This work opens a series of papers where we develop a general quasi-optical theory for mode-converting electromagnetic beams in plasma and implement it in a numerical algorithm. Here, the basic theory is introduced. We consider a general quasimonochromatic multicomponent wave in a weakly inhomogeneous linear medium with no sources. For any given dispersion operator that governs the wave field, we explicitly calculate the approximate operator that governs the wave envelope ψ to the second order in the geometrical-optics parameter. Then, we further simplify this envelope operator by assuming that the gradient of ψ transverse to the local group velocity is much larger than the corresponding parallel gradient. This leads to a parabolic differential equation for ψ (“quasioptical equation”) on the basis of the geometrical-optics polarization vectors. Scalar and mode-converting vector beams are described on the same footing. We also explain how to apply this model to electromagnetic waves in general. In the next papers of this series, we report successful quasioptical modeling of radio frequency wave beams in magnetized plasma based on this theory.

Describing the propagation of waves in inhomogeneous media is a classic problem with a long history.1–3 It is particularly important in fusion research, where quasistationary beams of electromagnetic (EM) radiation are commonly used for many purposes and need to be modeled with fidelity.4 Full-wave modeling, which involves solving complete Maxwell's equations, can be impractical at short (cm and mm) wavelengths, especially when multidimensional simulations with complex geometries are required, such as those of tokamak and stellarator plasmas. Hence, reduced methods have been widely used in practice. These methods are rooted in geometrical optics (GO)1 and include conventional ray tracing,5–8 complex ray tracing,9,10 beam tracing,11–14 and variations thereof.15 There are also other “quasioptical” models, such as in Refs. 16–19, which resolve the evolution of the beam transverse structure without adopting any particular ansatz for the intensity profile. Still, they assume that only one branch of the dispersion relation is excited in each given case,20 i.e., mode conversion does not occur.21 As a result, today's simulations of mode-converting beams mainly rely on full-wave codes19 and thus have to compromise the fidelity by reducing the number of dimensions resolved.8 

However, simulating mode conversion does not actually require the full-wave approach; in fact, it can be done within the quasioptical approach too if the latter is generalized properly. Developing the corresponding algorithms would facilitate not only fusion applications. For example, advanced quasioptical models could find use in optics and general relativity, where mode-converting beams are also possible and have been attracting much attention lately.22,23 Hence, it is potentially beneficial to approach the problem of quasioptical modeling as a general-physics problem, without restricting it to fusion applications.

Mode conversion can be described quite generally, particularly without restricting it to narrow regions in space, within “extended geometrical optics” (XGO), which was developed recently.24–28 XGO is a theory which calculates the leading-order correction U to the GO dispersion operator of a general vector wave and shows26 that this correction is analogous to (and a generalization of) the Stern–Gerlach Hamiltonian of a quantum spin-1/2 electron. Accordingly, U is responsible for two effects simultaneously: (i) it modifies the ray equations just like spin–orbital interactions affecting the electron motion and (ii) it also governs mode conversion, which appears as a direct analog of spin up–down transitions.29 An examination of the quasioptical algorithms such as those in Refs. 12 and 16 shows that they already involve calculations of terms similar to U. Hence, adding the mode-conversion capability to quasioptical codes should not be burdensome and is not expected to slow down the codes considerably. However, formulating the corresponding theory is easier to do using the abstract quantumlike formalism of XGO. One only needs to upgrade the existing XGO by adding diffraction, for which it also helps to introduce more general coordinates with curved metric.30 

In this series of papers (Papers I–III), we propose such an upgrade of XGO and apply it to numerical simulations. Our goal is to develop a framework which later could be applied to a broad variety of problems both in plasma physics and beyond. Within this framework to be presented, one can assume general dispersion, diffraction, and polarization effects, including mode conversion of not just two but many arbitrary resonant waves. Also importantly, our formulation below is not restricted to EM waves since we do not specify the dispersion operator in the governing wave equation.

Our series of papers is organized as follows. In Paper I, we introduce the basic theory of waves that diffract and mode-convert simultaneously. In Paper II31 and Paper III,32 we apply this theory to perform quasioptical modeling of radiofrequency-wave beams in magnetized plasma as an example. In particular, we consider applications to mode conversion caused by magnetic shear in edge plasma,29 which, for example, is a known problem33,34 in the Large Helical Device.35,36

In this first paper of our series, we consider an arbitrary quasimonochromatic multicomponent wave in a weakly inhomogeneous linear medium. Supposing that D̂ is some dispersion operator governing the wave dynamics, we simplify D̂ and obtain an approximate operator that governs the wave envelope ψ. Then, we derive a parabolic differential equation for ψ (“quasioptical equation”) by assuming that the gradient of the wave envelope transverse to the local group velocity is much larger than the corresponding parallel gradient. The resulting theory applies to both scalar and mode-converting vector beams. At the end of the paper, we also discuss how this model can be applied to EM waves in particular. However, readers who are mainly interested in simulations (as opposed to the general theory) are encouraged to proceed straight to Paper II, where our key equations are overviewed in a simplified form and without derivations.

This paper is organized as follows: In Sec. II, we introduce the general problem. In Sec. III, we formalize the concept of the envelope dispersion operator. In Sec. IV, we derive an approximation for the envelope dispersion operator for scalar waves, and we also derive its quasioptical approximation. In Sec. V, we extend this model to vector waves. In Sec. VI, we explain how to apply the resulting theory to EM waves in particular. In Sec. VII, we present our main conclusions. In  Appendix A, we summarize some of our notations. In  Appendix B, some auxiliary calculations are presented. Our paper also contains the supplementary material. There, we overview the Weyl calculus on a curved configuration space, which is used in this work.

Consider a wave propagating on an n-dimensional configuration space Mn with coordinates x{x0,x1,,xn1} and some general metric tensor g(x). For simplicity, assume that Mn is diffeomorphic to n, i.e., the n-dimensional Euclidean space or pseudo-Euclidean space with the same metric signature as Mn. (For more information on why the diffeomorphism with n is needed, see Sec. IV A 2.) Suppose that the wave field ΨΨ(x), which may have multiple components, is governed by a linear equation with no source terms,

D̂Ψ=0,
(1)

where D̂ is a differential or, more generally, integral dispersion operator. (For vector waves, D̂ is a matrix whose elements are operators; see Sec. V.) We shall assume that the GO parameter ϵ is small, namely,

ϵλ/L1
(2)

(the symbol denotes definitions), where λ is the characteristic wave period, or wavelength, and L is the least characteristic scale among those of the wave envelope and of the medium, including the metric.37 Below, we propose a systematic reduction of Eq. (1) using the smallness of ϵ and eventually obtain a quasioptical model based on this equation. The idea of quasioptical modeling will be formalized later (Secs. IV E and V F).

As a side note, we emphasize that the theory to be developed does not describe wave transformations near cutoffs, where the GO parameter (2) is not small. It is possible to waive this limitation, but formulating such generalized theory is left to future publications.

As the first step, let us introduce a unitary variable transformation

Ψ=Ûψ,Ûeiθ(x).
(3)

(Other Û may also be justified in some cases, e.g., for dealing with caustics or quasiperiodic media,38 but we shall not consider this possibility in the present work.) The phase θ, which we call the “reference phase,” serves as a gauge potential. It is a real function such that

k(x)θ(x)
(4)

is the wave vector identical or close to that predicted by the GO approximation. This implies that the envelope ψ and also k are slow functions [and that the wavelength substituted in Eq. (2) is λ2π/k]. Then, Eq. (1) becomes

D̂ψ=0,
(5)

where the “envelope dispersion operator” is D̂ÛD̂Û (the dagger denotes the adjoint, as usual) or, more explicitly,

D̂=eiθD̂eiθ.
(6)

The reference phase θ is treated as a prescribed function. As it will become clear later, knowing θ per se is not actually needed for our purposes; instead, it is k that matters. The latter can be calculated on some “reference rays” using the conventional ray equations,1

dxαdτ=Hkα,dkαdτ=Hxα
(7)

(τ is any parameter along the ray), which also leads to

H(x,k(x))=const.
(8)

Note that, in general, θ and k are defined uniquely only to the leading order, and so there exists some freedom in choosing reference rays and their Hamiltonian H. This means that more than one D̂ is possible. Still, envelope equations that (slightly) differ in the choice of k are equivalent in the sense that the total field Ψ that they describe is the same by construction. We shall discuss this in more detail in Secs. V E and V G.

Also note that our approach equally applies to stationary and nonstationary waves. In the case of a stationary wave, we assume that x is a coordinate in physical space (“spatial problem”) and the wave frequency ω serves as a constant parameter. In the case of a nonstationary wave, we assume that x is a coordinate in spacetime (“spacetime problem”), and then, ω is a part of k. In spacetime problems, we assume coordinates such that x0=ct, where c is the speed of light, t is the time, and the metric signature is (,+,+,); then, k0=ω/c (in the case of the Minkowski metric, this implies that k0=ω/c), and so ω=tθ, as usual. In other respects, spatial and spacetime problems are described on the same footing and will be distinguished only in Sec. VI.

Having defined this terminology, we shall now discuss how D̂ can be asymptotically expanded in ϵ for any D̂.

Our asymptotic theory of mode-converting beams is based on the well-known phase-space methods in quantum and classical wave theory. (For recent overviews, see Refs. 1 and 24. Specific key papers include, but are not limited to, Refs. 39–44. In particular, Ref. 43 is one of the first papers which explicitly discuss applications of the Weyl calculus to plasma problems; although also see Refs. 45–47.) In order to present our framework and notation in a self-contained manner, we restate some basics in Sec. IV A and the supplementary material. Also note that our calculations in Sec. IV B are similar in spirit to those in Ref. 43, except that we perform a higher-order expansion and assume a more general metric. We shall also comment throughout this paper on how our equations reproduce other known results in the corresponding limits.

1. Basic definitions

Until Sec. V, we shall assume that Ψ is a scalar function. Any given operator  acting on it maps Ψ to a new scalar function ÂΨ which can be expressed as follows:

(ÂΨ)(x)dnxg(x)A(x,x)Ψ(x).
(9)

Here, the integral is taken over n (and so are all integrals below, up to dimensions), g|detg|, and A is some kernel function that determines Â. Also, consider a family of all unitary operators Âu, which is a subset of all possible Â. For a given Ψ, all image functions ÂuΨ are mutually equivalent up to an isomorphism, and so their family {ÂuΨ} can be viewed as a single object, a “state vector” |Ψ, which belongs to a Hilbert space H1n with the inner product

Ψ|Φdnxg(x)Ψ*(x)Φ(x).
(10)

Then, Eq. (9) can be viewed as the “x representation” of Â, while the operator itself can be understood more generally as a transformation of |Ψ, i.e., of the whole family {ÂuΨ}. Using this invariant notation, one can formulate a machinery, called the Weyl calculus,48 which allows efficient asymptotic approximation of operators using the smallness of ϵ. Below, we overview the key theorems of the Weyl calculus which are used in our paper. Readers who are interested in details and proof of these theorems can find them in the supplementary material.

