldsymbol{p}}^{\prime},{\boldsymbol{p}})\mathcal{F}_{s^{\prime}s}({\boldsymbol{p}}^{\prime},{\boldsymbol{p}})
=s,sd𝒌(2\upi)nd𝒑d𝒑(𝒌𝒗s)kj𝒵ss(𝒌;𝒑,𝒑)ss(𝒑,𝒑).\displaystyle=\sum_{s^{\prime},s}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,\mathrm{d}{\boldsymbol{p}}\,({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})k_{j}\,\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime}). (571)

Like in the previous case, the third and the fifth lines are minus each other, whence sd𝒑s𝒞s=0\smash{\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\mathcal{C}_{s}=0}.

C.3 H{H}-theorem

From (279) and (280), one has

(dσdt)coll=sd𝒑(1+lnFs(𝒑))𝒞s=sd𝒑lnFs(𝒑)𝒞s,\displaystyle\left(\frac{\mathrm{d}\sigma}{\mathrm{d}t}\right)_{\text{coll}}=-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,(1+\ln F_{s}({\boldsymbol{p}}))\mathcal{C}_{s}=-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\ln F_{s}({\boldsymbol{p}})\mathcal{C}_{s}, (572)

where we used particle conservation, d𝒑𝒞s=0\smash{\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{C}_{s}=0}. Then,

(dσdt)coll\displaystyle\left(\frac{\mathrm{d}\sigma}{\mathrm{d}t}\right)_{\text{coll}} =ssd𝒑lnFs(𝒑)pid𝒌(2\upi)nd𝒑kikj𝒵ss(𝒌;𝒑,𝒑)ss(𝒑,𝒑)\displaystyle=-\sum_{ss^{\prime}}\int\mathrm{d}{\boldsymbol{p}}\,\ln F_{s}({\boldsymbol{p}})\,\frac{\partial}{\partial p_{i}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,k_{i}k_{j}\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})
=ssd𝒌(2\upi)nd𝒑d𝒑kikjlnFs(𝒑)pi𝒵ss(𝒌;𝒑,𝒑)ss(𝒑,𝒑).\displaystyle=\sum_{ss^{\prime}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,k_{i}k_{j}\,\frac{\partial\ln F_{s}({\boldsymbol{p}})}{\partial p_{i}}\,\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime}). (573)

Let us swap the dummy variables ss\smash{s\leftrightarrow s^{\prime}} and 𝒑𝒑\smash{{\boldsymbol{p}}\leftrightarrow{\boldsymbol{p}}^{\prime}} and then apply (568) to obtain

(dσdt)coll=ssd𝒌(2\upi)nd𝒑d𝒑kikjlnFs(𝒑)pi𝒵ss(𝒌;𝒑,𝒑)ss(𝒑,𝒑).\displaystyle\left(\frac{\mathrm{d}\sigma}{\mathrm{d}t}\right)_{\text{coll}}=-\sum_{ss^{\prime}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\int\mathrm{d}{\boldsymbol{p}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,k_{i}k_{j}\,\frac{\partial\ln F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial p_{i}^{\prime}}\,\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime}). (574)

Upon comparing (574) with (573), one can put the result in a symmetrized form:

(dσdt)coll=12ssd𝒌(2\upi)nd𝒑d𝒑kikj\displaystyle\left(\frac{\mathrm{d}\sigma}{\mathrm{d}t}\right)_{\text{coll}}=\frac{1}{2}\sum_{ss^{\prime}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\,k_{i}k_{j} 𝒵ss(𝒌;𝒑,𝒑)ss(𝒑,𝒑)\displaystyle\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})
×(lnFs(𝒑)pilnFs(𝒑)pi).\displaystyle\times\left(\frac{\partial\ln F_{s}({\boldsymbol{p}})}{\partial p_{i}}-\frac{\partial\ln F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial p_{i}^{\prime}}\right). (575)

But notice that

ss(𝒑,𝒑)=(lnFs(𝒑)pjlnFs(𝒑)pj)Fs(𝒑)Fs(𝒑).\displaystyle\mathcal{F}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})=\left(\frac{\partial\ln F_{s}({\boldsymbol{p}})}{\partial p_{j}}-\frac{\partial\ln F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial p_{j}^{\prime}}\right)F_{s}({\boldsymbol{p}})F_{s^{\prime}}({\boldsymbol{p}}^{\prime}). (576)

Thus,

(dσdt)coll=12ssd𝒌(2\upi)nd𝒑d𝒑\displaystyle\left(\frac{\mathrm{d}\sigma}{\mathrm{d}t}\right)_{\text{coll}}=\frac{1}{2}\sum_{ss^{\prime}}\int\frac{\mathrm{d}{\boldsymbol{k}}}{(2\upi)^{n}}\,\mathrm{d}{\boldsymbol{p}}\,\mathrm{d}{\boldsymbol{p}}^{\prime}\, 𝒵ss(𝒌;𝒑,𝒑)Fs(𝒑)Fs(𝒑)\displaystyle\mathcal{Z}_{ss^{\prime}}({\boldsymbol{k}};{\boldsymbol{p}},{\boldsymbol{p}}^{\prime})F_{s}({\boldsymbol{p}})F_{s^{\prime}}({\boldsymbol{p}}^{\prime})
×(𝒌lnFs(𝒑)𝒑𝒌lnFs(𝒑)𝒑)20.\displaystyle\times\left({\boldsymbol{k}}\cdot\frac{\partial\ln F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}-{\boldsymbol{k}}\cdot\frac{\partial\ln F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial{\boldsymbol{p}}^{\prime}}\right)^{2}\geq 0. (577)

Appendix D Conservation laws for on-shell waves

Here, we prove the momentum-conservation theorem (378) and the energy-conservation theorem (380) for QL interactions of plasmas with on-shell waves.

D.1 Momentum conservation

Let us multiply (374) by klk_{l} and integrate over 𝒌{\boldsymbol{k}}. Then, one obtains

0=\displaystyle 0= d𝒌klJt+d𝒌kl(vgiJ)xid𝒌klki(wxiJ)2d𝒌klγJ\displaystyle\int\mathrm{d}{\boldsymbol{k}}\,k_{l}\,\frac{\partial J}{\partial t}+\int\mathrm{d}{\boldsymbol{k}}\,k_{l}\,\frac{\partial(v_{\text{g}}^{i}J)}{\partial x^{i}}-\int\mathrm{d}{\boldsymbol{k}}\,k_{l}\frac{\partial}{\partial k_{i}}\left(\frac{\partial w}{\partial x^{i}}\,J\right)-2\int\mathrm{d}{\boldsymbol{k}}\,k_{l}\gamma J
=\displaystyle= td𝒌klJ+xid𝒌klvgiJ+d𝒌wxlJ2d𝒌wγJ.\displaystyle\,\frac{\partial}{\partial t}\int\mathrm{d}{\boldsymbol{k}}\,k_{l}J+\frac{\partial}{\partial x^{i}}\int\mathrm{d}{\boldsymbol{k}}\,k_{l}v_{\text{g}}^{i}J+\int\mathrm{d}{\boldsymbol{k}}\,\frac{\partial w}{\partial x^{l}}\,J-2\int\mathrm{d}{\boldsymbol{k}}\,w\gamma J. (578)

Similarly, multiplying (375) by s\smash{\mathcal{H}_{s}} and integrating over 𝒑\smash{{\boldsymbol{p}}} yields

0=\displaystyle 0= d𝒑plFst+d𝒑pl(vsiFs)xid𝒑plpi(sxiFs)\displaystyle\int\mathrm{d}{\boldsymbol{p}}\,p_{l}\,\frac{\partial F_{s}}{\partial t}+\int\mathrm{d}{\boldsymbol{p}}\,p_{l}\,\frac{\partial(v_{s}^{i}F_{s})}{\partial x^{i}}-\int\mathrm{d}{\boldsymbol{p}}\,p_{l}\,\frac{\partial}{\partial p_{i}}\left(\frac{\partial\mathcal{H}_{s}}{\partial x^{i}}\,F_{s}\right)
d𝒑plpi(𝖣s,ijFspj)d𝒑pl𝒞s\displaystyle-\int\mathrm{d}{\boldsymbol{p}}\,p_{l}\,\frac{\partial}{\partial p_{i}}\left(\mathsf{D}_{s,ij}\,\frac{\partial F_{s}}{\partial p_{j}}\right)-\int\mathrm{d}{\boldsymbol{p}}\,p_{l}\,\mathcal{C}_{s}
=\displaystyle= td𝒑plFs+xid𝒑plvsiFs+xld𝒑ΔsFs+d𝒑H0sxlFs\displaystyle\,\frac{\partial}{\partial t}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}F_{s}+\frac{\partial}{\partial x^{i}}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}v_{s}^{i}F_{s}+\frac{\partial}{\partial x^{l}}\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}F_{s}+\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial H_{0s}}{\partial x^{l}}\,F_{s}
d𝒑ΔsFsxl+d𝒑𝖣s,ljFspjd𝒑pl𝒞s.\displaystyle-\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}\,\frac{\partial F_{s}}{\partial x^{l}}+\int\mathrm{d}{\boldsymbol{p}}\,\mathsf{D}_{s,lj}\,\frac{\partial F_{s}}{\partial p_{j}}-\int\mathrm{d}{\boldsymbol{p}}\,p_{l}\,\mathcal{C}_{s}. (579)

Let us sum up (579) over species and also add it with (578). The contribution of the collision integral disappears due to (278), so one obtains

0=\displaystyle 0= t(sd𝒑plFs+d𝒌klJ)+xi(sd𝒑plvsiFs+d𝒌klvgiJ)\displaystyle\,\frac{\partial}{\partial t}\left(\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,k_{l}J\right)+\frac{\partial}{\partial x^{i}}\left(\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}v_{s}^{i}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,k_{l}v_{\text{g}}^{i}J\right)
+sxld𝒑ΔsFs+sd𝒑H0sxlFs\displaystyle+\sum_{s}\frac{\partial}{\partial x^{l}}\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}F_{s}+\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial H_{0s}}{\partial x^{l}}\,F_{s}
+sd𝒑𝖣s,ljFspj2d𝒌klγJ\displaystyle+\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\mathsf{D}_{s,lj}\,\frac{\partial F_{s}}{\partial p_{j}}-2\int\mathrm{d}{\boldsymbol{k}}\,k_{l}\gamma J
sd𝒑ΔsFsxl+d𝒌Jwxl.\displaystyle-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}\,\frac{\partial F_{s}}{\partial x^{l}}+\int\mathrm{d}{\boldsymbol{k}}\,J\,\frac{\partial w}{\partial x^{l}}. (580)

Next, notice that

2d𝒌klγJ\displaystyle 2\int\mathrm{d}{\boldsymbol{k}}\,k_{l}\gamma J =2\upisd𝒑d𝒌klkj|𝜶s𝜼|2ωΛJδ(w𝒌𝒗s)Fspj\displaystyle=2\upi\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\mathrm{d}{\boldsymbol{k}}\,k_{l}k_{j}\,\frac{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}{\partial_{\omega}\Lambda}\,J\delta(w-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})\,\frac{\partial F_{s}}{\partial p_{j}}
=sd𝒑𝖣s,ljFspj.\displaystyle=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\mathsf{D}_{s,lj}\,\frac{\partial F_{s}}{\partial p_{j}}. (581)

Also, assuming that 𝚵0\smash{{\boldsymbol{\Xi}}_{0}}, |𝜶s𝜼|2\smash{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}, and 𝜼s𝜼\smash{{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}}} are independent of 𝒙\smash{{\boldsymbol{x}}} and using (371), one gets

sd𝒑ΔsFsxl\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}\frac{\partial F_{s}}{\partial x^{l}}
=sd𝒑Fsxl(pidωd𝒌ki2(ω𝒌𝒗s)|𝜶s𝜼|2(h(𝒌)+h(𝒌))δ(ωw(𝒌))\displaystyle=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial F_{s}}{\partial x^{l}}\bigg{(}\frac{\partial}{\partial p_{i}}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\frac{k_{i}}{2(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})}\,|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\,\delta(\omega-w({\boldsymbol{k}}))
+12dωd𝒌(𝜼s𝜼)(h(𝒌)+h(𝒌))δ(ωw(𝒌)))\displaystyle\hphantom{=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial F_{s}}{\partial x^{l}}\bigg{(}}+\frac{1}{2}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}})(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\,\delta(\omega-w({\boldsymbol{k}}))\bigg{)}
=12sdωd𝒌d𝒑(h(𝒌)+h(𝒌))δ(ωw(𝒌))ki|𝜶s𝜼|2ω𝒌𝒗s2Fsxlpi\displaystyle=-\frac{1}{2}\sum_{s}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\mathrm{d}{\boldsymbol{p}}\,(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\,\delta(\omega-w({\boldsymbol{k}}))\,\frac{k_{i}|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}\,\frac{\partial^{2}F_{s}}{\partial x^{l}\partial p_{i}}
+12sdωd𝒌d𝒑(h(𝒌)+h(𝒌))δ(ωw(𝒌))xl(𝜼s𝜼Fs)\displaystyle\hphantom{\,=\,}+\frac{1}{2}\sum_{s}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\mathrm{d}{\boldsymbol{p}}\,(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\,\delta(\omega-w({\boldsymbol{k}}))\,\frac{\partial}{\partial x^{l}}\,({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}}F_{s})
=sdωd𝒌h(𝒌)δ(ωw(𝒌))xld𝒑(ki|𝜶s𝜼|2ω𝒌𝒗sFspi𝜼s𝜼Fs)\displaystyle=-\sum_{s}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\,\delta(\omega-w({\boldsymbol{k}}))\,\frac{\partial}{\partial x^{l}}\fint\mathrm{d}{\boldsymbol{p}}\,\left(\frac{k_{i}|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}\frac{\partial F_{s}}{\partial p_{i}}-{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}}F_{s}\right)
=sdωd𝒌h(𝒌)δ(ωw(𝒌))(𝜼𝚵𝜼)xl\displaystyle=-\sum_{s}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\,\delta(\omega-w({\boldsymbol{k}}))\,\frac{\partial({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}{\boldsymbol{\eta}})}{\partial x^{l}}
=sd𝒌h(𝒌)Λ(w(𝒌),𝒌)xl\displaystyle=-\sum_{s}\int\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\,\frac{\partial\Lambda(w({\boldsymbol{k}}),{\boldsymbol{k}})}{\partial x^{l}}
=sd𝒌Jwxl,\displaystyle=\sum_{s}\int\mathrm{d}{\boldsymbol{k}}\,J\,\frac{\partial w}{\partial x^{l}}, (582)

where we also used (366b). Substituting (581) and (582) into (580) leads to (378).