2. Coordinate and momentum operators

We start by defining the coordinate and momentum (wave-vector) operators,

x̂={x̂0,x̂1,,x̂n1},
(11)
p̂={p̂0,p̂1,,p̂n1}
(12)

such that the x representations of x̂μ and p̂μ are as follows:

x̂μΨ=xμΨ,
(13)
p̂μΨ=ig1/4μ(g1/4Ψ).
(14)

Here, μ/xμ,xμ and pμ are the corresponding eigenvalues, and the factors g±1/4 are introduced to keep p̂ self-adjoint under the inner product (10).49 (This would not be the case if Mn were not diffeomorphic to n.50) Since

p̂ν=iνiqν(x),qν14ν(lng),
(15)

and qν commutes with x̂μ, the usual commutation relation [x̂μ,p̂ν]=iδνμ is satisfied.

Let us consider the eigenvectors |x and |p of the coordinate and momentum operators, which are defined as

x̂|x=x|x,p̂|p=p|p.
(16)

Since the operators are self-adjoint, these eigenvectors can be chosen as mutually orthogonal, and we shall assume the following normalization:

x1|x2=G(x1,x2)δ(x1x2),
(17)
p1|p2=G¯(p1,p2)δ(p1p2).
(18)

Here, we introduced

G(x1,x2)[g(x1)g(x2)]1/4,
(19)
G¯(p1,p2)[g¯(p1)g¯(p2)]1/4.
(20)

The function g¯ can be chosen arbitrarily as long as it is kept positive. It plays a role similar to that of g in the Weyl calculus, but note that this is just a normalization factor, and we introduced it only to maintain the symmetry between x̂ and p̂. One can show then that our original function Ψ and the envelope ψ can be expressed through the corresponding state vectors as

Ψ(x)=x|Ψ,ψ(x)=x|ψ.
(21)

The p representations of these state vectors are introduced similarly as p|Ψ and p|ψ. One can also show that x1|Â|x2=A(x1,x2) and

x|p=p|x*=exp(ip·x),(2π)n/2[g(x)g¯(p)]1/4.
(22)

Here, p·xpμxμ, and summation over repeated indices is assumed from now on.

3. Wigner–Weyl transform

The set of all eigenvalues of the coordinate and momentum operators form a 2n-dimensional “phase space” z(x,p). For every given z, we introduce the so-called Wigner operatorΔ̂z, which is self-adjoint and defined as

Δ̂zdnsG(x,s)|xs/2x+s/2|eip·s,
(23)
G(x,s)[g(xs/2)g(x+s/2)]1/4.
(24)

(Note that G =1 if the x space is Euclidean or pseudo-Euclidean.) Using Δ̂z, we define the Wigner–Weyl transformWz:ÂA, which maps a given operator  on H1n to a function A (“Weyl symbol”) on the z space. Specifically, the Weyl symbol of a given operator  is A(x,p)tr(Δ̂zÂ) (“tr” stands for trace); i.e.,

A(x,p)dnsG(x,s)x+s/2|Â|xs/2eip·s.
(25)

As can be seen easily from this definition, if a given operator is self-adjoint, then its Weyl symbol is real. This also leads to the following corollary. Consider splitting a given  as Â=ÂH+iÂA, where the subscripts denote the Hermitian and anti-Hermitian parts,

ÂH12(Â+Â),ÂA12i(ÂÂ).
(26)

Both ÂH and ÂA (not to be confused with iÂA) are self-adjoint by definition. Thus, the corresponding Weyl images AH and AA are real.

We also define the inverse Wigner–Weyl transformW1:AÂ, which maps a given function A to the corresponding operator  via

Â=1(2π)ndnxdnpA(x,p)Δ̂z.
(27)

The direct and inverse transforms set the “Weyl correspondence” between operators and functions on the (x,p) space, ÂA(x,p). As can be checked by a direct calculation, for any function f, one has

f(x̂)f(x),f(p̂)f(p).
(28)

However, the Weyl symbols of operators that cannot be represented as f1(x̂)+f2(p̂) are generally more complicated. In particular, one can show that (see the supplementary material)

f(x̂)p̂αpαf(x)+i2αf(x),
(29)
p̂αf(x̂)pαf(x)i2αf(x),
(30)

and also

p̂αf(x̂)p̂βpαpβf(x)+14αβ2f(x)+i2pαβf(x)i2pβαf(x).
(31)

Overall, the Weyl symbol of an operator, that is, any given combination f(x̂,p̂) of x̂ and p̂, approaches f(x,p) in the GO limit, when [x̂μ,p̂ν] is negligible. However, in general, f(x̂,p̂) does not map simply to f(x,p).

Using the notation introduced in Sec. IV A, one can express Eq. (1) in the following invariant form:

D̂|Ψ=0.
(32)

Accordingly, Eq. (5) becomes

D̂|ψ=0,
(33)

where the invariant form of the envelope dispersion operator [Eq. (6)] is as follows:

D̂=eiθ(x̂)D̂eiθ(x̂).
(34)

We shall now approximate D̂ in three steps: (i) we map the right-hand side of Eq. (34) onto a function, or symbol, using the Wigner–Weyl transform; (ii) we approximate this symbol using the smallness of ϵ [Eq. (2)]; and (iii) we produce an operator out of the approximated symbol using the inverse Wigner–Weyl transform.

The first two steps can be done by expressing the Weyl symbol of D̂ as D(x,p)=eiθ(x)D(x,p)eiθ(x), where D is the Weyl symbol of D̂ and is the Moyal product (see the supplementary material). By expanding in the GO parameter, one can then obtain an approximate symbol of D̂ formally. But here we adopt a different approach, which is more transparent. We start by expressing D̂ through D explicitly,

D̂=1(2π)ndnxdnpdnsG(x,s)|xs/2eip·sD(x,p)x+s/2|,
(35)

where G is a metric factor given by Eq. (24). Using the fact that θ(x̂)|x±s/2=θ(x±s/2)|x±s/2, one can rewrite Eq. (34) as follows:

D̂=1(2π)ndnxdnpdnsG(x,s)ei[θ(x+s/2)θ(xs/2)p·s]|xs/2D(x,p)x+s/2|.
(36)

Consider a formal Taylor expansion of the reference phase θ in s,

θ(x±s/2)=θ(x)±sα2θ(x)xα+12sαsβ42θ(x)xαxβ±16sαsβsγ83θ(x)xαxβxγ+
(37)

This gives

θ(x+s/2)θ(xs/2)=sαθ(x)xα+sαsβsγ243θ(x)xαxβxγ+=k(x)·s+sαsβsγ242kγ(x)xαxβ+,
(38)

and so

ei[θ(x+s/2)θ(xs/2)p·s]=[1+i24sαsβsγ2kγ(x)xαxβ+]ei[k(x)p]·s=[1+1242kγ(x)xαxβ3pαpβpγ+]ei[k(x)p]·s.
(39)

This leads to the following expression for the envelope dispersion operator:

D̂=1(2π)ndnxdnpdnsG(x,s){[1+1242kγ(x)xαxβ3pαpβpγ+]ei[k(x)p]·s}|xs/2D(x,p)x+s/2|=1(2π)ndnxdnpdnsG(x,s)ei[k(x)p]·s|xs/2[D(x,p)1242kγ(x)xαxβ3D(x,p)pαpβpγ+]x+s/2|=1(2π)ndnxdnpdnsG(x,s)eip·s|xs/2D(x,k(x)+p)x+s/2|,
(40)

where we integrated by parts and introduced an “effective dispersion function,”

D(x,p)D(x,p)1242kγ(x)xαxβ3D(x,p)pαpβpγ+
(41)

The first term on the right-hand side of Eq. (41) is O(1) because at ϵ0, it becomes the GO dispersion function, which is assumed order-one.51 (At nonzero ϵ, the Weyl symbol D generally does not coincide with the dispersion function that governs the wave in a homogeneous medium; see Sec. IV A 3.) The second term can be estimated by assuming /x1/L and /p1/k, since we are interested in D(x,p) at p close to k. Then,

2kγxαxβ3DpαpβpγkL2Dk3ϵ2D,
(42)

meaning that the second term in Eq. (41) is roughly ϵ2 times smaller than the first one. This fact will be used below.

The inverse Wigner–Weyl transform will turn the coordinate p into the operator p̂. The latter will act on the wave envelope [Eq. (33)], which is considered slow in the coordinate representation, and so p̂|ψ=O(ϵ). In this sense, p=O(ϵ), and so by Taylor-expanding D in p, we are effectively expanding D̂ in ϵ. It is sufficient for our purposes to adopt the second-order expansion,

D̂1(2π)ndnxdnpdnsG(x,s)eip·s|xs/2[D(x)+pμVμ(x)+12pμpνΘμν(x)]x+s/2|,
(43)

where we introduced D(x)D(x,k(x)) and

Vμ(x)D(x,k(x))kμ,
(44)
Θμν(x)2D(x,k(x))kμkν.
(45)

By properties of the inverse Wigner–Weyl transform W1 (Sec. IV A 3), D̂ can be written as follows:

D̂W1[D(x)]+W1[pμVμ(x)]+12W1[pμpνΘμν(x)].
(46)

The inverse Wigner–Weyl transforms here can be calculated using Eqs. (28)–(31) and Θμν=Θνμ. This leads to

D̂D(x̂)+12[p̂μVμ(x̂)+Vμ(x̂)p̂μ]+12p̂μΘμν(x̂)p̂ν18(μν2Θμν)(x̂).
(47)

One can further simplify Eq. (47) as follows. Suppose that the medium, including the metric, is characterized by some parameters P. Within the accuracy of the theory developed in this paper, we shall neglect terms involving 2P and (P)(ψ) while generally retaining terms such as 2ψ (as to be explained Sec. IV E 2), where denotes a generic spatial derivative. Hence, we can ignore the difference between D and D; we can also omit μν2Θμν in Eq. (47), and so D̂ can be expressed as follows:

D̂D(x̂)+12[p̂μVμ(x̂)+Vμ(x̂)p̂μ]+12p̂μΘμν(x̂)p̂ν.
(48)

Also, Vμ and Θμν can be calculated through the zeroth-ϵ limit of D, denoted D0, which is just the dispersion function of the homogeneous medium,

Vμ(x)D0(x,k(x))kμ,Θμν(x)2D0(x,k(x))kμkν.

Importantly, D0(x,k) can depend only on k and on local parameters of the medium (but not on their derivatives). As a gradient of a scalar [Eq. (4)], k is a true covector; hence, D0(x,k) is a true scalar. This makes Vμ a vector and Θμν a tensor within the assumed accuracy.

As a side remark, note that vectors like Vμ, which are denoted with upper indices, belong to the space tangent to Mn. (Such vectors must not be confused with multicomponent fields denoted with Latin indices in Sec. V.) Similarly, covectors like kμ, which are denoted with lower indices, belong to the space cotangent to Mn. For any given X, one does not have to distinguish Xμ and Xμ only if the metric is Euclidean. In the case of a general metric gμν, one has

Xμ=gμνXν,Xμ=gμνXν,
(49)

where the matrix gμν is the inverse of gμν. Similar rules apply to manipulating tensors, as usual.

Let us derive the x representation of Eq. (48), which we need to obtain an explicit form of Eq. (5). From the x representations of x̂ and p̂ given by Eqs. (13) and (14), terms like p̂f(x̂) (for any f) must be interpreted as

[p̂μf(x̂)]ψ=ig1/4μ[g1/4f(x)ψ],
(50)

where g and ψ are also functions of x. Then,

D̂ψDψiVμμψi2Vμ;μψ12g1/4μ[Θμνν(g1/4ψ)],
(51)

where all terms are functions of x and ;μ denotes the covariant derivative with respect to xμ, so Xμ;μ (for any given vector Xμ) is the divergence, namely,

Xμ;μ1gxμ(gXμ).
(52)

Except for , all terms in Eq. (51) are covariant (i.e., invariant with respect to coordinate transformations) within the accuracy of our theory. Hence, the symbol D is also covariant within the accuracy of our theory [this includes not only the zeroth-order dispersion function D0 but also the leading-order correction DD0=O(P) that may appear for inhomogeneous media], even though this may not be obvious directly from the definition of the Weyl transform (Sec. IV A). To emphasize this fact, and also to ensure consistency of notation with the vector-wave theory to be discussed in Secs. V and VI, we denote this covariant function as D, and we also adopt the notation

DHReD,DAImD.
(53)

Then, the approximate form of the envelope equation (5) can be summarized as follows:

(D̂H+iD̂A)ψ=0.
(54)

Here, D̂H and D̂A are the Hermitian and anti-Hermitian parts of the approximated D̂, respectively, namely,

D̂Hψ=DHψiVμψ,μi2Vμ;μψ12g1/4[DH|μν(g1/4ψ),ν],μ,
(55)
D̂Aψ=DAψiVAμψ,μi2VAμ;μψ12g1/4[DA|μν(g1/4ψ),ν],μ
(56)

[from now on, we do not emphasize the approximate nature of Eq. (51), and so ≈ will be replaced with =], and

VμVHμ=DH|μ,
(57)

where we assume a standard notation

f,μfxμ,f,μν2fxμxν.
(58)

Note that ,μ is the “full” derivative in the sense that, for any given f(x,k(x)), it applies to both the first and the second argument of f; namely, the chain rule leads to

f,μ(x,k(x))=f|μ+f|νkν|μ.
(59)

The true partial derivatives are introduced as follows:

f|μf(x,k)xμ,f|μν2f(x,k)xμxν,
(60)
f|μf(x,k)kμ,f|μν2f(x,k)kμkν.
(61)

The derivatives ,μ and |μ are equivalent for functions that depend only on x. In particular, kννθ(x) satisfies

kν|μ=kν,μ=kμ,ν=θ,μν=θ,νμ.
(62)

For the envelope approximation to hold, i.e., for ψ(x) to remain a slow function, k(x) must be chosen such that D(x,k(x))ψO(ϵ). [The other terms in Eq. (5) are automatically small on the account of Eq. (2).] Let us adopt the usual GO ordering52 

DH=O(1),DA=O(ϵ),
(63)

or more rigorously, DAO(ϵ); i.e., DA much smaller than O(ϵ) is also allowed.53 Then, one can define k(x) such that

DH(x,k(x))=0.
(64)

This can be achieved by calculating k as discussed in Sec. III with the ray Hamiltonian H=DH and the initial condition [at any chosen x(0) and k(0)] such that

DH(x(0),k(0))=0.
(65)

To the first order in ϵ, Eq. (5) is

Vμψ,μ+12Vμ;μψ=DAψ,
(66)

and as a corollary,

Jμ;μ=2DA|ψ|2,JμVμ|ψ|2.
(67)

At zero DA, the dynamics is conservative, and Eq. (67) reflects conservation of the wave action, or quanta.54 Accordingly, DA determines the dissipation rate, Jμ can be identified (at least up to a constant factor) as the action flux density, and Vμ is proportional to the group velocity (in space or in spacetime, depending on the problem). Equations (64)–(67) coincide with the equations of the traditional GO theory.1,54 Below, we extend them by retaining the second-order terms neglected in Eq. (66).

1. Ray-based coordinates

Let us start by introducing ray-based coordinates as follows.55 Suppose multiple rays launched with different x(0) within the beam. Their trajectories can be found using the ray equations as discussed in Sec. IV D; this determines the path ζ along each ray as a function of the initial coordinate. We treat ζ as the longitudinal coordinate along the wave beam, and each of its isosurfaces is considered as the transverse space Mn1(ζ). The coordinates on this space, ϱ{ϱ1,ϱ2,,ϱn1}, can be introduced arbitrarily (yet they will be specified below). Then, we define n – 1 independent vector fields eσ via

eσx/ϱσ
(68)

and another vector e0 such that it is linearly independent of all eσ. Hence, any dx can be decomposed as follows:

dx=e0dζ+eσdϱσ,
(69)

and summation over repeated indices σ (and σ̃) is henceforth assumed from 1 to n – 1.

Let us also construct the dual basis {eμ}. (We treat vectors and covectors on the same footing; namely, X·Y=XμYμ=XμYμ and X=Y equally means Xμ=Yμ and Xμ=Yμ.) By definition,

eμ·eν=δνμ,
(70)

and so we adopt e0ζ. Then, the general form of e0 is

e0=±α2e0+β,
(71)

where the first sign is determined by the metric signature and α|e0·e0|1/2,βeσβσ, and βσ are arbitrary coefficients. This leads to the following metric representation in the coordinates {ζ,ϱ}:

g=(±α2+βThβ(hβ)Thβh).
(72)

Here, h is the matrix with elements hσσ̃eσ·eσ̃ and T denotes the transposition. By using a known theorem for the determinant of a block matrix,56 one obtains

g=α2h,h|deth|.
(73)

Let us also introduce the transverse projection of dϱ,

dϱ(1α2e0e0)dϱ,dϱσ=dϱσ+βσdζ,
(74)

where 1 is a unit matrix and e0e0 is a dyad formed out of e0. Then, one obtains from Eq. (69) that

dx·dx=±α2(dζ)2+hσσ̃dϱσdϱσ̃,
(75)

and so hσσ̃ serves as the transverse metric. Below, we assume β=0, and so ϱσ and ϱσ do not need to be distinguished, and we can define the inner product on Mn1 as follows:

ψ|ϕdn1ϱh(ζ,ϱ)ψ*(ζ,ϱ)ϕ(ζ,ϱ).
(76)

We also choose transverse coordinates specifically such that ϱ=const. In other words, the coordinate ϱ of a given point in space is the initial location of the ray that arrives at this point from the initial transverse surface Mn1(0). Then, the transverse group velocity is zero,

Vσeσ·dxdτ=dϱσdτ=0,
(77)

where we used Eqs. (69) and (70). This simplifies our equations below. A different definition of ϱ is assumed in Papers II and III, leading to straightforward modifications of the equations, which are not discussed below.

2. Quasioptical equation

Suppose a wave beam such that its longitudinal scale L||, which is defined via ψ,ζψ/L||, is much larger than its perpendicular scale L, which is defined via ψ,σψ/L and may or may not be the beam global width. (We assume the notation ψ,ζψ/ζ and ψ,σψ/ϱσ.) We introduce two GO parameters,

ϵ||λ/L||,ϵλ/L,ϵ||ϵ
(78)

[so the original parameter (2) is ϵ=ϵ], and we shall neglect terms smaller than O(ϵ||ϵ) from now on. Also, we require that either α is chosen as independent of ϱ, or, more generally, the relative variation of α across the wave beam is small enough such that α,σ is negligible. We also assume

(DH)|μO(ϵ||),gαβ,μO(ϵ||),
(79)

where the derivatives can be taken over the longitudinal or transverse coordinates. (Having gαβ,μ1/L|| is indeed common in typical applications.30)

Under these assumptions, Eq. (55) can be approximated as follows [also assuming Eq. (64)]:

D̂Hψ=iVψ,ζi2g(gV),ζψ+Ĝψ.
(80)

Here, VV0, and we used Vμψ,μ=Vψ,ζ due to Eq. (77). The operator Ĝ is self-adjoint under the inner product (76) and given by

Ĝψ12h1/4[DH|σσ̃(h1/4ψ),σ̃],σ.
(81)

(We kept only the transverse derivatives in this term because the longitudinal ones are beyond the assumed accuracy.) Similarly, Eq. (56) becomes

D̂Aψ=DAiVAσψ,σi2h(hVAσ),σψ,
(82)

where the higher-order terms are neglected because DA is assumed small [Eq. (63)]. Although the last term in Eq. (82) is negligible within the assumed accuracy, it is retained anyway to make D̂A (not to be confused with iD̂A) exactly self-adjoint under the inner product (76). Then, Eq. (5) becomes

iVψ,ζ+i2g(gV),ζψĜψ=iD̂Aψ,
(83)

and, as a corollary,

1αddζψ|αV|ψ=2ψ|D̂A|ψ,
(84)

where we used the fact that Ĝ and D̂A are self-adjoint. In particular, for zero DA, Eq. (84) predicts conservation of the wave-action flux through the beam cross section,

ψ|αV|ψ=const.
(85)

Equation (84) indicates that D̂A cannot be much larger than /ζ, and so the correct scaling to assume for the dissipation term is DAO(ϵ||). Then, Eq. (83) leads to the “quasioptical scaling”

ϵ||ϵ2.
(86)

Hence, the difference between D̂A and DA in Eq. (82) is O(ϵ3), and so it is beyond the accuracy of our theory, which is O(ϵ2). For this reason, we henceforth adopt57 

D̂A=DA,
(87)

which will also simplify our notation. [Also, as a reminder, μgαβO(ϵ||)ϵ2, and the parameters of the medium vary similarly.] This leads to

iVψ,ζ+i2g(gV),ζψĜψ=iDAψ.
(88)

Equation (88) is the general quasioptical equation for a scalar-wave beam in an inhomogeneous medium. Since it is a parabolic equation (it contains only the first-order derivative with respect to ζ), Eq. (88) is much easier to solve than Eq. (5) with the original expressions for D̂H and D̂A given by Eqs. (55) and (56).