D.2 Energy conservation

Let us multiply (374) by ww and integrate over 𝒌{\boldsymbol{k}}. Then, one obtains

0=\displaystyle 0= d𝒌wJt+d𝒌w(vgiJ)xid𝒌wki(wxiJ)2d𝒌wγJ\displaystyle\int\mathrm{d}{\boldsymbol{k}}\,w\,\frac{\partial J}{\partial t}+\int\mathrm{d}{\boldsymbol{k}}\,w\,\frac{\partial(v_{\text{g}}^{i}J)}{\partial x^{i}}-\int\mathrm{d}{\boldsymbol{k}}\,w\,\frac{\partial}{\partial k_{i}}\left(\frac{\partial w}{\partial x^{i}}J\right)-2\int\mathrm{d}{\boldsymbol{k}}\,w\gamma J
=\displaystyle= td𝒌wJd𝒌wtJ+xid𝒌wvgiJd𝒌wxivgiJ\displaystyle\,\frac{\partial}{\partial t}\int\mathrm{d}{\boldsymbol{k}}\,wJ-\int\mathrm{d}{\boldsymbol{k}}\,\frac{\partial w}{\partial t}\,J+\frac{\partial}{\partial x^{i}}\int\mathrm{d}{\boldsymbol{k}}\,wv_{\text{g}}^{i}J-\int\mathrm{d}{\boldsymbol{k}}\,\frac{\partial w}{\partial x^{i}}\,v_{\text{g}}^{i}J
+d𝒌vgiwxiJ2d𝒌wγJ\displaystyle+\int\mathrm{d}{\boldsymbol{k}}\,v_{\text{g}}^{i}\,\frac{\partial w}{\partial x^{i}}\,J-2\int\mathrm{d}{\boldsymbol{k}}\,w\gamma J
=\displaystyle= td𝒌wJ+xid𝒌wvgiJd𝒌wtJ2d𝒌wγJ.\displaystyle\,\frac{\partial}{\partial t}\int\mathrm{d}{\boldsymbol{k}}\,wJ+\frac{\partial}{\partial x^{i}}\int\mathrm{d}{\boldsymbol{k}}\,wv_{\text{g}}^{i}J-\int\mathrm{d}{\boldsymbol{k}}\,\frac{\partial w}{\partial t}\,J-2\int\mathrm{d}{\boldsymbol{k}}\,w\gamma J. (583)

Similarly, multiplying (375) by s\smash{\mathcal{H}_{s}} and integrating over 𝒑\smash{{\boldsymbol{p}}} yields

0=\displaystyle 0= d𝒑sFst+d𝒑s(vsiFs)xid𝒑spi(sxiFs)\displaystyle\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\,\frac{\partial F_{s}}{\partial t}+\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\,\frac{\partial(v_{s}^{i}F_{s})}{\partial x^{i}}-\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\,\frac{\partial}{\partial p_{i}}\left(\frac{\partial\mathcal{H}_{s}}{\partial x^{i}}\,F_{s}\right)
d𝒑spi(𝖣s,ijFspj)d𝒑s𝒞s\displaystyle-\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\,\frac{\partial}{\partial p_{i}}\left(\mathsf{D}_{s,ij}\,\frac{\partial F_{s}}{\partial p_{j}}\right)-\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\,\mathcal{C}_{s}
=\displaystyle= td𝒑sFsd𝒑stFs+xid𝒑svsiFsd𝒑sxivsiFs\displaystyle\,\frac{\partial}{\partial t}\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}F_{s}-\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial\mathcal{H}_{s}}{\partial t}\,F_{s}+\frac{\partial}{\partial x^{i}}\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}v_{s}^{i}F_{s}-\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial\mathcal{H}_{s}}{\partial x^{i}}\,v_{s}^{i}F_{s}
+d𝒑vsisxiFs+d𝒑vsi𝖣s,ijFspjd𝒑s𝒞s\displaystyle+\int\mathrm{d}{\boldsymbol{p}}\,v_{s}^{i}\,\frac{\partial\mathcal{H}_{s}}{\partial x^{i}}\,F_{s}+\int\mathrm{d}{\boldsymbol{p}}\,v_{s}^{i}\mathsf{D}_{s,ij}\,\frac{\partial F_{s}}{\partial p_{j}}-\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\,\mathcal{C}_{s}
=\displaystyle= td𝒑sFs+xid𝒑svsiFsd𝒑H0stFsd𝒑ΔstFs\displaystyle\,\frac{\partial}{\partial t}\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}F_{s}+\frac{\partial}{\partial x^{i}}\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}v_{s}^{i}F_{s}-\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial H_{0s}}{\partial t}\,F_{s}-\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial\Delta_{s}}{\partial t}\,F_{s}
+d𝒑vsi𝖣s,ijFspjd𝒑s𝒞s\displaystyle+\int\mathrm{d}{\boldsymbol{p}}\,v_{s}^{i}\mathsf{D}_{s,ij}\,\frac{\partial F_{s}}{\partial p_{j}}-\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\,\mathcal{C}_{s}
=\displaystyle= td𝒑H0sFs+xid𝒑H0svsiFs+xid𝒑ΔsvsiFsd𝒑H0stFs\displaystyle\,\frac{\partial}{\partial t}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s}+\frac{\partial}{\partial x^{i}}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}v_{s}^{i}F_{s}+\frac{\partial}{\partial x^{i}}\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}v_{s}^{i}F_{s}-\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial H_{0s}}{\partial t}\,F_{s}
+d𝒑ΔsFst+d𝒑vsi𝖣s,ijFspjd𝒑s𝒞s.\displaystyle+\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}\,\frac{\partial F_{s}}{\partial t}+\int\mathrm{d}{\boldsymbol{p}}\,v_{s}^{i}\mathsf{D}_{s,ij}\,\frac{\partial F_{s}}{\partial p_{j}}-\int\mathrm{d}{\boldsymbol{p}}\,\mathcal{H}_{s}\,\mathcal{C}_{s}. (584)

Let us sum up (584) over species and also add it with (583). The contribution of the collision integral disappears due to (278), so one obtains

0=\displaystyle 0= t(sd𝒑H0sFs+d𝒌wJ)+xi(sd𝒑H0svsiFs+d𝒌wvgiJ)\displaystyle\,\frac{\partial}{\partial t}\left(\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,wJ\right)+\frac{\partial}{\partial x^{i}}\left(\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}v_{s}^{i}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,wv_{\text{g}}^{i}J\right)
+xisd𝒑ΔsvsiFssd𝒑H0stFs\displaystyle+\frac{\partial}{\partial x^{i}}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}v_{s}^{i}F_{s}-\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial H_{0s}}{\partial t}\,F_{s}
+sd𝒑vsi𝖣s,ijFspj2d𝒌wγJ\displaystyle+\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,v_{s}^{i}\mathsf{D}_{s,ij}\,\frac{\partial F_{s}}{\partial p_{j}}-2\int\mathrm{d}{\boldsymbol{k}}\,w\gamma J
+sd𝒑ΔsFstd𝒌Jwt.\displaystyle+\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}\,\frac{\partial F_{s}}{\partial t}-\int\mathrm{d}{\boldsymbol{k}}\,J\,\frac{\partial w}{\partial t}. (585)

Next, notice that

2d𝒌wγJ\displaystyle 2\int\mathrm{d}{\boldsymbol{k}}\,w\gamma J =2\upisd𝒑d𝒌wkj|𝜶s𝜼|2ωΛJδ(w𝒌𝒗s)Fspj\displaystyle=2\upi\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\mathrm{d}{\boldsymbol{k}}\,wk_{j}\,\frac{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}{\partial_{\omega}\Lambda}\,J\delta(w-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})\,\frac{\partial F_{s}}{\partial p_{j}}
=sd𝒑vsi𝖣s,ijFspj.\displaystyle=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,v_{s}^{i}\mathsf{D}_{s,ij}\,\frac{\partial F_{s}}{\partial p_{j}}. (586)

Also, assuming that 𝚵0\smash{{\boldsymbol{\Xi}}_{0}}, |𝜶s𝜼|2\smash{|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}, and 𝜼s𝜼\smash{{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}}} are independent of t\smash{t} and using (371), one gets

sd𝒑ΔsFst\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\Delta_{s}\frac{\partial F_{s}}{\partial t}
=sd𝒑Fst(pidωd𝒌ki2(ω𝒌𝒗s)|𝜶s𝜼|2(h(𝒌)+h(𝒌))δ(ωw(𝒌))\displaystyle=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial F_{s}}{\partial t}\,\bigg{(}\frac{\partial}{\partial p_{i}}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\frac{k_{i}}{2(\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})}\,|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\,\delta(\omega-w({\boldsymbol{k}}))
+12dωd𝒌(𝜼s𝜼)(h(𝒌)+h(𝒌))δ(ωw(𝒌)))\displaystyle\hphantom{=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\frac{\partial F_{s}}{\partial t}\bigg{(}}+\frac{1}{2}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}})(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\,\delta(\omega-w({\boldsymbol{k}}))\bigg{)}
=12sdωd𝒌d𝒑(h(𝒌)+h(𝒌))δ(ωw(𝒌))ki|𝜶s𝜼|2ω𝒌𝒗s2Fstpi\displaystyle=-\frac{1}{2}\sum_{s}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\mathrm{d}{\boldsymbol{p}}\,(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\,\delta(\omega-w({\boldsymbol{k}}))\,\frac{k_{i}|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}\frac{\partial^{2}F_{s}}{\partial t\partial p_{i}}
+12sdωd𝒌d𝒑(h(𝒌)+h(𝒌))δ(ωw(𝒌))t(𝜼s𝜼Fs)\displaystyle\hphantom{\,=\,}+\frac{1}{2}\sum_{s}\fint\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\mathrm{d}{\boldsymbol{p}}\,(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))\,\delta(\omega-w({\boldsymbol{k}}))\,\frac{\partial}{\partial t}\,({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}}F_{s})
=sdωd𝒌h(𝒌)δ(ωw(𝒌))td𝒑(ki|𝜶s𝜼|2ω𝒌𝒗sFspi𝜼s𝜼Fs)\displaystyle=-\sum_{s}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\,\delta(\omega-w({\boldsymbol{k}}))\,\frac{\partial}{\partial t}\fint\mathrm{d}{\boldsymbol{p}}\left(\frac{k_{i}|{\boldsymbol{\alpha}}_{s}^{\dagger}{\boldsymbol{\eta}}|^{2}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}\frac{\partial F_{s}}{\partial p_{i}}-{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\wp}}_{s}{\boldsymbol{\eta}}F_{s}\right)
=sdωd𝒌h(𝒌)δ(ωw(𝒌))(𝜼𝚵𝜼)t\displaystyle=-\sum_{s}\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\,\delta(\omega-w({\boldsymbol{k}}))\,\frac{\partial({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\Xi}}{\boldsymbol{\eta}})}{\partial t}
=sd𝒌h(𝒌)Λ(w(𝒌),𝒌)t\displaystyle=-\sum_{s}\int\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\,\frac{\partial\Lambda(w({\boldsymbol{k}}),{\boldsymbol{k}})}{\partial t}
=sd𝒌Jwt,\displaystyle=\sum_{s}\int\mathrm{d}{\boldsymbol{k}}\,J\,\frac{\partial w}{\partial t}, (587)

where we also used (366a). Substituting (586) and (587) into (585) leads to (380).

Appendix E Uniqueness of the entropy-preserving distribution

Here, we prove that the Boltzmann–Gibbs distribution is the only distribution for which the entropy density σ\smash{\sigma} is conserved. According to (C.3), σ\smash{\sigma} is conserved when

δ(𝒌(𝒗s𝒗s))(𝒌𝑮ss(𝒑,𝒑))2=0\displaystyle\delta({\boldsymbol{k}}\cdot({\boldsymbol{v}}_{s}-{\boldsymbol{v}}^{\prime}_{s^{\prime}}))\,({\boldsymbol{k}}\cdot{\boldsymbol{G}}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime}))^{2}=0 (588)

(for all 𝒑\smash{{\boldsymbol{p}}}, 𝒑\smash{{\boldsymbol{p}}^{\prime}}, and 𝒌\smash{{\boldsymbol{k}}}, as well as all s\smash{s} and s\smash{s^{\prime}}), where

𝑮ss(𝒑,𝒑)lnFs(𝒑)𝒑lnFs(𝒑)𝒑.\displaystyle{\boldsymbol{G}}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})\doteq\frac{\partial\ln F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}-\frac{\partial\ln F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial{\boldsymbol{p}}^{\prime}}. (589)

Let us decompose the vector 𝑮ss(𝒑,𝒑)\smash{{\boldsymbol{G}}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})} into components parallel and perpendicular to the vector 𝒗s𝒗s\smash{{\boldsymbol{v}}_{s}-{\boldsymbol{v}}^{\prime}_{s^{\prime}}}:

𝑮ss(𝒑,𝒑)=λss(𝒗s,𝒗s)(𝒗s𝒗s)+𝑮ss(𝒑,𝒑),\displaystyle{\boldsymbol{G}}_{ss^{\prime}}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})=\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,({\boldsymbol{v}}_{s}-{\boldsymbol{v}}^{\prime}_{s^{\prime}})+{\boldsymbol{G}}_{ss^{\prime}}^{\perp}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime}), (590)

where λss(𝒗s,𝒗s)\smash{\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})} is a scalar function. (Because the velocities are functions of the momenta, one can as well consider λss\smash{\lambda_{ss^{\prime}}} as a function of 𝒑\smash{{\boldsymbol{p}}} and 𝒑\smash{{\boldsymbol{p}}^{\prime}}.) Due to the presence of the delta function in (588), the contribution of the first term to (588) is zero, so (588) can be written as

δ(𝒌(𝒗s𝒗s))(𝒌𝑮ss(𝒑,𝒑))2=0.\displaystyle\delta({\boldsymbol{k}}\cdot({\boldsymbol{v}}_{s}-{\boldsymbol{v}}^{\prime}_{s^{\prime}}))\,({\boldsymbol{k}}\cdot{\boldsymbol{G}}_{ss^{\prime}}^{\perp}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime}))^{2}=0. (591)