3. Simplified equations

A typical metric of interest for quasioptical modeling has the form gμν=δμν+Rμνσϱσ with Rμνσ=O(L||1).30 This satisfies the assumed scaling (79) and corresponds to lng=O(L/L||). Then,

(lng),ζL/L||2=O(ϵ||3/2),
(89)

which is negligible. [In contrast, (lng),σ=O(ϵ||).] The derivatives of g in Eq. (81) are negligible too. Hence, the metric factors in the quasioptical equation (88) can be replaced with unity. (However, the effect of the metric on the dispersion matrix may remain important in this case; see Sec. VI.) Then, Eq. (88) becomes

iVψ,ζ+i2V,ζψ+12(DH|σσ̃ψ,σ̃),σ=iDAψ.
(90)

Using the variable transformation ϕVψ, this equation can be further simplified as follows:

iϕ,ζ+12(Σσσ̃ϕ,σ̃),σ=iϒϕ,
(91)

which is just a dissipative Schrödinger equation with

Σσσ̃DH|σσ̃/V,ϒDA/V.
(92)

[In Eq. (91), we used that V,σϕ,σ̃ is negligible compared to Vϕ,σ̃σ=O(ϵ).] Also, Eq. (84) becomes

ddζϕ|ϕ=2ϕ|ʋ|ϕ.
(93)

Note that Σ can also be expressed alternatively as follows. Suppose one finds the group velocity and constructs a ray-based metric (as described in Sec. IV E 1) in the vicinity of a given x. Then, Eq. (64) can be considered as an equation for the local k0 as a function of the transverse wave-vector components kσ (and x) in this prescribed metric; i.e.,

DH(x,kσ,k0(x,kσ))=0.
(94)

By differentiating this with respect to kσ, we obtain k0|σ=DH|σ/DH|ζ. By differentiating the latter formula once again, now with respect to kσ̃, we also obtain

k0|σσ̃=1DH|ζ(DH|ζζk0|σk0|σ̃+DH|ζσ̃k0|σ+DH|ζσk0|σ̃+DH|σσ̃).
(95)

Recall that DH|σ=Vσ and Vσ=0 [Eq. (77)]. Then, k0|σ=0, and by comparing Eq. (95) with Eq. (92) and using Eq. (57), one finds that k0|σσ̃=Σσσ̃. Thus, Eq. (91) can be rewritten as

iϕ,ζ12(k0|σσ̃ϕ,σ̃),σ=iϒϕ.
(96)

In the case of a spacetime problem, k0|σσ̃ can also be expressed through the wave frequency in the laboratory frame. Then, the familiar Schrödinger equation for the wave envelope58 can be reproduced. We do not discuss it further because this subject is not essential to our paper.

Now, let us generalize the above results to vector waves. Suppose an m-component wave field ΨΨ(x), so

Ψ=(Ψ1Ψ2Ψm),|Ψ=(|Ψ1|Ψ2|Ψm).
(97)

Here, |Ψ is a vector on some Hilbert space Hmn for which we assume the following general inner product:

Ψ|Φmdnxg(x)γab(x)Ψa*(x)Φb(x).
(98)

Here, the Latin indices that characterize the field components span from 1 to m. (This is in contrast to the Greek indices introduced earlier, which characterize the components of x and span from 0 to n – 1.) The matrix γ(x) can be any m × m symmetric matrix. It can be considered as an additional metric. This metric does not have to be the same as g, as seen already from the fact that m and n do not have to be the same. (For example, Ref. 27 describes a six-dimensional field on a three-dimensional spacetime; also see Sec. VI B.) This means that the space to which Ψ(x) belongs is not necessarily the space tangent to Mn. With that said, having γ=g is possible as a special case; see Secs. VI A and VI C.

Using γ as a metric, we introduce the standard rules for manipulating the Latin indices,

ΨaγabΨb,Ψa=γabΨb,
(99)

where γab are the elements of γ1. In particular, in the x representation, one has

Ψ(x)Ψ|x=(Ψ1*,Ψ2*,,Ψm*),
(100)

and we shall also use the following local dot product:

Ψ·ΦΨ(x)Φ(x)=Ψa*(x)Φa(x).
(101)

Under this dot product, a matrix A with mixed-index elements Aab is self-adjoint [(AΨ)·Φ=Ψ·(AΦ)] if the matrix with the corresponding lower-index elements, Aab, is Hermitian. The difference between Hermitian and self-adjoint matrices can be ignored if the metric γ is Euclidean or pseudo-Euclidean.

Any operator  on Hmn can be understood as an m × m matrix with elements Âab which are operators on H1n. For any given Âab, we introduce ¯abγacÂcb, which is also an operator on H1n. Correspondingly, two types of Weyl images can be defined, namely, Wz[Âab] and Wz[¯ab]. Below, we assume the notation

A¯abWz[¯ab],AabγacA¯cb.
(102)

Also, A will be an index-free notation of the matrix with elements Aab, and A¯ will be an index-free notation of the matrix with elements A¯ab. Importantly, Aab should not be confused with Wz[Âab] because the Weyl symbols do not transform as tensors,

Wz[Âab]γacWz[¯cb],
(103)

except for those which are independent of p. Symbols Wz[Âab] will not be used below, and so no special notation is introduced for them.

Note that by Eq. (98), the following equality holds for any Φ,Ψ, and  on Hmn:

Ψ|ÂΦm=dnxgγabΨa*(ÂbcΦc)=dnxgΨa*(¯abΦb)=Ψa|¯abΦb,
(104)

where we invoked the definition of the inner product on H1n [Eq. (10)]. By the definition of the adjoint operator Â,

Ψ|ÂΦm=ÂΨ|Φm=dnxgγcb[(Â)caΨa]*Φb=dnxg[(¯)baΨa]*Φb.
(105)

In order to express  through Â, let us represent ¯ab in terms of the corresponding Weyl image A¯ab and the Wigner operator Δ̂z on H1n, with z(x,p) (Sec. IV A 3). Then, Eq. (105) leads to

Ψ|ÂΦm=d2nz(2π)nA¯ab(z)Ψa|Δ̂zΦb=d2nz(2π)nA¯ab*(z)Δ̂zΨa|Φb,
(106)

where we used that Δ̂z is self-adjoint on H1n. By comparing this with Eq. (105), one finds that, on one hand,

(¯)ba=d2nz(2π)nA¯ab*(z)Δ̂z.
(107)

On the other hand, by Eq. (27), for the inverse Wigner–Weyl transform, one has

(¯)ba=d2nz(2π)nWz[(¯)ba]Δ̂z.
(108)

Hence, the Weyl image of  is simply the conjugate transpose of the Weyl image of  in the sense that

Wz[(¯)ab]=A¯ba*(z).
(109)

In other words, for operators with both indices lowered, the operations Wz and commute.

Equation (109) implies that, if  is self-adjoint on Hmn, then A¯ is a Hermitian matrix, and thus, the corresponding matrix A is self-adjoint (Sec. V A). In particular, this means that the Hermitian and anti-Hermitian parts [Eq. (26)] of an operator are determined by the Hermitian and anti-Hermitian parts of its lower-index Weyl symbol, respectively. We summarize this as follows:

(AH)ab=12γac(A¯cb+A¯bc*),
(110)
(AA)ab=12iγac(A¯cbA¯bc*).
(111)

For lower-index matrices, the Hermitian and anti-Hermitian parts are defined as usual,

(A¯H)ab=12(A¯ab+A¯ba*),
(112)
(A¯A)ab=12i(A¯abA¯ba*).
(113)

Now, let us consider the dispersion operator D̂ in particular. According to the above definition, it is viewed as an m×m matrix with elements D̂ab. By lowering the index in the envelope equation (1), one can write this equation as follows:

D̂¯abΨb=0,a=1,2,m,
(114)

where D̂¯abγacD̂cb. Equivalently, this can be written as the envelope equation,

D̂¯abψb=0,D̂¯abeiθD̂¯abeiθ.
(115)

Then, the approximate operators D̂¯ab can be expressed just like D̂ for scalar waves (Sec. IV B),

D¯̂abD¯ab+12[p̂μ(V¯μ)ab+(V¯μ)abp̂μ]+12p̂μ(Θ¯μν)abp̂ν.
(116)

Here, the coefficients are x-dependent matrices; specifically, D¯ab(x)D¯ab(x,k(x)), and

[V¯μ(x)]abD¯ab(x,k(x))kμ,
(117)
[Θ¯μν(x)]ab2D¯ab(x,k(x))kμkν.
(118)

One can also multiply Eq. (115) by γ1 to raise the first index. Then, one obtains

D̂abψb=0,D̂abγacD¯̂cb.
(119)

Note that if D̂ is self-adjoint, then D¯ab is Hermitian; then, D̂ab is self-adjoint too.

Unlike in scalar waves (Sec. IV C), D¯ab(x,k) is not a covariant object, i.e., in this case, not a true tensor. (For example, see Sec. VI.) It is convenient to split it as

D¯ab(x,k)=D¯ab(x,k)+C¯ab(x,k),
(120)

where D¯is a true tensor and C¯=O(P) is a small remainder. [Here, P means the same as in Sec. IV C, and so C=O(ϵ) under the GO ordering and C=O(ϵ) under the quasioptical ordering.] The splitting of Eq. (120) is not unique and is a matter of convenience; for example, one can define D¯(x,k) simply as the zeroth-order dispersion tensor. Hence, we can redefine

[V¯μ(x)]abD¯ab(x,k(x))kμ,
(121)
[Θμν(x)]ab2D¯ab(x,k(x))kμkν,
(122)

which is equivalent to the above definitions within the accuracy of our theory. Then, assuming the same orderings as in the scalar-wave case (also see below), the following approximation is enough both for the first-order theory and for the quasioptical theory:

D̂DH+D̂,
(123)
D̂=CH+iDA+D̂H1+D̂H2.
(124)

Here, DD(x) and D=γ1D¯, and so Dab(x)=γacD¯cb(x); similar conventions are assumed for C and D. Accordingly, the matrices DH and DA are self-adjoint and so are the operators D̂H1 and D̂H2, which are given by

D̂H112γ1(p̂μV¯Hμ+V¯Hμp̂μ),
(125)
D̂H212γ1p̂μΘ¯Hμνp̂ν.
(126)

Using [γ1,p̂μ]=i(γ1),μ and (γ1γ),μ=0 to obtain

(γ1),μγ=γ1γ,μμ,
(127)

one can also express these as

D̂H1=12[p̂μVHμ(x)+VHμ(x)p̂μiμVHμ],
(128)
D̂H212p̂μΘHμν(x)p̂ν.
(129)

(The term μ must be kept in D̂H1, but a similar term in D̂H2 can be neglected.) Also,

VHμ(x)γ1V¯Hμ(x),ΘHμν(x)γ1Θ¯Hμν(x),
(130)

and since γ is k-independent, this leads to

[VHμ(x)]ab(DH)ab(x,k(x))kμ,
(131)
[ΘHμν(x)]ab2(DH)ab(x,k(x))kμkν.
(132)

If D̂H2 is neglected, the envelope equation (119) becomes a Dirac-type equation similar to those considered in the context of XGO24–28 and also, for instance, in Refs. 59 and 60. We shall also discuss this limit in Sec. V E.

1. Basis from the eigenvectors of DH

Let us assume the GO ordering as in Eq. (63), except that DH and DA are now matrices. Then, D̂=O(ϵ), and so the envelope equation [Eq. (5)] acquires the form

DH(x)ψ=O(ϵ).
(133)

Hence, it is convenient to express ψ through the eigenvectors of DH(x). Let us denote these eigenvectors as ηs(x) and the corresponding eigenvalues as Λs(x), and so

DHηs=Λsηs.
(134)

[Note that ηs(x)=ηs(x,k(x)), where ηs(x,p) are the eigenvectors of DH(x,p). Likewise, Λs(x)=Λs(x,k(x)), where Λs(x,p) are the eigenvalues of DH(x,p).] Since DH is self-adjoint, it has m eigenvectors ηs which form a complete basis {ηs}. Let us also introduce the corresponding dual basis {ηs} as usual, i.e., such that

ηs·ηs=δss.
(135)

Then, we can represent ψ as

ψ=ηsas,as=ηs·ψ(ηs)a*ψa,
(136)

where as serves as the components of ψ in the basis {ηs}. [Remember that the dot product (101) includes conjugation.] Since DH is self-adjoint, it is always possible to make the basis {ηs} orthonormal, and this choice is assumed below. Then, ηs·ηs=δss; thus, ηs=ηs, i.e.,

(ηs)a=(ηs)a=γab(ηs)b.
(137)

Also, for any two fields ψ=ηsas and ϕ=ηsbs, the inner product (98) can be written as follows:

ψ|ϕm=dnxg(x)as*(x)bs(x),
(138)

where asδssas. The matrix δ with elements δss serves as a Euclidean metric for manipulating the mode indices s and s. Those should not to be confused with the coordinate indices denoted with Greek letters and also with other Latin indices that are manipulated by the metric γ [Eq. (99)].

2. Amplitude vectors

The physical meaning of the expansion coefficients as is understood as follows. Consider multiplying Eq. (133) by ηs from the left. This gives Λsas=O(ϵ), where no summation over s is assumed. This shows that, for a given s, there are two possibilities: either as is small or Λs is small. In the first case, the polarization ηs does not correspond to a propagating wave mode per se; the small nonzero projection of ψ on ηs is only due to the fact that the wave field is not strictly sinusoidal in inhomogeneous medium. We call such “modes” passive. In the second case, the local k(x) approximately satisfies the local dispersion relation Λs(x,k(x))0. Then, as can be understood as the local scalar amplitude of an actual GO mode so that as=O(1) is allowed. We call such modes active, and mode conversion occurs when more than one active mode exists. In other words, for a given boundary or initial conditions, active modes are those that are excited resonantly, while passive modes are those that are nonresonant and, thus, adiabatically isolated. (However, this does not mean that the passive-mode amplitudes are simply negligible; see below.)

Assuming that there are N1 active modes, we shall order them such that they correspond to s=1,2,,N and the remaining N¯=mN modes with s=(N+1),(N+2),,m are passive. Let us also adopt the notation

a¯s=as+N,η¯s=ηs+N,Λ¯s=Λs+N
(139)

(s=1,2,,N¯) for the passive-mode amplitudes, polarizations, and eigenvalues, respectively. Then, it is convenient to introduce the “amplitude vectors”

a(a1aN),a¯(a¯1a¯N¯)
(140)

and the corresponding row vectors that are dual to the amplitude vectors under the Euclidean complex dot product, namely,

a=(a1*,a2*,,aN*),a¯=(a¯1*,a¯2*,,a¯N¯*),
(141)
asδssas,a¯sδssa¯s.
(142)

Below, we seek to derive an approximate envelope equation in terms of these amplitude vectors.

3. Polarization matrices

Before we proceed, let us introduce the following notation. First, consider the “polarization matrices”

Ξ=(η1,η2,,ηN),Ξ¯=(η¯1,η¯2,,η¯N¯).
(143)

These are nonsquare matrices that have active- and passive-mode polarizations as their columns, namely,

Ξas=(ηs)a,Ξ¯as=(η¯s)a.
(144)

Using these, Eq. (136) can be rewritten compactly as

ψ=Ξa+Ξ¯a¯.
(145)

Also, consider the auxiliary polarization matrices

Ξ+(η1*ηN*),Ξ¯+(η¯1*η¯N¯*),
(146)

which have ηs* as their rows,

(Ξ+)sa=(ηs)a*,(Ξ¯+)sa=(η¯s)a*.
(147)

Since

(ψ)a=(ηs)a*as*+(η¯s)a*a¯s*=(ηs)a*as*+(η¯s)a*a¯s*,
(148)

one obtains

ψ=aΞ++a¯Ξ¯+.
(149)

However, note that in general, + does not mean to represent the adjoint in the common sense [cf. Eq. (154)].

Next, notice that

(Ξ+Ξ)ss=(ηs)a*(ηs)a=ηs·ηs=δss,
(150)

where the indices s and s span from 1 to N. Analogous formulas apply to Ξ¯. Then,

(Ξ¯+Ξ)ss=(η¯s)a*(ηs)a=η¯s·ηs=0,
(151)

and similarly for Ξ+Ξ¯. In other words, one has

Ξ+Ξ=1,Ξ¯+Ξ¯=1,Ξ¯+Ξ=0,Ξ+Ξ¯=0,
(152)

where 1 is the unit square matrix and 0 is the zero square matrix. (The dimensions of 1 and 0 are different in different equations.) Hence, if one multiplies Eq. (145) by Ξ+ and, separately, by Ξ¯+, one obtains concise formulas for the amplitude vectors a and a¯,

a=Ξ+ψ,a¯=Ξ¯+ψ.
(153)

Another property that we shall use later on is

(Ξ+)sa=δss(ηs)a*=δss(ηs)b*γab=(δ1ΞγH)sa,

and similarly for Ξ¯+. Here, H denotes the conjugate transpose (ΞHΞT*) and δ1 is the Kronecker matrix with upper-index elements δss. Hence,

Ξ+=δ1ΞHγ,Ξ¯+=δ1Ξ¯Hγ.
(154)

The factor δ1 can be omitted if one ignores the difference between upper and lower indices s and s that refer to the mode number. If γ is Euclidean, one can also ignore the difference between upper and lower coordinate indices; then, Ξ+=ΞH.

As a side remark, note that Ξ and Ξ¯ are generally nonsquare, and so ΠΞΞ+ and Π¯Ξ¯Ξ¯+ are not unit matrices but rather projectors. Indeed, due to Eqs. (152), one has Π2=Π and Π¯2=Π¯. Also, by applying Π and Π¯ to Eq. (145), one obtains

Ξa=Πψ,Ξ¯a=Π¯ψ.
(155)

Thus, Πψ is the projection of ψ on the active-mode space and Π¯ψ is the projection of ψ on the passive-mode space.

4. Equation for the active modes

Using the eigendecomposition theorem and Eq. (152), one obtains

DH=ηsΛsηs=ΞΛΞ++Ξ¯Λ¯Ξ¯+,
(156)

where Λ and Λ¯ are the diagonal eigenvalue matrices,

Λdiag{Λ1,Λ2,,ΛN}=O(ϵ),
(157)
Λ¯diag{Λ¯1,Λ¯2,,Λ¯N¯}=O(1).
(158)

These and Eq. (152) also lead to

DHΞ=ΞΛ=O(ϵ),DHΞ¯=Ξ¯Λ¯=O(1).
(159)

Hence, the envelope equation (5) can be written as

0=DHΞa+DHΞ¯a¯+D̂Ξa+D̂Ξ¯a¯=ΞΛa+Ξ¯Λ¯a¯+D̂Ξa+D̂Ξ¯a¯,
(160)

where we used Eq. (159). Let us multiply this by Ξ+ and, separately, by Ξ¯+. Then, due to Eq. (152), one obtains

Λa+Ξ+D̂Ξa+Ξ+D̂Ξ¯a¯=0,
(161)
Λ¯a¯+Ξ¯+D̂Ξa+Ξ¯+D̂Ξ¯a¯=0.
(162)

Let us also use Eq. (162) to express a¯ through a and substitute the result into Eq. (161). Since both a¯ and D̂ are of order ϵ, it is sufficient to solve Eq. (162) for a¯ approximately to the leading order,

a¯Λ¯1Ξ¯+D̂Ξa.
(163)

Then, Eq. (161) becomes an equation for just the N-dimensional active-mode amplitude vector,

(Λ+Ξ+D̂ΞΞ+D̂Ξ¯Λ¯1Ξ¯+D̂Ξ)a=0.
(164)

As a reminder, this equation is valid up to O(ϵ2).