By considering this formula for 𝒌\smash{{\boldsymbol{k}}} parallel to 𝑮ss(𝒑,𝒑)\smash{{\boldsymbol{G}}_{ss^{\prime}}^{\perp}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})} (and thus perpendicular to 𝒗s𝒗s\smash{{\boldsymbol{v}}_{s}-{\boldsymbol{v}}^{\prime}_{s^{\prime}}}), one finds that 𝑮ss(𝒑,𝒑)=0\smash{{\boldsymbol{G}}_{ss^{\prime}}^{\perp}({\boldsymbol{p}},{\boldsymbol{p}}^{\prime})=0}. Combined with (589) and (590), this yields

lnFs(𝒑)𝒑lnFs(𝒑)𝒑=λss(𝒗s,𝒗s)(𝒗s𝒗s).\displaystyle\frac{\partial\ln F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}-\frac{\partial\ln F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial{\boldsymbol{p}}^{\prime}}=\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,({\boldsymbol{v}}_{s}-{\boldsymbol{v}}^{\prime}_{s^{\prime}}). (592)

Also, by swapping 𝒑𝒑\smash{{\boldsymbol{p}}\leftrightarrow{\boldsymbol{p}}^{\prime}} and ss\smash{s\leftrightarrow s^{\prime}}, one finds that

λss(𝒗s,𝒗s)=λss(𝒗s,𝒗s).\displaystyle\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})=\lambda_{s^{\prime}s}({\boldsymbol{v}}^{\prime}_{s^{\prime}},{\boldsymbol{v}}_{s}). (593)

Equation (592) yields, in particular, that464646The idea of this argument was brought to author’s attention by G. W. Hammett and is taken from (Landreman, 2017), where it is applied to single-species plasmas with a specific s\smash{\mathcal{H}_{s}}.

lnFs(𝒑)p2lnFs(𝒑)p2=λss(𝒗s,𝒗s)(vs,2vs,2),\displaystyle\frac{\partial\ln F_{s}({\boldsymbol{p}})}{\partial p_{2}}-\frac{\partial\ln F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial p_{2}^{\prime}}=\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,(v_{s,2}-v^{\prime}_{s^{\prime},2}), (594a)
lnFs(𝒑)p3lnFs(𝒑)p3=λss(𝒗s,𝒗s)(vs,3vs,3),\displaystyle\frac{\partial\ln F_{s}({\boldsymbol{p}})}{\partial p_{3}}-\frac{\partial\ln F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial p_{3}^{\prime}}=\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})\,(v_{s,3}-v^{\prime}_{s^{\prime},3}), (594b)

where we have assumed some coordinate axes in the momentum and velocity space labeled (1,2,3,)\smash{(1,2,3,\ldots)}. Then,

2lnFs(𝒑)p2vs,1=λss(𝒗s,𝒗s)vs,1(vs,2vs,2),\displaystyle\frac{\partial^{2}\ln F_{s}({\boldsymbol{p}})}{\partial p_{2}\partial v_{s,1}}=\frac{\partial\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})}{\partial v_{s,1}}\,(v_{s,2}-v^{\prime}_{s^{\prime},2}), (595a)
2lnFs(𝒑)p3vs,1=λss(𝒗s,𝒗s)vs,1(vs,3vs,3),\displaystyle\frac{\partial^{2}\ln F_{s}({\boldsymbol{p}})}{\partial p_{3}\partial v_{s,1}}=\frac{\partial\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})}{\partial v_{s,1}}\,(v_{s,3}-v^{\prime}_{s^{\prime},3}), (595b)

where the derivative with respect to vs,1\smash{v_{s,1}} is taken at fixed vs,i1\smash{v_{s,i\neq 1}} and at fixed 𝒗s\smash{{\boldsymbol{v}}^{\prime}_{s^{\prime}}}. Due to (592), λss(𝒗s,𝒗s)\smash{\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})} is continuous for all Fs\smash{F_{s}} and Fs\smash{F_{s^{\prime}}}. (Here we consider only physical distributions, which are always differentiable.) Then, (595) leads to

1vs,2vs,22lnFs(𝒑)p2vs,1=1vs,3vs,32lnFs(𝒑)p3vs,1.\displaystyle\frac{1}{v_{s,2}-v^{\prime}_{s^{\prime},2}}\frac{\partial^{2}\ln F_{s}({\boldsymbol{p}})}{\partial p_{2}\partial v_{s,1}}=\frac{1}{v_{s,3}-v^{\prime}_{s^{\prime},3}}\frac{\partial^{2}\ln F_{s}({\boldsymbol{p}})}{\partial p_{3}\partial v_{s,1}}. (596)

By differentiating this with respect to vs,2\smash{v^{\prime}_{s^{\prime},2}}, one obtains

2lnFs(𝒑)p2vs,1=0,\displaystyle\frac{\partial^{2}\ln F_{s}({\boldsymbol{p}})}{\partial p_{2}\partial v_{s,1}}=0, (597)

whence (595a) yields

λss(𝒗s,𝒗s)vs,1=0.\displaystyle\frac{\partial\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})}{\partial v_{s,1}}=0. (598)

By repeating this argument for other axes and for 𝒗\smash{{\boldsymbol{v}}^{\prime}} instead of 𝒗\smash{{\boldsymbol{v}}}, one can also extend (598) to

λss(𝒗s,𝒗s)𝒗=0,λss(𝒗s,𝒗s)𝒗=0.\displaystyle\frac{\partial\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})}{\partial{\boldsymbol{v}}}=0,\qquad\frac{\partial\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})}{\partial{\boldsymbol{v}}^{\prime}}=0. (599)

Hence, λss(𝒗s,𝒗s)\smash{\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})} is actually independent of the velocities; i.e. λss(𝒗s,𝒗s)=λss\smash{\lambda_{ss^{\prime}}({\boldsymbol{v}}_{s},{\boldsymbol{v}}^{\prime}_{s^{\prime}})=\lambda_{ss^{\prime}}}. Using this along with (593), one also finds that

λss=λss.\displaystyle\lambda_{ss^{\prime}}=\lambda_{s^{\prime}s}. (600)

Let us rewrite (592) as follows:

lnFs(𝒑)𝒑λss𝒗s=lnFs(𝒑)𝒑λss𝒗s.\displaystyle\frac{\partial\ln F_{s}({\boldsymbol{p}})}{\partial{\boldsymbol{p}}}-\lambda_{ss^{\prime}}{\boldsymbol{v}}_{s}=\frac{\partial\ln F_{s^{\prime}}({\boldsymbol{p}}^{\prime})}{\partial{\boldsymbol{p}}^{\prime}}-\lambda_{s^{\prime}s}{\boldsymbol{v}}^{\prime}_{s^{\prime}}. (601)

Here, the left-hand side is independent of 𝒑\smash{{\boldsymbol{p}}^{\prime}} and the right-hand side is independent of 𝒑\smash{{\boldsymbol{p}}}, so both must be equal to some vector

𝝁ss=𝝁ss{\boldsymbol{\mu}}_{ss^{\prime}}={\boldsymbol{\mu}}_{s^{\prime}s} (602)

that is independent of both 𝒑\smash{{\boldsymbol{p}}} and 𝒑\smash{{\boldsymbol{p}}^{\prime}}. Because 𝒗s=𝒑s\smash{{\boldsymbol{v}}_{s}=\partial_{{\boldsymbol{p}}}\mathcal{H}_{s}}, this is equivalent to

lnFs(𝒑)λsss(𝒑)=𝝁ss𝒑+ηss\displaystyle\ln F_{s}({\boldsymbol{p}})-\lambda_{ss^{\prime}}\mathcal{H}_{s}({\boldsymbol{p}})={\boldsymbol{\mu}}_{ss^{\prime}}\cdot{\boldsymbol{p}}+\eta_{ss^{\prime}} (603a)
(and similarly for 𝒑\smash{{\boldsymbol{p}}^{\prime}}), where the integration constant ηss\smash{\eta_{ss^{\prime}}} is independent of both 𝒑\smash{{\boldsymbol{p}}} and 𝒑\smash{{\boldsymbol{p}}^{\prime}}. This is supposed to hold for any s\smash{s^{\prime}}, so one can also write
lnFs(𝒑)λss′′s(𝒑)=𝝁ss′′𝒑+ηss′′,\displaystyle\ln F_{s}({\boldsymbol{p}})-\lambda_{ss^{\prime\prime}}\mathcal{H}_{s}({\boldsymbol{p}})={\boldsymbol{\mu}}_{ss^{\prime\prime}}\cdot{\boldsymbol{p}}+\eta_{ss^{\prime\prime}}, (603b)

where s′′\smash{s^{\prime\prime}} is any other species index. Subtracting equations (603) from each other gives

(λssλss′′)s(𝒑)=(𝝁ss𝝁ss′′)𝒑+ηssηss′′.\displaystyle(\lambda_{ss^{\prime}}-\lambda_{ss^{\prime\prime}})\mathcal{H}_{s}({\boldsymbol{p}})=({\boldsymbol{\mu}}_{ss^{\prime}}-{\boldsymbol{\mu}}_{ss^{\prime\prime}})\cdot{\boldsymbol{p}}+\eta_{ss^{\prime}}-\eta_{ss^{\prime\prime}}. (604)

By differentiating this with respect to 𝒑\smash{{\boldsymbol{p}}}, one finds

(λssλss′′)𝒗s=𝝁ss𝝁ss′′.\displaystyle(\lambda_{ss^{\prime}}-\lambda_{ss^{\prime\prime}}){\boldsymbol{v}}_{s}={\boldsymbol{\mu}}_{ss^{\prime}}-{\boldsymbol{\mu}}_{ss^{\prime\prime}}. (605)

By differentiating this further with respect to 𝒗s\smash{{\boldsymbol{v}}_{s}}, one obtains λss=λss′′\smash{\lambda_{ss^{\prime}}=\lambda_{ss^{\prime\prime}}}. Then, (605) yields 𝝁ss=𝝁ss′′\smash{{\boldsymbol{\mu}}_{ss^{\prime}}={\boldsymbol{\mu}}_{ss^{\prime\prime}}}, and (604) yields ηss=ηss′′\smash{\eta_{ss^{\prime}}=\eta_{ss^{\prime\prime}}}. In other words, the functions λss\smash{\lambda_{ss^{\prime}}}, 𝝁ss\smash{{\boldsymbol{\mu}}_{ss^{\prime}}}, and ηss\smash{\eta_{ss^{\prime}}} are independent of their second index and thus can as well be written as

λss=λs,𝝁ss=𝝁s,ηss=ηs.\displaystyle\lambda_{ss^{\prime}}=\lambda_{s},\qquad{\boldsymbol{\mu}}_{ss^{\prime}}={\boldsymbol{\mu}}_{s},\qquad\eta_{ss^{\prime}}=\eta_{s}. (606)

But then, (600) and (602) also yield λs=λsλ\smash{\lambda_{s}=\lambda_{s^{\prime}}\equiv\lambda} and 𝝁s=𝝁s𝝁\smash{{\boldsymbol{\mu}}_{s}={\boldsymbol{\mu}}_{s^{\prime}}\equiv{\boldsymbol{\mu}}}. Therefore, (603) can be written as

Fs(𝒑)=consts×exp(λs(𝒑)+𝝁𝒑),\displaystyle F_{s}({\boldsymbol{p}})=\text{const}_{s}\times\exp(\lambda\mathcal{H}_{s}({\boldsymbol{p}})+{\boldsymbol{\mu}}\cdot{\boldsymbol{p}}), (607)

which is the Boltzmann–Gibbs distribution (section 8.1). This proves that a plasma that conserves its entropy density necessarily has the Boltzmann–Gibbs distribution.

Appendix F Total momentum and energy

Here, we show that the total momentum and energy in the OC–wave representation equals the total momentum and energy in the particle–field representation.

F.1 Nonrelativistic electrostatic interactions

F.1.1 Momentum

Assuming the notation 𝒫lsd𝒑plf¯s\smash{\mathcal{P}_{l}\doteq\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}\overline{f}_{s}} and using (431) for 𝚯s\smash{{\boldsymbol{\Uptheta}}_{s}}, one can represent the OC momentum density as follows:

sd𝒑plFs\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}F_{s} =𝒫l+12sd𝒑plpi(Θs,ijf¯spj)\displaystyle=\mathcal{P}_{l}+\frac{1}{2}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}\,\frac{\partial}{\partial p_{i}}\left(\Uptheta_{s,ij}\,\frac{\partial\overline{f}_{s}}{\partial p_{j}}\right)
𝒫l12sd𝒑Θs,ljFspj\displaystyle\approx\mathcal{P}_{l}-\frac{1}{2}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\Uptheta_{s,lj}\,\frac{\partial F_{s}}{\partial p_{j}}
=𝒫ld𝒌h(𝒌)ϑses2d𝒑klw(𝒌)𝒌𝒗s+ϑ𝒌Fs𝒑|ϑ=0\displaystyle=\mathcal{P}_{l}-\int\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\left.\frac{\partial}{\partial\vartheta}\sum_{s}e_{s}^{2}\fint\mathrm{d}{\boldsymbol{p}}\,\frac{k_{l}}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}\right|_{\vartheta=0}
=𝒫ld𝒌klh(𝒌)ϑ(k2(ϵH(w(𝒌)+ϑ,𝒌)1)4\upi)|ϑ=0\displaystyle=\mathcal{P}_{l}-\int\mathrm{d}{\boldsymbol{k}}\,k_{l}h({\boldsymbol{k}})\left.\frac{\partial}{\partial\vartheta}\left(\frac{k^{2}(\epsilon_{\parallel\text{H}}(w({\boldsymbol{k}})+\vartheta,{\boldsymbol{k}})-1)}{4\upi}\right)\right|_{\vartheta=0}
=𝒫ld𝒌klJ,\displaystyle=\mathcal{P}_{l}-\int\mathrm{d}{\boldsymbol{k}}\,k_{l}J, (608)

where we substituted (430). This leads to (432).