Upon substituting Eqs. (124)(126) into Eq. (164), we obtain the following equation to lowest (first) order in ϵ:

(Λ+iΓ+K̂)a=0,
(165)

where we introduced

ΓΞ+DAΞ,
(166)
K̂Ξ+(D̂H1+CH)Ξ.
(167)

As shown in Appendix B 1,

K̂a=iVμa,μi2Vμ;μaUa,
(168)
VμΞ+VHμΞΛ|μ,
(169)
U=δ1(Ξ,μHV¯HμΞ)AΞ+CHΞ,
(170)

where Ξ is considered as a function of x. [As a reminder, ΞHΞT* is the conjugate transpose of Ξ; V¯Hμ is given by Eq. (121); and δ1 is a unit matrix which only raises the mode index.] Alternatively, Ξ can be considered as a function of (x,k). Then, as shown in Appendix B 2,

Uδ1(Λ|μΞHγΞ|μΛ|μΞHγΞ|μ+ΞH|μD¯HΞ|μ)AΞ+CHΞ,
(171)

where the partial derivatives |μ and |μ are defined in Sec. IV C. The term U is the Stern–Gerlach Hamiltonian mentioned in the introduction (Sec. I), except here it is generalized to an arbitrary metric. This term causes polarization-driven bending of the ray trajectories, which is missed in traditional GO; also, it causes mode conversion, if more than one active mode is present.24–28 These effects are discussed in further detail in Sec. V G.

More explicitly, Eq. (165) can be written as

ΛaiVμa,μi2Vμ;μa=(UiΓ)a.
(172)

This generalizes the XGO equation derived in Refs. 24–29 to dissipative waves and curved coordinates. Since the vector a belongs to the space where the metric is Euclidean by definition (Sec. V D 2), the elements of this vector are covariant, i.e., invariant with respect to coordinate transformations in the physical space. The same applies to the eigenvalues comprising Λ; thus, the whole left-hand side of Eq. (172) is covariant too. This means that the operator on the right-hand side is also covariant, which includes its Hermitian and anti-Hermitian parts separately. Hence, U and Γ are covariant.

Also, note that Eq. (172) is similar to Eq. (66) for scalar waves and has a similar corollary,

Jμ;μ=2aΓa,JμaVμa.
(173)

Using Eqs. (100), (130), (145), (149), (155), and (169), one obtains

Jμ=aΞ+VHμΞa=(Πψ)VHμ(Πψ)ψVHμψ=ψ*V¯Hμψ,
(174)

and so Eq. (173) can be viewed as a generalization of Eq. (67). At zero Γ, the dynamics is conservative, and Eq. (173) reflects conservation of the wave action or quanta.54 Accordingly, Γ determines the dissipation rate, Jμ can be identified (up to a constant factor) as the action flux density summed over all active modes (for example, see Sec. VI B), and the elements of the diagonal matrix Λ|μ are proportional to the group velocities of the corresponding active modes. We shall call them the group velocities (without “proportional to”) for brevity. We also emphasize that this model applies even when the group velocities of the active modes are very different, unlike the quasioptical model discussed in Sec. V F.

For scalar waves studied in Sec. IV, we had Λ=DH and we defined k(x) such that this term be zero [Eq. (64)]. Now, Λ is a diagonal matrix with N1 nonzero elements, and so it cannot be zeroed entirely by imposing just one scalar constraint on k(x). Hence, there can be more than one natural way to define k(x). One esthetically pleasing and convenient29 option is to require that Λ or ΛU be traceless. This amounts to choosing the reference-ray Hamiltonian in Sec. III as H=trΛ/N or H=tr(ΛU)/N, respectively. [Arbitrary constant factors can be introduced instead of N, for that only redefines τ in Eq. (7).] Another option is to adopt Λs=0 for some single sN, since all active modes have close wave vectors anyway; then, H=Λs. As mentioned in Sec. III, all such choices of k(x) are equally justified. Although they lead to slightly different envelope equations, those equations are equivalent in the sense that they all describe the same total field by construction. (One might call this a gauge freedom.) However, for the case N =1, choosing H=ΛU is preferable, as discussed in Sec. V G 2.

1. Quasioptical equation

Suppose that a wave propagates largely as a single beam. This implies that all active modes have group velocities close to their average group velocity Vavr, which can be defined via NVavrμtrΛ|μ. Then,

Vμ=Vavrμ1+ΔVμ,ΔVμVμVavrμ1,
(175)

where ΔVμ are matrices with eigenvalues much smaller than Vavr. Let us assume a ray-based coordinate system aligned with Vavr with the inner product on the transverse space Mn1 defined similar to Eq. (76) [cf. also Eq. (138)], namely,

a|bNdn1ϱh(ζ,ϱ)as*(ζ,ϱ)bs(ζ,ϱ).
(176)

Let us also adopt the quasioptical ordering as we did for scalar waves in Sec. IV E. In particular, we allow ΔVμ=O(ϵ). Then, Eq. (164) becomes

(Λ+iΓ+K̂+Ĝ)a=0.
(177)

Here, K̂ is defined as in Eq. (167) and can be approximated as follows:

K̂a=iVa,ζi2g(gV),ζa+ΔK̂a,
(178)
ΔK̂aUaiΔVσa,σi2h(hΔVσ),σa,
(179)
ΔVσ=Ξ+DH|σΞ,
(180)

where VVavr0 (we used the fact that Vavrσ=0 by definition of the ray-based coordinates), U=O(ϵ||) is given by Eq. (171), and Γ is given by Eq. (166). Finally,

ĜΞ+D̂H2ΞΞ+D̂H1Ξ¯Λ¯1Ξ¯+D̂H1Ξ.
(181)

As shown in Appendix B 3, Ĝ can also be simplified as

Ĝa12h1/4[Λ|σσ̃(h1/4a),σ̃],σ,
(182)

and so it is also self-adjoint under the inner product (176).

In summary, the quasioptical equation for vector waves can be written as follows:

Λa+iΓaiVa,ζi2g(gV),ζa+Ĝa+ΔK̂a=0.
(183)

Equation (183) is the main result of our paper. Like Eq. (172), this equation is covariant within the accuracy of our theory. As a reminder, a is generally a vector (dima=N1), whose elements are the scalar amplitudes of active modes (Sec. V D 2); Λ is the diagonal matrix formed by the eigenvalues of DH; Γ is given by Eq. (166); V=N1trΛ|ζ is a scalar; |ζ/k0; Ĝ is given by Eq. (182); and ΔK̂ is given by Eq. (179). There, U is given by Eq. (171) and ΔV is given by Eq. (180). The derivatives of h could be neglected within the assumed accuracy, but a purist might want to retain them in order to keep ΔK̂ and Ĝ precisely self-adjoint under the inner product (176).

Like in Sec. IV E 2, we require the α parameter of the ray-based coordinates to be defined as ϱ-independent. Then, Eq. (183) has the following corollary:

1αddζa|αV|aN=2a|Γ|aN,
(184)

which is similar to Eq. (84). For the case of zero Γ, Eq. (184) predicts conservation of the wave-action flux through the beam cross section,

ψ|αV|ψN=const.
(185)

Unlike in Eq. (85), this is the flux of all active modes combined, and the fluxes of individual active modes may not be conserved separately.

2. Simplified equations

If the metric is nearly Euclidean (or pseudo-Euclidean) as in Sec. IV E 3, we can drop the metric factors and rewrite Eq. (183) as follows:

Λa+iΓaiVa,ζi2V,ζa12(Λ|σσ̃a,σ̃),σUaiΔVσa,σi2ΔVσ,σa=0.
(186)

Using the variable transformation ϕVa, this equation can be further simplified as

iϕ,ζ=χ̂ϕ+iϒϕ.
(187)

Here, χ̂ is an operator self-adjoint under the inner product (176) and given by

χ̂ϕQϕ12(Σσσ̃ϕ,σ̃),σΔVσViϕ,σi2(ΔVσV),σϕ.

Also, Q, Σ, and ϒ are the self-adjoint matrices given by

Q(ΛU)/V,Σσσ̃Λ|σσ̃/V,ϒΓ/V.
(188)

Accordingly, Eq. (184) becomes

ddζϕ|ϕN=2ϕ|ϒ|ϕN,
(189)

and so ϕ|ϕN is conserved if ϒ is zero.

The quasioptical model proposed above is equally applicable to scalar beams, single-mode vector beams (N =1), and multimode vector beams (N >1).

1. Scalar beams

In the case of a scalar beam (Sec. IV), one has Ξ=1, and so

Λ=DH,Γ=DA,U=0,ΔVσ=0.
(190)

Then, the equations from Sec. IV E are reproduced. In particular, Λ is made zero by the choice of k [Eq. (64)]. This implies that the ray Hamiltonian is H=Λ, which leads to the following ray equations:

dxαdτ=Λkα,dkαdτ=Λxα.
(191)

2. Single-mode vector beams

In the case of a single-mode vector beam (N =1), one has Ξ=η, where η is the polarization vector. Then,

Λ=ηDHη,Γ=ηDAη,
(192)

and so they are scalars. Also, ΔVσ=0, but U is generally a nonzero scalar function. There are two natural ways to define k in this case. One is to require that Λ=0, i.e., to adopt H=Λ. In this case, U is left in the envelope equation and can intensify the envelope inhomogeneity in the direction perpendicular to the group velocity. This may eventually result in the violation of the envelope approximation. (For multimode beams, this is less of a concern because they typically split into single-mode beams before the issue becomes significant.) Hence, it is potentially advantageous to define k by requiring ΛU=0, i.e., by adopting H=ΛU. In this case, the amplitude equation is identical to that of a scalar wave, while the effect of U is absorbed in the reference phase, leading to modified ray equations,

dxαdτ=(ΛU)kα,dkαdτ=(ΛU)xα.
(193)

The effect of U on the ray trajectories is known as the (spin) Hall effect of light in optics.22,23,61–64 In plasma physics, this effect is typically neglected. (To our knowledge, all existing ray-tracing codes ignore U entirely.)