F.1.2 Energy

Assuming the notation 𝒦sd𝒑H0sf¯s\smash{\mathcal{K}\doteq\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}\overline{f}_{s}} and using (431) for 𝚯s\smash{{\boldsymbol{\Uptheta}}_{s}}, one can represent the OC energy density as follows:

sd𝒑H0sFs\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s} =𝒦+12sd𝒑p22mspi(Θs,ijf¯spj)\displaystyle=\mathcal{K}+\frac{1}{2}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\frac{p^{2}}{2m_{s}}\,\frac{\partial}{\partial p_{i}}\left(\Uptheta_{s,ij}\,\frac{\partial\overline{f}_{s}}{\partial p_{j}}\right)
𝒦12sd𝒑vsiΘs,ijFspj\displaystyle\approx\mathcal{K}-\frac{1}{2}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,v_{s}^{i}\Uptheta_{s,ij}\,\frac{\partial F_{s}}{\partial p_{j}}
=𝒦d𝒌h(𝒌)ϑses2d𝒑𝒌𝒗w(𝒌)𝒌𝒗s+ϑ𝒌Fs𝒑|ϑ=0.\displaystyle=\mathcal{K}-\int\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\,\frac{\partial}{\partial\vartheta}\sum_{s}e_{s}^{2}\fint\mathrm{d}{\boldsymbol{p}}\left.\frac{{\boldsymbol{k}}\cdot{{\boldsymbol{v}}}}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}\right|_{\vartheta=0}. (609)

Notice that

ϑ𝒌𝒗w(𝒌)𝒌𝒗s+ϑ\displaystyle\frac{\partial}{\partial\vartheta}\frac{{\boldsymbol{k}}\cdot{{\boldsymbol{v}}}}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta} =ϑ(1+w(𝒌)+ϑw(𝒌)𝒌𝒗s+ϑ)\displaystyle=\frac{\partial}{\partial\vartheta}\left(-1+\frac{w({\boldsymbol{k}})+\vartheta}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\right)
=1w(𝒌)𝒌𝒗s+ϑ+w(𝒌)ϑ1w(𝒌)𝒌𝒗s+ϑ.\displaystyle=\frac{1}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}+w({\boldsymbol{k}})\,\frac{\partial}{\partial\vartheta}\frac{1}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}. (610)

Then,

sd𝒑H0sFs=\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s}= 𝒦d𝒌w(𝒌)h(𝒌)ϑses2d𝒑𝒌w(𝒌)𝒌𝒗s+ϑFs𝒑|ϑ=0\displaystyle\,\mathcal{K}-\int\mathrm{d}{\boldsymbol{k}}\,w({\boldsymbol{k}})h({\boldsymbol{k}})\,\frac{\partial}{\partial\vartheta}\sum_{s}e_{s}^{2}\fint\mathrm{d}{\boldsymbol{p}}\left.\frac{{\boldsymbol{k}}}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\,\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}\right|_{\vartheta=0}
d𝒌h(𝒌)ses2d𝒑𝒌w(𝒌)𝒌𝒗sFs𝒑\displaystyle-\int\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\sum_{s}e_{s}^{2}\fint\mathrm{d}{\boldsymbol{p}}\,\frac{{\boldsymbol{k}}}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}\,\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}
=\displaystyle= 𝒦d𝒌w(𝒌)h(𝒌)ϑ(k2(ϵH(w(𝒌)+ϑ,𝒌)1)4\upi)|ϑ=0\displaystyle\,\mathcal{K}-\int\mathrm{d}{\boldsymbol{k}}\,w({\boldsymbol{k}})h({\boldsymbol{k}})\left.\frac{\partial}{\partial\vartheta}\left(\frac{k^{2}(\epsilon_{\parallel\text{H}}(w({\boldsymbol{k}})+\vartheta,{\boldsymbol{k}})-1)}{4\upi}\right)\right|_{\vartheta=0}
d𝒌h(𝒌)k2(ϵH(w(𝒌),𝒌)1)4\upi.\displaystyle-\int\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\,\frac{k^{2}(\epsilon_{\parallel\text{H}}(w({\boldsymbol{k}}),{\boldsymbol{k}})-1)}{4\upi}. (611)

Using (428) and (430), one obtains that the sum of the OC and wave energy is given by

sd𝒑H0sFs+d𝒌wJ\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,wJ =𝒦+d𝒌h(𝒌)k24\upi\displaystyle=\mathcal{K}+\int\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\,\frac{k^{2}}{4\upi}
=𝒦+σk¯σ2|φ˘σ|216\upi\displaystyle=\mathcal{K}+\sum_{\sigma}\frac{\overline{k}_{\sigma}^{2}|{\breve{\varphi}}_{\sigma}|^{2}}{16\upi}
=sd𝒑(p22ms+esφ¯s)f¯s+18\upi𝑬~𝑬~¯,\displaystyle=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\left(\frac{p^{2}}{2m_{s}}+e_{s}\overline{\varphi}_{s}\right)\overline{f}_{s}+\frac{1}{8\upi}\,\overline{\smash{\widetilde{{\boldsymbol{E}}}}^{\dagger}\widetilde{{\boldsymbol{E}}}}, (612)

where we also substituted (355).

F.2 Relativistic electromagnetic interactions

F.2.1 Momentum

Let us assume the notation 𝓟sd𝒑𝒑f¯s\smash{{\boldsymbol{\mathcal{P}}}\doteq\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{p}}\overline{f}_{s}} and

𝝌(ω,𝒌)s4\upies2ω2d𝒑𝒗s𝒗sω𝒌𝒗s𝒌Fs𝒑=ϵ(ω,𝒌)𝟏+𝖜pω2.\displaystyle{\boldsymbol{\chi}}(\omega,{\boldsymbol{k}})\doteq\sum_{s}\frac{4\upi e_{s}^{2}}{\omega^{2}}\fint\mathrm{d}{\boldsymbol{p}}\,\frac{{\boldsymbol{v}}_{s}{\boldsymbol{v}}_{s}^{\dagger}}{\omega-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}={\boldsymbol{\epsilon}}(\omega,{\boldsymbol{k}})-{\boldsymbol{1}}+\frac{{\boldsymbol{\mathfrak{w}}}_{p}}{\omega^{2}}. (613)

Then, using (466) for 𝚯s\smash{{\boldsymbol{\Uptheta}}_{s}}, one can represent the OC momentum density as follows:

sd𝒑plFs\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,p_{l}F_{s} 𝒫l12sd𝒑Θs,ljFspj\displaystyle\approx\mathcal{P}_{l}-\frac{1}{2}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\Uptheta_{s,lj}\,\frac{\partial F_{s}}{\partial p_{j}}
=𝒫ld𝒌klh(𝒌)ϑses2d𝒑1w2(𝒌)(𝜼𝒗s𝒗s𝜼)w(𝒌)𝒌𝒗s+ϑ𝒌Fspj|ϑ=0\displaystyle=\mathcal{P}_{l}-\int\mathrm{d}{\boldsymbol{k}}\,k_{l}h({\boldsymbol{k}})\left.\frac{\partial}{\partial\vartheta}\sum_{s}e_{s}^{2}\fint\mathrm{d}{\boldsymbol{p}}\,\frac{1}{w^{2}({\boldsymbol{k}})}\frac{({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{v}}_{s}{\boldsymbol{v}}_{s}^{\dagger}{\boldsymbol{\eta}})}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}}{\partial p_{j}}\right|_{\vartheta=0}
=𝒫ld𝒌klh(𝒌)4\upiw2(𝒌)𝜼ω(ω2𝝌(ω,𝒌))𝜼|ω=w(𝒌)\displaystyle=\mathcal{P}_{l}-\int\mathrm{d}{\boldsymbol{k}}\,\frac{k_{l}h({\boldsymbol{k}})}{4\upi w^{2}({\boldsymbol{k}})}\left.{\boldsymbol{\eta}}^{\dagger}\frac{\partial}{\partial\omega}\left(\omega^{2}{\boldsymbol{\chi}}(\omega,{\boldsymbol{k}})\right){\boldsymbol{\eta}}\right|_{\omega=w({\boldsymbol{k}})}
=𝒫l+d𝒌klh(𝒌)2\upiw(𝒌)d𝒌klh(𝒌)4\upiw2(𝒌)𝜼ω(ω2ϵ(ω,𝒌))𝜼|ω=w(𝒌)\displaystyle=\mathcal{P}_{l}+\int\mathrm{d}{\boldsymbol{k}}\,\frac{k_{l}h({\boldsymbol{k}})}{2\upi w({\boldsymbol{k}})}-\int\mathrm{d}{\boldsymbol{k}}\,\frac{k_{l}h({\boldsymbol{k}})}{4\upi w^{2}({\boldsymbol{k}})}\left.{\boldsymbol{\eta}}^{\dagger}\frac{\partial}{\partial\omega}\left(\omega^{2}{\boldsymbol{\epsilon}}(\omega,{\boldsymbol{k}})\right){\boldsymbol{\eta}}\right|_{\omega=w({\boldsymbol{k}})}
=𝒫l+d𝒌kl4\upiw(𝒌)(h(𝒌)+h(𝒌))d𝒌klJ,\displaystyle=\mathcal{P}_{l}+\int\mathrm{d}{\boldsymbol{k}}\,\frac{k_{l}}{4\upi w({\boldsymbol{k}})}\,(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))-\int\mathrm{d}{\boldsymbol{k}}\,k_{l}J, (614)

where we substituted (465) and used (314). Next, let us rewrite (468) as

𝓟=𝓟(kin)+(𝑨~/c)sesd𝒑fs¯=𝓟(kin)+14\upic𝑨~(𝑬~)¯,\displaystyle{\boldsymbol{\mathcal{P}}}={\boldsymbol{\mathcal{P}}}^{(\text{kin})}+\overline{(\widetilde{{\boldsymbol{A}}}/c)\sum_{s}e_{s}\textstyle\int\mathrm{d}{\boldsymbol{p}}\,f_{s}}={\boldsymbol{\mathcal{P}}}^{(\text{kin})}+\frac{1}{4\upi c}\,\overline{\widetilde{{\boldsymbol{A}}}(\nabla\cdot\widetilde{{\boldsymbol{E}}})}, (615)

where the last equality is due to Gauss’s law. This gives

𝒫l𝒫l(kin)=14\upi(iω^1E~l)(jE~j)¯=i4\upi(ω^1E~l)(jE~j)¯.\displaystyle\mathcal{P}_{l}-\mathcal{P}^{(\text{kin})}_{l}=\frac{1}{4\upi}\,\overline{(-\mathrm{i}\widehat{\omega}^{-1}\widetilde{E}_{l})(\partial_{j}\widetilde{E}^{j})}=-\frac{\mathrm{i}}{4\upi}\,\overline{(\widehat{\omega}^{-1}\widetilde{E}_{l})(\partial_{j}\widetilde{E}^{j})^{*}}. (616)

Then, using (53) and also (354) for 𝗨\smash{{\boldsymbol{\mathsf{U}}}}, one obtains

𝒫l𝒫l(kin)\displaystyle\mathcal{P}_{l}-\mathcal{P}^{(\text{kin})}_{l} i4\upidωd𝒌ω1𝖴l(ω,𝒌)j(ikj)\displaystyle\approx-\frac{\mathrm{i}}{4\upi}\,\int\mathrm{d}\omega\,\mathrm{d}{\boldsymbol{k}}\,\omega^{-1}\mathsf{U}_{l}{}^{j}(\omega,{\boldsymbol{k}})(\mathrm{i}k_{j})^{*}
d𝒌𝒌𝜼4\upiw(𝒌)ηl(h(𝒌)+h(𝒌)),\displaystyle\approx-\int\mathrm{d}{\boldsymbol{k}}\,\frac{{\boldsymbol{k}}\cdot{\boldsymbol{\eta}}^{*}}{4\upi w({\boldsymbol{k}})}\,\eta_{l}(h({\boldsymbol{k}})+h(-{\boldsymbol{k}})), (617)

and thus (614) can be written as follows:

sd𝒑𝒑Fs\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{p}}F_{s} +d𝒌𝒌J𝓟+d𝒌𝒌4\upiw(𝒌)(h(𝒌)+h(𝒌))\displaystyle+\int\mathrm{d}{\boldsymbol{k}}\,{\boldsymbol{k}}J\approx{\boldsymbol{\mathcal{P}}}+\int\mathrm{d}{\boldsymbol{k}}\,\frac{{\boldsymbol{k}}}{4\upi w({\boldsymbol{k}})}\,(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))
𝓟(kin)+d𝒌14\upiw(𝒌)(𝒌𝜼(𝒌𝜼))(h(𝒌)+h(𝒌))\displaystyle\approx{\boldsymbol{\mathcal{P}}}^{(\text{kin})}+\int\mathrm{d}{\boldsymbol{k}}\,\frac{1}{4\upi w({\boldsymbol{k}})}\,({\boldsymbol{k}}-{\boldsymbol{\eta}}({\boldsymbol{k}}\cdot{\boldsymbol{\eta}}^{*}))\,(h({\boldsymbol{k}})+h(-{\boldsymbol{k}}))
=𝓟(kin)+red𝒌h(𝒌)2\upiw(𝒌)𝜼×(𝒌×𝜼),\displaystyle={\boldsymbol{\mathcal{P}}}^{(\text{kin})}+\operatorname{re}\int\mathrm{d}{\boldsymbol{k}}\,\frac{h({\boldsymbol{k}})}{2\upi w({\boldsymbol{k}})}\,{\boldsymbol{\eta}}^{*}\times({\boldsymbol{k}}\times{\boldsymbol{\eta}}), (618)

where we used (298) and (314) again. For an eikonal wave (471), which has h(𝒌)=δ(𝒌𝒌¯)|𝑬˘|2/4\smash{h({\boldsymbol{k}})=\delta({\boldsymbol{k}}-\overline{{\boldsymbol{k}}})|{\breve{{\boldsymbol{E}}}}|^{2}/4} (section 7.4.1), this gives

red𝒌h(𝒌)2\upiw(𝒌)𝜼×(𝒌×𝜼)=18\upicre(𝑬˘×(c𝒌¯ω¯×𝑬˘))=𝑬~×𝑩~¯4\upic.\displaystyle\operatorname{re}\int\mathrm{d}{\boldsymbol{k}}\,\frac{h({\boldsymbol{k}})}{2\upi w({\boldsymbol{k}})}\,{\boldsymbol{\eta}}^{*}\times({\boldsymbol{k}}\times{\boldsymbol{\eta}})=\frac{1}{8\upi c}\,\operatorname{re}\left(\smash{{\breve{{\boldsymbol{E}}}}}^{*}\times\left(\frac{c\overline{{\boldsymbol{k}}}}{\overline{\omega}}\times{\breve{{\boldsymbol{E}}}}\right)\right)=\frac{\overline{\widetilde{{\boldsymbol{E}}}\times\widetilde{{\boldsymbol{B}}}}}{4\upi c}. (619)

In case of a broadband spectrum, the same equality applies as well, because contributions of the individual eikonal waves to both left-hand side and the right-hand side are additive. (Alternatively, one can invoke (53) again.) This leads to (467).

F.2.2 Energy

Assuming the notation 𝒦sd𝒑H0sf¯s\smash{\mathcal{K}\doteq\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,H_{0s}\overline{f}_{s}} and using (466) for 𝚯s\smash{{\boldsymbol{\Uptheta}}_{s}}, one can represent the OC energy density as follows:

s\displaystyle\sum_{s}\int d𝒑H0sFs𝒦12sd𝒑vsiΘs,ijFspj\displaystyle\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s}\approx\mathcal{K}-\frac{1}{2}\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,v_{s}^{i}\Uptheta_{s,ij}\,\frac{\partial F_{s}}{\partial p_{j}}
𝒦d𝒌h(𝒌)ϑses2d𝒑(𝒌𝒗s)w2(𝒌)(𝜼𝒗s𝒗s𝜼)w(𝒌)𝒌𝒗s+ϑ𝒌Fs𝒑|ϑ=0.\displaystyle\approx\mathcal{K}-\int\mathrm{d}{\boldsymbol{k}}\,h({\boldsymbol{k}})\,\frac{\partial}{\partial\vartheta}\sum_{s}e_{s}^{2}\fint\mathrm{d}{\boldsymbol{p}}\left.\frac{({\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s})}{w^{2}({\boldsymbol{k}})}\frac{({\boldsymbol{\eta}}^{\dagger}{\boldsymbol{v}}_{s}{\boldsymbol{v}}_{s}^{\dagger}{\boldsymbol{\eta}})}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}\right|_{\vartheta=0}.

Using (610) and (613) for 𝝌\smash{{\boldsymbol{\chi}}}, one further obtains

s\displaystyle\sum_{s}\int d𝒑H0sFs\displaystyle\mathrm{d}{\boldsymbol{p}}\,H_{0s}F_{s}
=\displaystyle= 𝒦d𝒌h(𝒌)4\upiw(𝒌)𝜼ϑ(s4\upies2d𝒑𝒗s𝒗sw(𝒌)𝒌𝒗s+ϑ𝒌Fs𝒑)𝜼|ϑ=0\displaystyle\,\mathcal{K}-\int\mathrm{d}{\boldsymbol{k}}\,\frac{h({\boldsymbol{k}})}{4\upi w({\boldsymbol{k}})}\left.{\boldsymbol{\eta}}^{\dagger}\,\frac{\partial}{\partial\vartheta}\bigg{(}\sum_{s}4\upi e_{s}^{2}\fint\mathrm{d}{\boldsymbol{p}}\,\frac{{\boldsymbol{v}}_{s}{\boldsymbol{v}}_{s}^{\dagger}}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}+\vartheta}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}\bigg{)}{\boldsymbol{\eta}}\right|_{\vartheta=0}
d𝒌h(𝒌)4\upi𝜼(s4\upies2w2(𝒌)d𝒑𝒗s𝒗sw(𝒌)𝒌𝒗s𝒌Fs𝒑)𝜼\displaystyle-\int\mathrm{d}{\boldsymbol{k}}\,\frac{h({\boldsymbol{k}})}{4\upi}\,{\boldsymbol{\eta}}^{\dagger}\,\bigg{(}\sum_{s}\frac{4\upi e_{s}^{2}}{w^{2}({\boldsymbol{k}})}\fint\mathrm{d}{\boldsymbol{p}}\,\frac{{\boldsymbol{v}}_{s}{\boldsymbol{v}}_{s}^{\dagger}}{w({\boldsymbol{k}})-{\boldsymbol{k}}\cdot{\boldsymbol{v}}_{s}}\,{\boldsymbol{k}}\cdot\frac{\partial F_{s}}{\partial{\boldsymbol{p}}}\bigg{)}{\boldsymbol{\eta}}
=\displaystyle= 𝒦d𝒌h(𝒌)4\upiw(𝒌)𝜼(ω2𝝌(ω,𝒌))ω𝜼|ω=w(𝒌)d𝒌h(𝒌)4\upi𝜼𝝌𝜼\displaystyle\,\mathcal{K}-\int\mathrm{d}{\boldsymbol{k}}\,\frac{h({\boldsymbol{k}})}{4\upi w({\boldsymbol{k}})}\left.{\boldsymbol{\eta}}^{\dagger}\frac{\partial(\omega^{2}{\boldsymbol{\chi}}(\omega,{\boldsymbol{k}}))}{\partial\omega}{\boldsymbol{\eta}}\right|_{\omega=w({\boldsymbol{k}})}-\int\mathrm{d}{\boldsymbol{k}}\,\frac{h({\boldsymbol{k}})}{4\upi}\,{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\chi}}{\boldsymbol{\eta}}
=\displaystyle= 𝒦d𝒌h(𝒌)4\upiw(𝒌)𝜼(ω2ϵ(ω,𝒌))ω𝜼|ω=w(𝒌)+d𝒌h(𝒌)2\upi\displaystyle\,\mathcal{K}-\int\mathrm{d}{\boldsymbol{k}}\,\frac{h({\boldsymbol{k}})}{4\upi w({\boldsymbol{k}})}\left.{\boldsymbol{\eta}}^{\dagger}\frac{\partial(\omega^{2}{\boldsymbol{\epsilon}}(\omega,{\boldsymbol{k}}))}{\partial\omega}{\boldsymbol{\eta}}\right|_{\omega=w({\boldsymbol{k}})}+\int\mathrm{d}{\boldsymbol{k}}\,\frac{h({\boldsymbol{k}})}{2\upi}
d𝒌h(𝒌)4\upi𝜼(ϵ(w(𝒌),𝒌)𝟏+𝖜pw2(𝒌))𝜼\displaystyle-\int\mathrm{d}{\boldsymbol{k}}\,\frac{h({\boldsymbol{k}})}{4\upi}\,{\boldsymbol{\eta}}^{\dagger}\left({\boldsymbol{\epsilon}}(w({\boldsymbol{k}}),{\boldsymbol{k}})-{\boldsymbol{1}}+\frac{{\boldsymbol{\mathfrak{w}}}_{p}}{w^{2}({\boldsymbol{k}})}\right){\boldsymbol{\eta}}
=\displaystyle= 𝒦d𝒌wJ+d𝒌3h(𝒌)4\upid𝒌h(𝒌)4\upi𝜼ϵ(w(𝒌),𝒌)𝜼d𝒌h(𝒌)4\upiw2(𝒌)𝜼𝖜p𝜼.\displaystyle\,\mathcal{K}-\int\mathrm{d}{\boldsymbol{k}}\,wJ+\int\mathrm{d}{\boldsymbol{k}}\,\frac{3h({\boldsymbol{k}})}{4\upi}-\int\mathrm{d}{\boldsymbol{k}}\,\frac{h({\boldsymbol{k}})}{4\upi}\,{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\epsilon}}(w({\boldsymbol{k}}),{\boldsymbol{k}}){\boldsymbol{\eta}}-\int\mathrm{d}{\boldsymbol{k}}\,\frac{h({\boldsymbol{k}})}{4\upi w^{2}({\boldsymbol{k}})}\,{\boldsymbol{\eta}}^{\dagger}{\boldsymbol{\mathfrak{w}}}_{p}{\boldsymbol{\eta}}.

Using (464) and proceeding as in section F.2.1, one can also cast this as follows:

sd𝒑H0sFs+d𝒌wJ=𝒦+18\upi(3𝑬~𝑬~¯𝑩~𝑩~¯)18\upic2𝑨~𝖜p𝑨~¯.\displaystyle\sum_{s}\int\mathrm{d}{\boldsymbol{p}}H_{0s}F_{s}+\int\mathrm{d}{\boldsymbol{k}}\,wJ=\mathcal{K}+\frac{1}{8\upi}\,\big{(}3\overline{\smash{\widetilde{{\boldsymbol{E}}}}^{\dagger}\widetilde{{\boldsymbol{E}}}}-\overline{\smash{\widetilde{{\boldsymbol{B}}}}^{\dagger}\widetilde{{\boldsymbol{B}}}}\big{)}-\frac{1}{8\upi c^{2}}\,\overline{\smash{\widetilde{{\boldsymbol{A}}}}^{\dagger}{\boldsymbol{\mathfrak{w}}}_{p}\widetilde{{\boldsymbol{A}}}}. (620)

Now, notice that

𝒦\displaystyle\mathcal{K} =sd𝒑H0s(𝒑)fs(kin)(𝒑es𝑨~/c)¯\displaystyle=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\overline{H_{0s}({\boldsymbol{p}})f_{s}^{(\text{kin})}({\boldsymbol{p}}-e_{s}\widetilde{{\boldsymbol{A}}}/c)}
=sd𝒑H0s(𝒑+es𝑨~/c)fs(kin)(𝒑)¯\displaystyle=\sum_{s}\int\mathrm{d}{\boldsymbol{p}}\,\overline{H_{0s}({\boldsymbol{p}}+e_{s}\widetilde{{\boldsymbol{A}}}/c)f_{s}^{(\text{kin})}({\boldsymbol{p}})}
𝒦(kin)+(𝑨~/c)sesd𝒑𝒗sfs(kin)¯+ses22c2d𝒑(𝑨~𝝁s1𝑨~)fs(kin)¯\displaystyle\approx\mathcal{K}^{(\text{kin})}+\overline{(\widetilde{{\boldsymbol{A}}}/c)\cdot\sum_{s}e_{s}\textstyle\int\mathrm{d}{\boldsymbol{p}}\,{\boldsymbol{v}}_{s}f_{s}^{(\text{kin})}}+\sum_{s}\frac{e_{s}^{2}}{2c^{2}}\int\mathrm{d}{\boldsymbol{p}}\,\overline{(\widetilde{{\boldsymbol{A}}}{\boldsymbol{\mu}}_{s}^{-1}\widetilde{{\boldsymbol{A}}})\,f_{s}^{(\text{kin})}}
𝒦(kin)+1c𝑨~𝒋~¯+18\upic2𝑨~𝖜p𝑨~¯,\displaystyle\approx\mathcal{K}^{(\text{kin})}+\frac{1}{c}\,\overline{\widetilde{{\boldsymbol{A}}}\cdot\widetilde{{\boldsymbol{j}}}}+\frac{1}{8\upi c^{2}}\,\overline{\smash{\widetilde{{\boldsymbol{A}}}}^{\dagger}{\boldsymbol{\mathfrak{w}}}_{p}\widetilde{{\boldsymbol{A}}}}, (621)

where 𝒋~\smash{\widetilde{{\boldsymbol{j}}}} is the oscillating-current density. From Ampere’s law,

1c𝑨~𝒋~¯\displaystyle\frac{1}{c}\,\overline{\widetilde{{\boldsymbol{A}}}\cdot\widetilde{{\boldsymbol{j}}}} =(iω^1𝑬~)4\upi(ic𝒌^×𝑩~+iω^𝑬~)¯\displaystyle=\overline{\frac{(-\mathrm{i}\widehat{\omega}^{-1}\widetilde{{\boldsymbol{E}}})^{\dagger}}{4\upi}\big{(}\mathrm{i}c\widehat{\boldsymbol{k}}\times\widetilde{{\boldsymbol{B}}}+\mathrm{i}\widehat{\omega}\widetilde{{\boldsymbol{E}}}\big{)}}
𝑬~𝑬~4\upi¯𝑬~4\upi(c𝒌^ω^×𝑩~)¯\displaystyle\approx-\overline{\frac{\smash{\widetilde{{\boldsymbol{E}}}}^{\dagger}\widetilde{{\boldsymbol{E}}}}{4\upi}}-\overline{\frac{\widetilde{{\boldsymbol{E}}}}{4\upi}\cdot\bigg{(}\frac{c\widehat{\boldsymbol{k}}}{\widehat{\omega}}\times\widetilde{{\boldsymbol{B}}}\bigg{)}}
𝑬~𝑬~4\upi¯(𝑬~×c𝒌^ω^)𝑩~4\upi¯\displaystyle\approx-\overline{\frac{\smash{\widetilde{{\boldsymbol{E}}}}^{\dagger}\widetilde{{\boldsymbol{E}}}}{4\upi}}-\overline{\bigg{(}\widetilde{{\boldsymbol{E}}}\times\frac{c\widehat{\boldsymbol{k}}}{\widehat{\omega}}\bigg{)}\cdot\frac{\widetilde{{\boldsymbol{B}}}}{4\upi}}
14\upi(𝑩~𝑩~¯𝑬~𝑬~¯).\displaystyle\approx\frac{1}{4\upi}\,\big{(}\overline{\smash{\widetilde{{\boldsymbol{B}}}}^{\dagger}\widetilde{{\boldsymbol{B}}}}-\overline{\smash{\widetilde{{\boldsymbol{E}}}}^{\dagger}\widetilde{{\boldsymbol{E}}}}\big{)}. (622)

Substituting (621) and (622) into (620) leads to (469).