Finally, note that the ray dynamics can also be represented in an alternative form. Since Λ is a scalar, one can rewrite Eq. (171) for U(x,k) as follows:

U=U0Λ|μIm(ηη|μ)+Λ|μIm(ηη|μ)U0ẋμAμ(x)k̇μA(k)μ.
(194)

Since there is only one mode, there is no mode index to manipulate, and so the “metric” factor δ has been dropped. Likewise, the anti-Hermitian parts are simply the imaginary parts in this case. We also substituted the GO ray equations (191), where we replaced d/dτ with dots, and adopted

U0Im(ηH|μD¯Hη|μ)ηHC¯Hη,
(195)
Aμ(x)Im(ηη|μ),A(k)μIm(ηη|μ).
(196)

Then, the phase-space Lagrangian of a ray, L[x,k]=kμẋμH, has a noncanonical structure, namely,

L=[kμ+Aμ(x)]ẋμ+A(k)μk̇μ(ΛU0).
(197)

The equations originating from a special case of this noncanonical phase-space Lagrangian were used, for example, in Ref. 62 to study the Hall effect for light propagating in a nonbirefringent material. A comparison of the resulting noncanonical ray equations with the canonical ray equations (193) can be found in Ref. 27. Also notably, the terms Aμ(x) and A(k)μ are known as the Berry connection and are linked to the Berry phase or the geometrical phase of light.27,44,62

3. Multimode vector beams

In the case of multiple active modes (N >1), a is an N-dimensional vector, and the coefficients in the amplitude equation are matrices (except for V, which is a scalar). In particular, the matrices Γ,ΔVσ, and U are generally nondiagonal, and so they can cause mode conversion. In the simplest case, when both the transverse gradients and dissipation are negligible, this process is most transparently described by Eq. (187), which then becomes

iϕ,ζ=Qϕ,
(198)

with Q being a self-adjoint matrix. The modes decouple when Q is close to diagonal. More generally, the dynamics governed by Eq. (198) is similar to that of an N-level quantum system with a Hamiltonian Q (and to the dynamics of N coupled classical oscillators whose parameters do not change too rapidly with ζ65). Alternatively, it can be mapped to a precession equation for a real (N21)-dimensional “spin” vector.26 This approach is particularly intuitive at N=2, when the spin is three-dimensional. A detailed analysis of mode conversion for this case can be found in Ref. 29.

As a reminder, the model presented here (Sec. V F) relies on the assumption that the group velocities of the active modes are close to each other. Otherwise, beam splitting occurs rapidly, and the assumption that a,ζa,σ does not hold. The quasioptical description is inapplicable to such beams; however, the first-order XGO model described in Sec. V E can be used.

Here, we shall explain how the above theory applies to EM waves. We start with Maxwell's equation68 

1gα(gFαβ)=4πcJβ,
(199)

where Fαβ is the EM tensor, namely,

Fαβ=gαμgβνFμν,FαβαAββAα,
(200)

A is the vector potential, and J is the current density. This equation can also be represented as follows:

D̂vac(A)A=4πcJ.
(201)

Here, D̂vac(A) is the vacuum dispersion operator,

[D̂vac(A)]αβAβ=1gλ{ggαμgλν×[ν(gμβAβ)μ(gνβAβ)]},
(202)

which is self-adjoint under the inner product (98). By assuming J=cϑ̂A, where ϑ̂ is some linear operator, one can cast the equation for A in the form (1), namely,

[D̂vac(A)+4πϑ̂]A=0.
(203)

Hence, the theory developed in the previous sections readily applies, specifically, with the dimension of the configuration space being n =4, the dimension of the vector field A being m =4 (so m = n), and γ=g. The corresponding Weyl symbols are calculated as follows. It can be shown straightforwardly that [D¯̂vac(A)]αβgαγ[D̂vac(A)]γβ is approximately given by

[D¯̂vac(A)]αβp̂αp̂βp̂μgμνgαβp̂ν+i(qαp̂βqβp̂α)+i[(β)μα(α)μβ]p̂μ.
(204)

Here, we omitted second-order derivatives of g, which are negligible within the accuracy of our theory. Also, q is defined as in Eq. (15), and μg1g,μ, which is in agreement with Eq. (127) because g=γ. Finally, in the form introduced in Sec. V C, the above result is

[D¯(A)]αβpαpβgαβgμνpμpν+4π(ϑ¯0)αβ,Cαβ(A)i(qαkβqβkγ)+i[(β)μα(α)μβ]kμ+4π(Δϑ¯)αβ

where (ϑ¯0)αβ is the GO limit of ϑ¯,(Δϑ¯)αβ is the remaining part of ϑ¯, and ϑ¯ itself is the symbol of ϑ̂¯.

As a side remark, these equations can be used to calculate the Hall effect of light in gravitational fields, i.e., the deviation of vacuum light rays from geodesics in curved spacetime. (Special cases of this effect are discussed in Refs. 23, 63, and 64.)

As a special case, suppose a metric of the form

g=(100h),
(205)

with time-independent spatial metric h. Then, the above equations can be simplified as follows. Let us assume the Weyl gauge (A0=0). Then, it is sufficient to consider just the spatial part of the four-dimensional Eq. (203) and replace the vector potential with the electric field,

EaF0a=ip̂0Aa.
(206)

(As usual, p̂0ic1t, which is a self-adjoint operator. Assuming we work with wave fields that have zero time average, p̂0 can also be considered reversible.) Unlike earlier, we now use the Latin indices to denote spatial components, and we shall adopt the bold font for spatial vectors and matrices when using the index-free notation. Specifically, one obtains

D̂(E)E=0,
(207)
D̂(E)(p̂0)1[D̂vac(A)+4πϑ̂](p̂0)1,
(208)

where we also multiplied the equation by (p̂0)1. This has the form (1) with n =4, m =3, and γ=h.

Let us introduce the conductivity operator σ̂ via J=σ̂E and notice that 4πσ̂=icp̂0χ̂ by the definition of the susceptibility operator χ̂. Then, 4πϑ̂=p̂0χ̂p̂0, and so D̂(E) can be expressed as follows:

D̂(E)(p̂0)1D̂vac(A)(p̂0)1+χ̂,
(209)

or equivalently, D̂(E)=(p̂0)2D̂vac(A)+χ̂. The corresponding Weyl image is D¯ab(E)=(p0)2[D¯vac(A)]ab+χ¯ab, or

D¯ab(E)=(p0)2[papbpcpdhcdhab+C¯ab(A)]+ε¯ab,
(210)

where ε¯abhab+(χ¯0)ab serves as the dielectric tensor and χ¯0 is the GO limit of χ¯.

Using the fact that the dispersion operator is defined only up to a constant factor, let us introduce an additional factor 1/(16π). Then, Eq. (210) becomes

D¯ab(E)=D¯ab(E)+C¯ab(E),
(211)
D¯ab(E)=116πp02(papbpcpdhcdhab)+ε¯ab16π,
(212)
C¯ab(E)116πp02C¯ab(A).
(213)

The quantity C¯ab(A) was derived in Sec. VI A, and the vector q [Eq. (15)] and the matrices a [Eq. (127)] can be written as

qa=14aln|deth|,a=h1ah,
(214)

and so they both can be of order ϵ||; hence, C=O(ϵ||). Note that such C is generally non-negligible even for near-Euclidean metrics such as those discussed in Sec. IV E 3.

From Eq. (174), we obtain

Jα=Ea*Ebkα[D¯H(E)(t,x,ω,k)]ab,
(215)

which acts as the true action-flux density.54 As a reminder, the action density IJ0 can be written as

IEa*Eb16πω2ω{ω2[ε¯H(t,x,ω,k)]ab},
(216)

where we used [D¯H(E)]abEb0. Using the latter and Faraday's law Bck×E/ω for the magnetic field B, one can also rewrite I in another common form,4 

IEa*Eb16πωω{ω[ε¯H(t,x,ω,k)]ab}+Ba*Ba16πω.
(217)

One can also show54 that the spatial component of the action flux density (215) can be expressed as

JSωEa*Eb16πk[ε¯H(t,x,ω,k)]ab,
(218)

where S is the (time- or space-averaged) Poynting vector,

S=c8πRe(E×B*).
(219)

The GO equation (173) can be written as

It+1hxa(hJa)=Ea*Eb8π[ε¯A(t,x,ω,k)]ab.

The corresponding density of the wave energy is ωI, and the density of the energy flux is ωJ, as flows from the variational principle.54 

For stationary waves with fixed frequency, one has p̂0=ω/c and ω can be treated as a constant parameter. Such waves can be studied on the three-dimensional configuration space, namely, the physical space x. In this case, n=m=3 and γ=h. If there is no spatial dispersion, then χ̂=χ(x̂), where the dependence on ω is assumed but not emphasized. Assuming that the underlying space is flat (which is always the case in laboratory applications), the symbol χab is then identical to the susceptibility of the homogeneous medium. (Other formulations of the susceptibility have also been proposed in this case, e.g., for light in nondispersive dielectric media27 and for waves in cold plasmas.24,25,67)

These properties can also be extended, approximately, to media with weak spatial dispersion, i.e., such that

χ¯ab=χ¯ab(0)(x)+χ¯ab(k)(x,p),χ¯ab(k)χ¯ab(0).
(220)

For example, in plasma, χ¯ab(k) is proportional to the temperature, and for EM waves, thermal effects often can be considered as small perturbations.4 Although the Weyl symbol χ¯ab(k) is, strictly speaking, different from the corresponding part of the susceptibility in homogeneous medium, the difference is of order ϵχ¯ab(k), and so it is much smaller than ϵ and thus can be neglected. In other words, one can simply replace the small χ¯ab(k) with its GO limit, which is usually well known.4 

In principle, our general method is also applicable to waves in media with strong spatial dispersion. However, χ̂ may be hard to calculate in that case unless a medium is homogeneous, and no extrapolation of a known homogeneous-medium model is guaranteed to yield the true general χ̂. A related discussion can also be found in Ref. 43.

Let us also discuss the special case when the configuration space is flat and Eq. (220) applies, as in typical laboratory applications. Although the ray-based metric h is still non-Euclidean in this case (unless the reference rays are straight), one can use the flatness of the configuration space in the following sense. In the Euclidean laboratory coordinates x̃, the spatial metric has the form h̃ab=δab and C̃ vanishes. (Here and further, the tildes denote that the corresponding quantities are evaluated in the laboratory coordinates. Note that Papers II and III assume a different notation; there, tildes are quantities in the ray-based coordinates, and nontilded are those in the laboratory coordinates.) This simplifies the expression for the dispersion matrix and makes DH a true tensor (modulo the insignificant corrections caused by weak spatial dispersion discussed in Sec. VI C). Then, since U and Γ are frame-invariant (Sec. V E), one can calculate them in the laboratory coordinates using

U=Ũ(Λ̃|μΞ̃HΞ̃|μΛ̃|μΞ̃HΞ̃|μ+Ξ̃H|μD̃HΞ̃|μ)A,
(221)
Γ=Γ̃=Ξ̃HD̃AΞ̃,
(222)

where both δ1 and γ̃ are omitted for simplicity, since they are unit matrices. This also extends to the quasioptical model. Since calculations in the laboratory coordinates are generally easier than the corresponding calculations in the ray-based coordinates, this simplifies applications of our theory to numerical simulations, as discussed further in Paper II.