Appendix G Selected notation

This paper uses the following notation (also see section 2 for the index convention):

Symbol Definition Explanation
\smash{\sqbullet} placeholder
\smash{\sqbullet^{*}} complex conjugate
1\smash{\sqbullet^{-1}} inverse
\smash{\sqbullet^{\dagger}} Hermitian adjoint
\smash{\sqbullet^{-{\dagger}}} ()1\smash{(\sqbullet^{\dagger})^{-1}} inverse Hermitian adjoint
\smash{\sqbullet^{\intercal}} transpose
|a\smash{\sqbullet^{|a}} section 4.2 auxiliary notation
(μ)\smash{\sqbullet^{(\mu)}} section 6.5 contribution from the microscopic part
(m)\smash{\sqbullet^{\text{(m)}}} section 6.5 contribution from the macroscopic part
¯\smash{\overline{\sqbullet}} average part or, for eikonal waves, a quantity evaluated
on the local wavevector
~\smash{\widetilde{\sqbullet}} oscillatory part
¯\smash{\underline{\sqbullet}} macroscopic part
\textstyle\sqbullet  ~\widetilde{\vrule height=0.0pt,depth=0.0pt,width=9.72217pt} microscopic part
^\smash{\widehat{\sqbullet}} operator
˙\smash{\dot{\sqbullet}} time derivative
̊\smash{\mathring{\sqbullet}} (23) Fourier image
˘\smash{{\breve{\sqbullet}}} envelope of an eikonal (or monochromatic) wave
A\smash{\sqbullet_{\text{A}}} section 2.1.2 anti-Hermitian part
H\smash{\sqbullet_{\text{H}}} section 2.1.2 Hermitian part
\smash{\partial_{\sqbullet}} /\smash{\partial/\partial\sqbullet} partial derivative (but i/xi\smash{\partial_{i}\doteq\partial/\partial x^{i}}, α/zα\smash{\partial_{\alpha}\doteq\partial/\partial z^{\alpha}}, a/Xa\smash{\partial_{a}\doteq\partial/\partial X^{a}})
a,i\smash{\partial^{a},\partial^{i}} /Ka,/pi\smash{\partial/\partial K_{a},\partial/\partial p_{i}} partial derivative with respect to a lower-index quantity
ð\smash{\eth_{\sqbullet}} (149) auxiliary notation
dt\smash{\mathrm{d}_{t}} (116), (327) convective time derivative
{,}𝗑\smash{\{\sqbullet,\sqbullet\}_{\mathsf{x}}} (32) Poisson bracket on (𝘅,𝗸)\smash{({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})}
{,}\smash{\{\sqbullet,\sqbullet\}} (56) Poisson bracket on (𝒙,𝒌)\smash{({\boldsymbol{x}},{\boldsymbol{k}})}
[,]\smash{[\sqbullet,\sqbullet]} commutator
|\smash{\braket{\sqbullet}{\sqbullet}} (1), (59) inner product on 𝗑\smash{\mathscr{H}_{\mathsf{x}}} or on X\smash{\mathscr{H}_{X}}
\smash{\doteq} definition
\smash{\cdot} section 2.1.3 scalar product
\smash{\star} (30) Moyal product on (𝘅,𝗸)\smash{({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})}
🟊\smash{\bigstar} (72) Moyal product on (𝑿,𝑲)\smash{({\boldsymbol{X}},{\boldsymbol{K}})}
i0\smash{\mathrm{i}0} section 4.2 i\smash{\mathrm{i}} times an infinitesimally small positive number
\smash{\fint} principal-value integral
eigv eigenvalue
im\smash{\operatorname{im}} imaginary part
pv\smash{\operatorname{pv}} (146) auxiliary notation
re\smash{\operatorname{re}} real part
operX\smash{\text{oper}_{X}} operator corresponding to a Weyl symbol on (𝑿,𝑲)\smash{({\boldsymbol{X}},{\boldsymbol{K}})}
oper𝗑\smash{\text{oper}_{\mathsf{x}}} operator corresponding to a Weyl symbol on (𝘅,𝗸)\smash{({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})}
symb same as symbX\smash{\text{symb}_{X}} or symb𝗑\smash{\text{symb}_{\mathsf{x}}} when the two are equal
symbX\smash{\text{symb}_{X}} Weyl symbol of an operator on X\smash{\mathscr{H}_{X}}
symb𝗑\smash{\text{symb}_{\mathsf{x}}} Weyl symbol of an operator on 𝗑\smash{\mathscr{H}_{\mathsf{x}}}
sgn\smash{\operatorname{sgn}} sign
tr\smash{\operatorname{tr}} trace
Γ,Γs\smash{\Gamma,\Gamma_{s}} section 6.7 part of a collision operator
Δs\smash{\Delta_{s}} section 6.6 particle’s total ponderomotive energy in on-shell waves
Θαβc\smash{\Theta^{\alpha\beta c}} (169) auxiliary notation
Θαβ\smash{\Uptheta^{\alpha\beta}} (182) dressing function (since section 5.3)
Θij,𝚯\smash{\Uptheta_{ij},{\boldsymbol{\Uptheta}}} (202) dressing function (a part of Θαβ\smash{\Uptheta^{\alpha\beta}})
Λ\smash{\Lambda} section 7.1 dispersion function (one of Λb\smash{\Lambda_{b}})
Λb\smash{\Lambda_{b}} section 7.1 b\smash{b}th eigenvalue of 𝚵\smash{{\boldsymbol{\Xi}}}
𝚷s\smash{{\boldsymbol{\Pi}}_{s}} table 1 OC momentum flux density of species s\smash{s}
𝚷w\smash{{\boldsymbol{\Pi}}_{\text{w}}} table 1 wave momentum flux density
𝚵\smash{{\boldsymbol{\Xi}}} section 6.3 dispersion matrix
𝚵^\smash{\widehat{\boldsymbol{\Xi}}} (220), (224) dispersion operator
𝚵0\smash{{\boldsymbol{\Xi}}_{0}} section 6.3 vacuum dispersion matrix
𝚵^0\smash{\widehat{\boldsymbol{\Xi}}_{0}} (208) vacuum dispersion operator
𝚵p\smash{{\boldsymbol{\Xi}}_{p}} (228) Weyl symbol of 𝚵^p\smash{\widehat{\boldsymbol{\Xi}}_{p}}
𝚵^p\smash{\widehat{\boldsymbol{\Xi}}_{p}} section 6.1 auxiliary operator
Φ,Φs\smash{\Phi,\Phi_{s}} (184), (203) ponderomotive energy
Ψi,𝚿\smash{\Psi^{i},{\boldsymbol{\Psi}}} section 6.1 generic interaction field
Ψ~ci,𝚿~c\smash{\widetilde{\Psi}^{i}_{\text{c}},\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}} (308) complexified interaction field
Ω\smash{\Omega} (139) auxiliary notation
αs,i,𝜶s\smash{\alpha_{s,i},{\boldsymbol{\alpha}}_{s}} section 6.1 Weyl symbols of α^s,i\smash{\widehat{\alpha}_{s,i}} and 𝜶^s\smash{\widehat{\boldsymbol{\alpha}}_{s}}
α^s,i,𝜶^s\smash{\widehat{\alpha}_{s,i},\widehat{\boldsymbol{\alpha}}_{s}} section 6.1 coupling operators
γ\smash{\gamma} (338), (377) linear dissipation rate as a function of (t,𝒙,𝒌)\smash{(t,{\boldsymbol{x}},{\boldsymbol{k}})}
γ¯\smash{\overline{\gamma}} section 7.3 local linear dissipation rate of an eikonal wave
δ\smash{\delta} Kronecker symbol or delta function
ϵ\smash{\epsilon} section 3.1.1 geometrical-optics parameter
ϵ\smash{{\boldsymbol{\epsilon}}} (456) dielectric tensor
ϵ,ϵ\smash{\epsilon_{\parallel},\epsilon_{\perp}} (423) parallel and transverse permittivity
ε\smash{\varepsilon} section 3.1.1 small parameter proportional to the oscillation amplitude
𝜼\smash{{\boldsymbol{\eta}}} section 7.1 polarization vector (one of 𝜼b\smash{{\boldsymbol{\eta}}_{b}})
𝜼b\smash{{\boldsymbol{\eta}}_{b}} section 7.1 b\smash{b}th eigenvector of 𝚵\smash{{\boldsymbol{\Xi}}}
θ\smash{\theta} eikonal phase
κ\smash{\kappa} section 7.2.3 auxiliary notation
κx,κp\smash{\kappa_{x},\kappa_{p}} (86) characteristic inverse scales in 𝒙\smash{{\boldsymbol{x}}} and 𝒑\smash{{\boldsymbol{p}}}, respectively
ρs\smash{\rho_{s}} charge density of species s\smash{s}
ϱαβ\smash{\varrho^{\alpha\beta}} (178) auxiliary notation
s\smash{{\boldsymbol{\wp}}_{s}} (229) Weyl symbol of ^s\smash{\widehat{\boldsymbol{\wp}}_{s}}
^s\smash{\widehat{\boldsymbol{\wp}}_{s}} (229) coupling operator
σ\smash{\sigma} (280) entropy density
ς𝒌\smash{\varsigma_{{\boldsymbol{k}}}} (373) sign of the action density
φ\smash{\varphi} electrostatic potential
ψ,𝝍\smash{\psi,{\boldsymbol{\psi}}} any field
ω\smash{\omega} coordinate in the frequency space dual to t\smash{t}
ω¯\smash{\overline{\omega}} tθ\smash{-\partial_{t}\theta} local frequency of an eikonal wave
ω^\smash{\widehat{\omega}} it\smash{\mathrm{i}\partial_{t}} frequency operator
C\smash{C} (89) Fourier image of W\smash{W}
C,𝑪\smash{C_{\sqbullet},{\boldsymbol{C}}_{\sqbullet}} (77), (82) Fourier images of W\smash{W_{\sqbullet}} and 𝑾\smash{{\boldsymbol{W}}_{\sqbullet}}
𝒞s\smash{\mathcal{C}_{s}} section 6.8 collision operator of species s\smash{s}
𝖢\smash{\mathsf{C}_{\sqbullet}}, 𝗖\smash{{\boldsymbol{\mathsf{C}}}_{\sqbullet}} (44), (52) Fourier images of 𝖶\smash{\mathsf{W}_{\sqbullet}} and 𝗪\smash{{\boldsymbol{\mathsf{W}}}_{\sqbullet}}
Dαβ\smash{D^{\alpha\beta}} section 5.1 Weyl symbol of D^αβ\smash{\widehat{D}^{\alpha\beta}}
𝖣αβ\smash{\mathsf{D}^{\alpha\beta}} (177) phase-space-diffusion coefficient
𝖣ij,𝗗s\smash{\mathsf{D}_{ij},{\boldsymbol{\mathsf{D}}}_{s}} (201), (284a) momentum-diffusion coefficient (part of 𝖣αβ\smash{\mathsf{D}^{\alpha\beta}})
D0αβ\smash{D_{0}^{\alpha\beta}} (168) auxiliary notation
D^αβ\smash{\widehat{D}^{\alpha\beta}} (118) diffusion operator on X\smash{\mathscr{H}_{X}}
s\smash{\mathcal{E}_{s}} table 1 OC energy density of species s\smash{s}
w\smash{\mathcal{E}_{\text{w}}} table 1 wave energy density (also see (333) for eikonal waves)
F,Fs\smash{F,F_{s}} (187), (198) OC distribution functions
𝔉s\smash{\mathfrak{F}_{s}} (267) polarization drag for species s\smash{s}
G,Gs\smash{G,G_{s}} section 4.2 Weyl symbols of G^\smash{\widehat{G}} and G^s\smash{\widehat{G}_{s}}
G0\smash{G_{0}} (145) approximation of G\smash{G} to the zeroth order in ϵ\smash{\epsilon}
G^,G^s\smash{\widehat{G},\widehat{G}_{s}} (109) effective Green’s operators on X\smash{\mathscr{H}_{X}}
𝔊ss\smash{\mathfrak{G}_{ss^{\prime}}} (237) spectrum of the correlations between gs\smash{g_{s}} and gs\smash{g_{s^{\prime}}}
𝒢^\smash{\widehat{\mathscr{G}}} section 3.2 Green’s operator on X\smash{\mathscr{H}_{X}}
\smash{\mathcal{I}} (318) action density of an eikonal wave
J\smash{J} section 7.4.3 phase-space action density
Jαβ,𝑱\smash{J^{\alpha\beta},{\boldsymbol{J}}} (57) canonical Poisson structure
𝒥i,𝓙\smash{\mathcal{J}^{i},{\boldsymbol{\mathcal{J}}}} (319) action flux density of an eikonal wave
H,Hs\smash{H,H_{s}} particle Hamiltonian
s\smash{\mathcal{H}_{s}} (200) OC Hamiltonian of species s\smash{s}
𝗑\smash{\mathscr{H}_{\mathsf{x}}} Hilbert space formed by functions on 𝘅\smash{{\boldsymbol{\mathsf{x}}}}
X\smash{\mathscr{H}_{X}} Hilbert space formed by functions on 𝑿\smash{{\boldsymbol{X}}}
𝑲\smash{{\boldsymbol{K}}} (ω,𝒒)\smash{(-\omega,{\boldsymbol{q}})} coordinate in the wavevector space dual to 𝑿\smash{{\boldsymbol{X}}}
𝑲^\smash{\widehat{\boldsymbol{K}}} (ω^,𝒒^)\smash{(-\widehat{\omega},\widehat{\boldsymbol{q}})} wavevector operator on X\smash{\mathscr{H}_{X}}
L^\smash{\widehat{L}} (98) extended Liouvillian (up to a factor i\smash{\mathrm{i}})
𝑳^s\smash{\widehat{\boldsymbol{L}}_{s}} (209) coupling operator
^X\smash{\widehat{\mathcal{L}}_{X}} (73) same as the Poisson bracket on (𝑿,𝑲)\smash{({\boldsymbol{X}},{\boldsymbol{K}})}
^𝗑\smash{\widehat{\mathcal{L}}_{\mathsf{x}}} (32) same as the Poisson bracket on (𝘅,𝗸)\smash{({\boldsymbol{\mathsf{x}}},{\boldsymbol{\mathsf{k}}})}
𝔏\smash{\mathfrak{L}} (213) 𝚿\smash{{\boldsymbol{\Psi}}}-dependent part of the plasma Lagrangian density
𝔏0\smash{\mathfrak{L}_{0}} (208) Lagrangian density of 𝚿\smash{{\boldsymbol{\Psi}}} in vacuum
M\smash{M} number of components of 𝚿\smash{{\boldsymbol{\Psi}}} or of another vector field
N\smash{N} 2n+1\smash{2n+1} dimension of the extended phase space 𝑿\smash{{\boldsymbol{X}}}
𝒩,𝒩s\smash{\mathcal{N},\mathcal{N}_{s}} (199) OC density
𝒪\smash{\mathcal{O}} big O (‘at most of the order of’)
Psi,𝑷s\smash{P_{s}^{i},{\boldsymbol{P}}_{s}} table 1 OC momentum density of species s\smash{s}
Pwi,𝑷w\smash{P^{i}_{\text{w}},{\boldsymbol{P}}_{\text{w}}} table 1 wave momentum density (also see (333) for eikonal waves)
Qsi,𝑸s\smash{Q_{s}^{i},{\boldsymbol{Q}}_{s}} table 1 OC energy flux density of species s\smash{s}
Qwi,𝑸w\smash{Q^{i}_{\text{w}},{\boldsymbol{Q}}_{\text{w}}} table 1 wave energy flux density
𝒬ss\smash{\mathcal{Q}_{ss^{\prime}}} section 6.8 symmetrized coefficient in the collision operator
𝑹^s\smash{\widehat{\boldsymbol{R}}_{s}} (209) coupling operator
\smash{\mathbb{R}} real axis
S\smash{S} section 7.1 action integral
Sad\smash{S_{\text{ad}}} (223) adiabatic action integral
𝗦\smash{{\boldsymbol{\mathsf{S}}}} (399) spectrum of the macroscopic oscillations
T^τ\smash{\widehat{T}_{\tau}} (105) shift operator (see also section 4.1)
𝖴ij,𝗨\smash{\mathsf{U}^{ij},{\boldsymbol{\mathsf{U}}}} (285) average Wigner function of the macroscopic field 𝚿¯~(𝘅)\smash{\underline{\widetilde{{\boldsymbol{\Psi}}}}({\boldsymbol{\mathsf{x}}})}
𝗨c±,𝗨c\smash{{\boldsymbol{\mathsf{U}}}_{\text{c}\pm},{\boldsymbol{\mathsf{U}}}_{\text{c}}} section 7.4.2 average Wigner matrix of 𝚿~c\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}} and 𝚿~c\smash{\smash{\widetilde{{\boldsymbol{\Psi}}}}_{\text{c}}^{*}} (𝗨c𝗨c+\smash{{\boldsymbol{\mathsf{U}}}_{\text{c}}\equiv{\boldsymbol{\mathsf{U}}}_{\text{c}+}})
Va,𝑽\smash{V^{a},{\boldsymbol{V}}} section 3.2 unperturbed velocity in the 𝑿\smash{{\boldsymbol{X}}} space
𝒱n\smash{\mathscr{V}_{n}} volume of n\smash{n}-dimensional homogeneous plasma
W\smash{W} (88) Wigner function of H~(𝑿)\smash{\widetilde{H}({\boldsymbol{X}})}
W\smash{W_{\sqbullet}}, 𝑾\smash{{\boldsymbol{W}}_{\sqbullet}} (76), (81) Weyl symbol of W^\smash{\widehat{W}_{\sqbullet}} (Wigner function or matrix)
W^\smash{\widehat{W}} (87) density operator on X\smash{\mathscr{H}_{X}} of H~\smash{\widetilde{H}}
W^\smash{\widehat{W}_{\sqbullet}}, 𝑾^\smash{\widehat{\boldsymbol{W}}_{\sqbullet}} (75), (80) density operator on X\smash{\mathscr{H}_{X}} of a given field
𝖶,𝖶s\smash{\mathsf{W},\mathsf{W}_{s}} section 5.6 Wigner functions of H~\smash{\widetilde{H}} and H~s\smash{\widetilde{H}_{s}} with 𝒑\smash{{\boldsymbol{p}}} as a parameter
𝖶,𝗪\smash{\mathsf{W}_{\sqbullet},{\boldsymbol{\mathsf{W}}}_{\sqbullet}} (43), (51) Weyl symbol of 𝖶^\smash{\widehat{\mathsf{W}}_{\sqbullet}} (Wigner function or matrix)
𝖶^,𝗪^\smash{\widehat{\mathsf{W}}_{\sqbullet},\widehat{\boldsymbol{{\boldsymbol{\mathsf{W}}}}}_{\sqbullet}} (42), (50) density operator on X\smash{\mathscr{H}_{X}} of a given field
𝖂\smash{{\boldsymbol{\mathfrak{W}}}} section 6.4 average Wigner matrix of the microscopic field ~Ψ ~  (𝘅)\smash{\mathchoice{{\vtop{\hbox{$\displaystyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\textstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}{{\vtop{\hbox{$\scriptscriptstyle\widetilde{{\boldsymbol{\Psi}}}$}\hbox{$\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}$}\vskip-4.30554pt}}}({\boldsymbol{\mathsf{x}}})}
Xa,𝑿\smash{X^{a},{\boldsymbol{X}}} (t,𝒛)\smash{(t,{\boldsymbol{z}})} coordinate in the extended phase space
X^a,𝑿^\smash{\widehat{X}^{a},\widehat{\boldsymbol{X}}} (t^,𝒛^)\smash{(\widehat{t},\widehat{\boldsymbol{z}})} operator of the position in the extended phase space
𝒳ss\smash{\mathcal{X}_{ss^{\prime}}} (243) Weyl symbol of 𝒳^ss\smash{\widehat{\mathcal{X}}_{ss^{\prime}}}
𝒳^ss\smash{\widehat{\mathcal{X}}_{ss^{\prime}}} (242) coupling operators on 𝗑\smash{\mathscr{H}_{\mathsf{x}}} that enter H~\textstyle\widetilde{H}  ~\widetilde{\vrule height=0.0pt,depth=0.0pt,width=5.55557pt}
d\smash{\mathrm{d}} differential
es\smash{e_{s}} charge of species s\smash{s}
f,fs\smash{f,f_{s}} distribution function
g,gs\smash{g,g_{s}} section 4.3 initial conditions for f~\smash{\widetilde{f}} and f~s\smash{\widetilde{f}_{s}}
h\smash{h} (355), (365) rescaled phase-space action density
hc±,hc\smash{h_{\text{c}\pm},h_{\text{c}}} section 7.4.2 auxiliary notation (hc+hc\smash{h_{\text{c}+}\equiv h_{\text{c}}})
ki,𝒌\smash{k_{i},{\boldsymbol{k}}} coordinate in the wavevector space dual to 𝒙\smash{{\boldsymbol{x}}}
k¯i,𝒌¯\smash{\overline{k}_{i},\overline{{\boldsymbol{k}}}} iθ,𝒙θ\smash{\partial_{i}\theta,\partial_{\boldsymbol{x}}\theta} local wavevector of an eikonal wave
k^i,𝒌^\smash{\widehat{k}_{i},\widehat{\boldsymbol{k}}} ii,i𝒙\smash{-\mathrm{i}\partial_{i},-\mathrm{i}\partial_{\boldsymbol{x}}} wavevector operator
𝗸\smash{{\boldsymbol{\mathsf{k}}}} (ω,𝒌)\smash{(-\omega,{\boldsymbol{k}})} coordinate in the wavevector space dual to 𝘅\smash{{\boldsymbol{\mathsf{x}}}}
𝗸¯\smash{\overline{{\boldsymbol{\mathsf{k}}}}} (ω¯,𝒌¯)\smash{(-\overline{\omega},\overline{{\boldsymbol{k}}})} local spacetime-wavevector of an eikonal wave
𝗸^\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}}} i𝘅\smash{-\mathrm{i}\partial_{\boldsymbol{\mathsf{x}}}} spacetime-wavevector operator
|𝗸\smash{\ket{{\boldsymbol{\mathsf{k}}}}} (16), (17) eigenvector of 𝗸^\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{k}}}}}} corresponding to the eigenvalue 𝗸\smash{{\boldsymbol{\mathsf{k}}}}
τ\smash{{\boldsymbol{\ell}}_{\tau}} section 4.1.1 displacement in 𝑿\smash{{\boldsymbol{X}}} along unperturbed characteristics
ms\smash{m_{s}} mass of species s\smash{s}
n\smash{n} dim𝒙\smash{\dim{\boldsymbol{x}}} number of spatial dimensions
𝗇\smash{\mathsf{n}} n+1\smash{n+1} number of spacetime dimensions
pi,𝒑\smash{p_{i},{\boldsymbol{p}}} coordinate in the momentum space
p^i,𝒑^\smash{\widehat{p}_{i},\widehat{\boldsymbol{p}}} 𝒑\smash{{\boldsymbol{p}}} position operator corresponding to the coordinate 𝒑\smash{{\boldsymbol{p}}}
qi,𝒒\smash{q^{i},{\boldsymbol{q}}} (𝒌,𝒓)\smash{({\boldsymbol{k}},{\boldsymbol{r}})} coordinate in the wavevector space dual to 𝒛\smash{{\boldsymbol{z}}}
q^i,𝒒^\smash{\widehat{q}^{i},\widehat{\boldsymbol{q}}} (𝒌^,𝒓^)\smash{(\widehat{\boldsymbol{k}},\widehat{\boldsymbol{r}})} wavevector operator corresponding to the coordinate 𝒛\smash{{\boldsymbol{z}}}
ri,𝒓\smash{r_{i},{\boldsymbol{r}}} coordinate in the wavevector space dual to 𝒑\smash{{\boldsymbol{p}}}
r^i,𝒓^\smash{\widehat{r}^{i},\widehat{\boldsymbol{r}}} ii,i𝒑\smash{-\mathrm{i}\partial^{i},-\mathrm{i}\partial_{\boldsymbol{p}}} wavevector operator corresponding to the coordinate 𝒑\smash{{\boldsymbol{p}}}
s\smash{s} species index
t\smash{t} time
t^\smash{\widehat{t}} t\smash{t} time operator
uα\smash{u^{\alpha}} (92) oscillating part of the phase-space velocity
u^α\smash{\widehat{u}^{\alpha}} (114) uα\smash{u^{\alpha}} as an operator on X\smash{\mathscr{H}_{X}}
vα,vi,𝒗\smash{v^{\alpha},v^{i},{\boldsymbol{v}}} (92) average velocity in phase space or in physical space,
(192) or, since section 5.4, OC velocity
vgi,𝒗g\smash{v_{\text{g}}^{i},{\boldsymbol{v}}_{\text{g}}} (316) group velocity as a function of (t,𝒙,𝒌)\smash{(t,{\boldsymbol{x}},{\boldsymbol{k}})}
v¯gi,𝒗¯g\smash{\overline{v}_{\text{g}}^{i},\overline{{\boldsymbol{v}}}_{\text{g}}} (320) local group velocity of an eikonal wave
w\smash{w} (313) eikonal-wave frequency as a function of (t,𝒙,𝒌)\smash{(t,{\boldsymbol{x}},{\boldsymbol{k}})}
xi,𝒙\smash{x^{i},{\boldsymbol{x}}} coordinate in space
x¯i,𝒙¯\smash{\overline{x}^{i},\overline{{\boldsymbol{x}}}} ray coordinate in space
x^i,𝒙^\smash{\widehat{x}^{i},\widehat{\boldsymbol{x}}} 𝒙\smash{{\boldsymbol{x}}} space-position operator
𝗑i,𝘅\smash{\mathsf{x}^{i},{\boldsymbol{\mathsf{x}}}} (t,𝒙)\smash{(t,{\boldsymbol{x}})} coordinate in spacetime
𝗑^i,𝘅^\smash{\widehat{\mathsf{x}}^{i},\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}}} 𝘅\smash{{\boldsymbol{\mathsf{x}}}} spacetime-position operator
|𝘅\smash{\ket{{\boldsymbol{\mathsf{x}}}}} (16), (17) eigenvector of 𝘅^\smash{\widehat{\boldsymbol{{\boldsymbol{\mathsf{x}}}}}} corresponding to the eigenvalue 𝘅\smash{{\boldsymbol{\mathsf{x}}}}
zα,𝒛\smash{z^{\alpha},{\boldsymbol{z}}} (𝒙,𝒑)\smash{({\boldsymbol{x}},{\boldsymbol{p}})} coordinate in phase space
z^α,𝒛^\smash{\widehat{z}^{\alpha},\widehat{\boldsymbol{z}}} (𝒙^,𝒑^)\smash{(\widehat{\boldsymbol{x}},\widehat{\boldsymbol{p}})} phase-space position operator

References

  • Andersson et al. (2021) Andersson, L., Joudioux, J., Oancea, M. A. & Raj, A. 2021 Propagation of polarized gravitational waves. Phys. Rev. D 103, 044053.
  • Balakin et al. (2016) Balakin, A. A., Dodin, I. Y., Fraiman, G. M. & Fisch, N. J. 2016 Backward Raman amplification of broad-band pulses. Phys. Plasmas 23, 083115.
  • Besse et al. (2011) Besse, N., Elskens, Y., Escande, D. F. & Bertrand, P. 2011 Validity of quasilinear theory: refutations and new numerical confirmation. Plasma Phys. Control. Fusion 53, 025012.
  • Binney & Tremaine (2008) Binney, J. & Tremaine, S. 2008 Galactic Dynamics: Second Edition. Princeton: Princeton Univ. Press.
  • Bliokh et al. (2015) Bliokh, K. Y., Rodríguez-Fortuño, F. J., Nori, F. & Zayats, A. V. 2015 Spin-orbit interactions of light. Nat. Photonics 9, 796.
  • Brizard et al. (1993) Brizard, A. J., Cook, D. R. & Kaufman, A. N. 1993 Wave-action conservation for pseudo-Hermitian fields. Phys. Rev. Lett. 70, 521.
  • Cartwright (1976) Cartwright, N. D. 1976 A non-negative Wigner-type distribution. Physica A: Stat. Mech. Appl. 83, 210.
  • Cary & Brizard (2009) Cary, J. R. & Brizard, A. J. 2009 Hamiltonian theory of guiding-center motion. Rev. Mod. Phys. 81, 693.
  • Cary & Kaufman (1977) Cary, J. R. & Kaufman, A. N. 1977 Ponderomotive force and linear susceptibility in Vlasov plasma. Phys. Rev. Lett. 39, 402.
  • Cary & Kaufman (1981) Cary, J. R. & Kaufman, A. N. 1981 Ponderomotive effects in collisionless plasma: a Lie transform approach. Phys. Fluids 24, 1238.
  • Catto et al. (2017) Catto, P. J., Lee, J. & Ram, A. K. 2017 A quasilinear operator retaining magnetic drift effects in tokamak geometry. J. Plasma Phys. 83, 905830611.
  • Chavanis (2012) Chavanis, P.-H. 2012 Kinetic theory of long-range interacting systems with angle–action variables and collective effects. Physica A 391, 3680.
  • Crews & Shumlak (2022) Crews, D. W. & Shumlak, U. 2022 On the validity of quasilinear theory applied to the electron bump-on-tail instability. Phys. Plasmas 29, 043902.
  • Dewar (1972) Dewar, R. L. 1972 A Lagrangian theory for nonlinear wave packets in a collisionless plasma. J. Plasma Phys. 7, 267.
  • Dewar (1973) Dewar, R. L. 1973 Oscillation center quasilinear theory. Phys. Fluids 16, 1102.
  • Dewar (1977) Dewar, R. L. 1977 Energy-momentum tensors for dispersive electromagnetic waves. Austral. J. Phys. 30, 533.
  • Dodin (2014) Dodin, I. Y. 2014 On variational methods in the physics of plasma waves. Fusion Sci. Tech. 65, 54.
  • Dodin & Fisch (2010a) Dodin, I. Y. & Fisch, N. J. 2010a On generalizing the KK-χ\chi theorem. Phys. Lett. A 374, 3472.
  • Dodin & Fisch (2010b) Dodin, I. Y. & Fisch, N. J. 2010b On the evolution of linear waves in cosmological plasmas. Phys. Rev. D 82, 044044.
  • Dodin & Fisch (2012) Dodin, I. Y. & Fisch, N. J. 2012 Axiomatic geometrical optics, Abraham–Minkowski controversy, and photon properties derived classically. Phys. Rev. A 86, 053834.
  • Dodin & Fisch (2014) Dodin, I. Y. & Fisch, N. J. 2014 Ponderomotive forces on waves in modulated media. Phys. Rev. Lett. 112, 205002.
  • Dodin et al. (2009) Dodin, I. Y., Geyko, V. I. & Fisch, N. J. 2009 Langmuir wave linear evolution in inhomogeneous nonstationary anisotropic plasma. Phys. Plasmas 16, 112101.
  • Dodin et al. (2019) Dodin, I. Y., Ruiz, D. E., Yanagihara, K., Zhou, Y. & Kubo, S. 2019 Quasioptical modeling of wave beams with and without mode conversion. I. Basic theory. Phys. Plasmas 26, 072110.
  • Dodin et al. (2017) Dodin, I. Y., Zhmoginov, A. I. & Ruiz, D. E. 2017 Variational principles for dissipative (sub)systems, with applications to the theory of linear dispersion and geometrical optics. Phys. Lett. A 381, 1411.
  • Drummond & Pines (1962) Drummond, W. E. & Pines, D. 1962 Non-linear stability of plasma oscillations. Nucl. Fusion Suppl. 3, 1049.
  • Eriksson & Helander (1994) Eriksson, L.-G. & Helander, P. 1994 Monte Carlo operators for orbit-averaged Fokker–Planck equations. Phys. Plasmas 1, 308.
  • Escande et al. (2018) Escande, D. F., Bénisti, D., Elskens, Y., Zarzoso, D. & Doveil, F. 2018 Basic microscopic plasma physics from N{N}-body mechanics. Rev. Mod. Plasma Phys. 2, 1, arXiv: 1805.11408.
  • Fetterman & Fisch (2008) Fetterman, A. J. & Fisch, N. J. 2008 α\alpha channeling in a rotating plasma. Phys. Rev. Lett. 101, 205003.
  • Fisch (1987) Fisch, N. J. 1987 Theory of current drive in plasmas. Rev. Mod. Phys. 59, 175.
  • Fisch & Rax (1992) Fisch, N. J. & Rax, J. M. 1992 Interaction of energetic alpha-particles with intense lower hybrid waves. Phys. Rev. Lett. 69, 612.
  • Fraiman & Kostyukov (1995) Fraiman, G. M. & Kostyukov, I. Yu. 1995 Influence of external inhomogeneous static fields on interaction between beam of charged-particles and packet of electromagnetic waves. Phys. Plasmas 2, 923.
  • Galeev & Sagdeev (1985) Galeev, A. A. & Sagdeev, R. Z. 1985 Theory of Weakly Turbulent Plasma, Part 4 in ‘Basic Plasma Physics I’. New York: North–Holland, edited by A. A. Galeev and R. N. Sudan.
  • Gaponov & Miller (1958) Gaponov, A. V. & Miller, M. A. 1958 Potential wells for charged particles in a high-frequency electromagnetic field. Zh. Eksp. Teor. Fiz. 34, 242.
  • Garg & Dodin (2020) Garg, D. & Dodin, I. Y. 2020 Average nonlinear dynamics of particles in gravitational pulses: effective Hamiltonian, secular acceleration, and gravitational susceptibility. Phys. Rev. D 102, 064012.
  • Garg & Dodin (2021a) Garg, G. & Dodin, I. Y. 2021a Gauge-invariant gravitational waves in matter beyond linearized gravity, arXiv: 2106.05062.
  • Garg & Dodin (2021b) Garg, G. & Dodin, I. Y. 2021b Gauge invariants of linearized gravity with a general background metric, arXiv: 2105.04680.
  • Garg & Dodin (2022) Garg, G. & Dodin, I. Y. 2022 Gravitational wave modes in matter, arXiv: 2204.09095.
  • Hamilton (2020) Hamilton, C. 2020 A simple, heuristic derivation of the Balescu-Lenard kinetic equation for stellar systems. Mon. Notices Royal Astron. Soc. 501, 3371.
  • Hayes (1973) Hayes, W. D. 1973 Group velocity and nonlinear dispersive wave propagation. Proc. R. Soc. Lond. A 332, 199.
  • Hizanidis et al. (1983) Hizanidis, K., Molvig, K. & Swartz, K. 1983 A retarded time superposition principle and the relativistic collision operator. J. Plasma Phys. 30, 223.
  • Kaufman (1972) Kaufman, A. N. 1972 Quasilinear diffusion of an axisymmetric toroidal plasma. Phys. Fluids 15, 1063.
  • Kaufman (1987) Kaufman, A. N. 1987 Phase-space-Lagrangian action principle and the generalized KK-χ\chi theorem. Phys. Rev. A 36, 982.
  • Kaufman & Holm (1984) Kaufman, A. N. & Holm, D. D. 1984 The Lie-transformed Vlasov action principle: relativistically covariant wave propagation and self-consistent ponderomotive effects. Phys. Lett. A 105, 277.
  • Kennel & Engelmann (1966) Kennel, C. F. & Engelmann, F. 1966 Velocity space diffusion from weak plasma turbulence in a magnetic field. Phys. Fluids 9, 2377.
  • Kentwell (1987) Kentwell, G. W. 1987 Oscillation-center theory at resonance. Phys. Rev. A 35, 4703.
  • Kentwell & Jones (1987) Kentwell, G. W. & Jones, D. A. 1987 The time-dependent ponderomotive force. Phys. Rep. 145, 319.
  • Krall & Trivelpiece (1973) Krall, N. A. & Trivelpiece, A. W. 1973 Principles of Plasma Physics. New York: McGraw-Hill.
  • Landau & Lifshitz (1976) Landau, L. D. & Lifshitz, E. M. 1976 Mechanics. Oxford: Butterworth–Heinemann.
  • Landreman (2017) Landreman, M. 2017 The H theorem for the Landau–Fokker–Planck collision operator. Unpublished.
  • Lee et al. (2018) Lee, J., Smithe, D., Wright, J. & Bonoli, P. 2018 A positive-definite form of bounce-averaged quasilinear velocity diffusion for the parallel inhomogeneity in a tokamak. Plasma Phys. Control. Fusion 60, 025007.
  • Lichtenberg & Lieberman (1992) Lichtenberg, A. J. & Lieberman, M. A. 1992 Regular and Chaotic Dynamics. New York: Springer–Verlag, second edition.
  • Lifshitz & Pitaevskii (1981) Lifshitz, E. M. & Pitaevskii, L. P. 1981 Physical Kinetics. New York: Pergamon Press.
  • Littlejohn (1979) Littlejohn, R. G. 1979 A guiding center Hamiltonian: a new approach. J. Math. Phys. 20, 2445.
  • Littlejohn (1981) Littlejohn, R. G. 1981 Hamiltonian formulation of guiding center motion. Phys. Fluids 24, 1730.
  • Littlejohn (1983) Littlejohn, R. G. 1983 Variational principles of guiding centre motion. J. Plasma Phys. 29, 111.
  • Littlejohn (1986) Littlejohn, R. G. 1986 The semiclassical evolution of wave packets. Phys. Rep. 138, 193.
  • Liu & Dodin (2015) Liu, C. & Dodin, I. Y. 2015 Nonlinear frequency shift of electrostatic waves in general collisionless plasma: unifying theory of fluid and kinetic nonlinearities. Phys. Plasmas 22, 082117.
  • Magorrian (2021) Magorrian, J. 2021 Stellar dynamics in the periodic cube. Mon. Notices Royal Astron. Soc. 507, 4840.
  • McDonald (1988) McDonald, S. W. 1988 Phase-space representations of wave equations with applications to the eikonal approximation for short-wavelength waves. Phys. Rep. 158, 337.
  • McDonald (1991) McDonald, S. W. 1991 Wave kinetic equation in a fluctuating medium. Phys. Rev. A 43, 4484.
  • McDonald et al. (1985) McDonald, S. W., Grebogi, C. & Kaufman, A. N. 1985 Locally coupled evolution of wave and particle distribution in general magnetoplasma geometry. Phys. Lett. A 111, 19.
  • McDonald & Kaufman (1985) McDonald, S. W. & Kaufman, A. N. 1985 Weyl representation for electromagnetic waves: The wave kinetic equation. Phys. Rev. A 32, 1708.
  • Motz & Watson (1967) Motz, H. & Watson, C. J. H. 1967 The radio-frequency confinement and acceleration of plasmas. Adv. Electron. Electron Phys. 23, 153.
  • Moyal (1949) Moyal, J. E. 1949 Quantum mechanics as a statistical theory. Proc. Cambridge Philosoph. Soc. 45, 99.
  • Mynick (1988) Mynick, H. E. 1988 The generalized Balescu-Lenard collision operator. J. Plasma Phys. 39, 303.
  • Oancea et al. (2020) Oancea, M. A., Joudioux, J., Dodin, I. Y., Ruiz, D. E., Paganini, C. F. & Andersson, L. 2020 Gravitational spin Hall effect of light. Phys. Rev. D 102, 024075.
  • Ochs (2021) Ochs, I. E. 2021 Controlling and exploiting perpendicular rotation in magnetized plasmas. PhD thesis, Princeton University.
  • Ochs & Fisch (2021a) Ochs, I. E. & Fisch, N. J. 2021a Nonresonant diffusion in alpha channeling. Phys. Rev. Lett. 127, 025003.
  • Ochs & Fisch (2021b) Ochs, I. E. & Fisch, N. J. 2021b Wave-driven torques to drive current and rotation. Phys. Plasmas 28, 102506.
  • Ochs & Fisch (2022) Ochs, I. E. & Fisch, N. J. 2022 Momentum conservation in current drive and alpha-channeling-mediated rotation drive, arXiv: 2201.07853.
  • Pinsker (2001) Pinsker, R. I. 2001 Introduction to wave heating and current drive in magnetized plasmas. Phys. Plasmas 8, 1219.
  • Rigas et al. (2011) Rigas, I., Sánchez-Soto, L. L., Klimov, A., Řeháček, J. & Hradil, Z. 2011 Orbital angular momentum in phase space. Ann. Phys. 326, 426.
  • Rogister & Oberman (1968) Rogister, A. & Oberman, C. 1968 On the kinetic theory of stable and weakly unstable plasma. Part 1. J. Plasma Phys. 2, 33.
  • Rogister & Oberman (1969) Rogister, A. & Oberman, C. 1969 On the kinetic theory of stable and weakly unstable plasma. Part 2. J. Plasma Phys. 3, 119.
  • Rostoker (1964) Rostoker, N. 1964 Superposition of dressed test particles. Phys. Fluids 7, 479.
  • Ruiz (2017) Ruiz, D. E. 2017 Geometric theory of waves and its applications to plasma physics. PhD thesis, Princeton University, arXiv:1708.05423.
  • Ruiz & Dodin (2015a) Ruiz, D. E. & Dodin, I. Y. 2015a First-principles variational formulation of polarization effects in geometrical optics. Phys. Rev. A 92, 043805.
  • Ruiz & Dodin (2015b) Ruiz, D. E. & Dodin, I. Y. 2015b On the correspondence between quantum and classical variational principles. Phys. Lett. A 379, 2623.
  • Ruiz & Dodin (2017a) Ruiz, D. E. & Dodin, I. Y. 2017a Extending geometrical optics: a Lagrangian theory for vector waves. Phys. Plasmas 24, 055704.
  • Ruiz & Dodin (2017b) Ruiz, D. E. & Dodin, I. Y. 2017b Ponderomotive dynamics of waves in quasiperiodically modulated media. Phys. Rev. A 95, 032114.
  • Ruiz et al. (2019) Ruiz, D. E., Glinsky, M. E. & Dodin, I. Y. 2019 Wave kinetic equation for inhomogeneous drift-wave turbulence beyond the quasilinear approximation. J. Plasma Phys. 85, 905850101.
  • Schlickeiser & Yoon (2014) Schlickeiser, R. & Yoon, P. H. 2014 Quasilinear theory of general electromagnetic fluctuations in unmagnetized plasmas. Phys. Plasmas 21, 092102.
  • Schmit et al. (2010) Schmit, P. F., Dodin, I. Y. & Fisch, N. J. 2010 Controlling hot electrons by wave amplification and decay in compressing plasma. Phys. Rev. Lett. 105, 175003.
  • Silin (1961) Silin, V. P. 1961 Collision integral for charged particles. Zh. Eksp. Teor. Fiz. 40, 1768.
  • Stix (1992) Stix, T. H. 1992 Waves in Plasmas. New York: AIP, second edition.
  • Tracy et al. (2014) Tracy, E. R., Brizard, A. J., Richardson, A. S. & Kaufman, A. N. 2014 Ray Tracing and Beyond: Phase Space Methods in Plasma Wave Theory. New York: Cambridge University Press.
  • Trigger et al. (2004) Trigger, S. A., Ershkovich, A. I., van Heijst, G. J. F. & Schram, P. P. J. M. 2004 Kinetic theory of Jeans instability. Phys. Rev. E 69, 066403.
  • Vedenov et al. (1961) Vedenov, A. A., Velikhov, E. P. & Sagdeev, R. Z. 1961 Nonlinear oscillations of rarified plasma. Nucl. Fusion 1, 82.
  • Weibel (1981) Weibel, E. S. 1981 Quasi-linear theory without the random phase approximation. Phys. Fluids 24, 413.
  • Whitham (1974) Whitham, G. B. 1974 Linear and Nonlinear Waves. New York: Wiley.
  • Wong (2000) Wong, H. V. 2000 Particle canonical variables and guiding center Hamiltonian up to second order in the Larmor radius. Phys. Plasmas 7, 73.
  • Yasseen (1983) Yasseen, F. 1983 Quasilinear theory of inhomogeneous magnetized plasmas. Phys. Fluids 26, 468.
  • Yasseen & Vaclavik (1986) Yasseen, F. & Vaclavik, J. 1986 Quasilinear theory of uniformly magnetized inhomogeneous plasmas: electromagnetic fluctuations. Phys. Fluids 29, 450.
  • Ye & Kaufman (1992) Ye, H. & Kaufman, A. N. 1992 Self-consistent theory for ion gyroresonance. Phys. Fluids B 4, 1735.
  • Yoon et al. (2016) Yoon, P. H., Ziebell, L. F., Kontar, E. P. & Schlickeiser, R. 2016 Weak turbulence theory for collisional plasmas. Phys. Rev. E 93, 033203.
  • Zakharov et al. (1992) Zakharov, V. E., L’vov, V. S. & Falkovich, G. 1992 Kolmogorov Spectra of Turbulence I: Wave Turbulence. New York: Springer-Verlag.
  • Zhu & Dodin (2021) Zhu, H. & Dodin, I. Y. 2021 Wave-kinetic approach to zonal-flow dynamics: recent advances. Phys. Plasmas 28, 032303.