Specific applications of this theory, including those to waves in fusion plasmas, will be discussed in future publications. Here, we only note that our formulation relies on the GO ordering and thus is inapplicable to wave transformations near caustics, for example, as in the O–X conversion caused by the density gradient in dense magnetized plasmas.68 In contrast, the O–X conversion caused by the magnetic shear in dilute plasmas29 can be modeled naturally, as also demonstrated explicitly in Paper III. Any adiabatic transformations of waves along continuous dispersion curves can also be modeled naturally, specifically, within the single-mode approximation summarized in Sec. V G 2. (In special cases like the X–B transformation,4 when the group velocity becomes zero on a ray, while k remains large, one may want to replace the coordinate ζ with τ, where dτ=dζ/V, to remove the singularity in the amplitude equation.) In all these cases, our theory is expected to yield results which agree with full-wave modeling up to local errors of order O(ϵ), because this is the accuracy within which the passive-mode contribution is calculated (Sec. V D 4).

In summary, we propose a quasioptical theory of mode-converting wave beams in inhomogeneous media such as plasma. This includes the following. For any given dispersion operator D̂ that governs the original wave field Ψ, we explicitly calculate the approximate operator D̂ that governs the wave envelope ψ to the second order in the GO parameter ϵ. Then, we further simplify this envelope operator by assuming that the gradient of ψ transverse to the local group velocity is much larger than the corresponding parallel gradient. This leads to a parabolic differential equation for ψ (“quasioptical equation”) on the basis of the GO polarization vectors [Eq. (136)]. Our main results can be found in Sec. V F, which includes the general quasioptical equation (183) and its special case (187). Scalar and mode-converting vector beams are described on the same footing (Sec. V G). We also explain how to apply this model to EM waves considered as a special case (Sec. VI). In the follow-up papers, we report successful quasioptical modeling of radio frequency wave beams in magnetized plasma based on this theory.

See the supplementary material for an overview of the Weyl calculus on a curved configuration space. There, we present proofs of the statements summarized in Sec. IV A.

The first author (I.Y.D.) acknowledges the support and hospitality of the National Institute of Fusion Science, Japan. The authors are also thankful to M. A. Oancea for stimulating discussions regarding the XGO generalization to non-Euclidean metrics. This work was supported by JSPS KAKENHI Grant No. JP17H03514. This work was also supported by the U.S. DOE through Contract No. DE-AC02-09CH11466 and by the Laboratory Directed Research and Development program at Sandia National Laboratories. Sandia National Laboratories is a multimission laboratory managed and operated by National Technology and Engineering Solutions of Sandia, LLC., a wholly owned subsidiary of Honeywell International, Inc., for the U.S. DOE National Nuclear Security Administration under Contract No. DE-NA-0003525. This paper describes objective technical results and analysis. Any subjective views or opinions that might be expressed in this paper do not necessarily represent the views of the U.S. Department of Energy or the United States Government.

Here, we present a summary of the selected notations used in the main text.

1. Basic symbols

denotes a definition.

̂ denotes an operator.

denotes either a dual vector [Eqs. (100) and (141)] or an adjoint operator.

T denotes the matrix transpose.

H=T* denotes the conjugate matrix transpose.

+ is used only in Ξ+ and Ξ¯+ defined in Eqs. (146).

H and A denote the Hermitian and anti-Hermitian parts of an operator [Eqs. (26)] or those of a matrix (Sec. V B). When applied to a scalar, H and A denote the real and imaginary parts, respectively.

• The underline notation _ is explained in Sec. V B and is used for matrices with lower indices only.

Wz and W1 denote the Wigner–Weyl transform and its inverse (Sec. IV A 3), respectively.

2. Index manipulation

On the n-dimensional configuration space Mn, we assume a general metric tensor g with components gμν. For (co)vector fields on the space (co)tangent to Mn, the indices are manipulated as usual, namely, via

Xμ=gμνXν,μ,ν=0,1,,(n1).
(A1)

(Summation over repeated indices is always assumed unless specified otherwise.) For vector waves, we also introduce an additional vector space that may or may not be the same as the space tangent to Mn (Sec. V A). On that space, an additional metric γ is introduced, and the indices are manipulated as follows:

Ψa=γabΨb,a,b=1,2,,m.
(A2)

In special cases, Eqs. (A1) and (A2) can be equivalent, but that is not a generic situation.

We also introduce the Euclidean metric δ on the space of amplitude vectors a (Sec. V D 2),

as=δssas.
(A3)

For active modes, which are of our primary interest, the mode indices s and s range from 1 to N. The distinction between as and as is made only for esthetic reasons (consistency of notation) and can be neglected otherwise.

3. Derivatives

For spatial derivatives, we use the notation that is standard, for example, in general relativity.69 In particular, Xμ;μ is the divergence, namely,

Xμ;μ=1g(gXμ),μ.
(A4)

Here, g|detg|,

f,μμffxμ, f,μν2fxμxν.
(A5)

For functions of the form f(x,k), we also introduce the following partial derivatives:

f|μf(x,k)xμ,f|μν2f(x,k)xμxν,
(A6)
f|μf(x,k)kμ,f|μν2f(x,k)kμkν.
(A7)

Since k=θ(x), one has

f,μ(x,k(x))=f|μ+f|νkν|μ,
(A8)
kν|μ=kν,μ=kμ,ν=θ,μν=θ,νμ.
(A9)
4. Inner products

The inner product for scalar waves on Mn is

Ψ|Φ=dnxg(x)Ψ*(x)Φ(x),
(A10)

and for m-dimensional vector waves on Mn,

Ψ|Φmdnxg(x)γab(x)Ψa*(x)Φb(x).
(A11)

In particular, in the x representation,

Ψ(x)Ψ|x=(Ψ1*,Ψ2*,,Ψm*).
(A12)

We also use the dot product

Ψ·ΦΨ(x)Φ(x)=Ψa*(x)Φa(x).
(A13)

The inner product on the transverse space Mn1 is defined for scalar fields as

ψ|ϕdn1ϱh(ζ,ϱ)ψ*(ζ,ϱ)ϕ(ζ,ϱ)
(A14)

and for the N-dimensional “active” parts of the vector fields (Sec. V D 2) as

a|bNdn1ϱh(ζ,ϱ)as*(ζ,ϱ)bs(ζ,ϱ).
(A15)

Here, we present some auxiliary calculations whose results are used in Sec. V.

1. Calculation of K̂ and U(x)

We start with Eq. (167) for K̂. This equation can be rewritten as follows:

K̂a=12Ξ+(p̂μVHμ+VHμp̂μiμVHμ)Ξa+Ξ+CHΞ=i2g1/4[Ξ+(g1/4VHμΞa),μ+Ξ+VHμ(g1/4Ξa),μ]i2Ξ+μVHμΞa+Ξ+CHΞ=ig1/4(g1/4),μΞ+VHμΞaiΞ+VHμΞa,μi2[Ξ+(VHμΞ),μ+Ξ+VHμΞ,μ]ai2Ξ+μVHμΞa+Ξ+CHΞ=i2(lng),μVμaiVμa,μi2[Vμ,μ(Ξ+),μVHμΞ+Ξ+VHμΞ,μ+Ξ+μVHμΞ]a+Ξ+CHΞ=iVμa,μi2Vμ;μai2[Ξ+VHμΞ,μ(Ξ+),μVHμΞ+Ξ+μVHμΞ]a+Ξ+CHΞ,
(B1)

where we used Eq. (52) and introduced

VμΞ+VHμΞ=(Ξ+DHΞ)|μΞ+|μDHΞΞ+DHΞ|μ=Λ|μΞ+|μΞΛΛΞ+Ξ|μΛ|μ,
(B2)

where, at the end, we used Λ=O(ϵ) (Sec. V D 4), so that the last two terms in Eq. (B2) can be neglected.

From Eq. (154), we have (Ξ+),μ=δ1Ξ,μHγ+Ξ+μ. Using this and also Eq. (130), one obtains

K̂a=iVμa,μi2Vμ;μaUa,U=U0Ξ+CHΞ,
(B3)
U0=i2δ1(ΞHV¯HμΞ,μΞ,μHV¯HμΞ)=δ1(Ξ,μHV¯HμΞ)A,
(B4)

where the subscript A denotes the anti-Hermitian part, as usual.

2. Calculation of U0(x,k)

Let us also derive an alternative expression for U0 [Eq. (B4)] in terms of the partial derivatives |μ instead of the full derivatives ,μ. Using Eq. (59), one obtains Ξ,μ=Ξ|νkν,μ+Ξ|μ. Then, Eq. (B4) can be rewritten as follows:

U0=i2δ1(ΞHV¯HμΞ|νΞH|νV¯HμΞ)kν,μ+i2δ1(ΞHV¯HμΞ|μΞ|μHV¯HμΞ).
(B5)

By differentiating Eq. (159) with respect to kμ, we obtain

VHμΞ=Ξ|μΛ+ΞΛ|μDHΞ|μΞΛ|μDHΞ|μ,
(B6)

where we used that Λ (but not its partial derivatives) is small. Let us multiply this by γ to obtain

V¯HμΞγΞΛ|μD¯HΞ|μ,ΞHV¯HμΛ|μΞHγΞH|μD¯H,
(B7)

where the latter equality is the complex conjugate of the former. Using these and kν,μ=θ,νμ=θ,μν=O(ϵ), we further obtain, to the leading order,

(ΞHV¯HμΞ|νΞH|νV¯HμΞ)kν,μ(Λ|μΞHγΞ|νΞH|μD¯HΞ|νΞH|νγΞΛ|μ+ΞH|νD¯HΞ|μ)kν,μ=(Λ|μΞHγΞ|νΞH|νγΞΛ|μ)kν,μ=(Λ,μΞHγΞ|μΞH|μγΞΛ,μ)(Λ|μΞHγΞ|μΞH|μγΞΛ|μ)(Λ|μΞHγΞ|μΞH|μγΞΛ|μ),
(B8)

where we neglected terms of the second order in ϵ. We also used the fact that Λs(x,k(x))O(ϵ) by the definition of active modes, in addition, Λs are changing slowly, and so Λ,μO(ϵ2). (Note that this is only true for Λ,μ but not for Λ|μ.) Using Eqs. (B7), we also obtain similarly that

ΞHV¯HμΞ|μΞ|μHV¯HμΞΛ|μΞHγΞ|μΞH|μD¯HΞ|μΞ|μHγΞΛ|μ+Ξ|μHD¯HΞ|μ.
(B9)

Then,

U0=δ1(Λ|μΞHγΞ|μΞH|μγΞΛ|μΛ|μΞHγΞ|μ+ΞH|μD¯HΞ|μ+Ξ|μHγΞΛ|μΞ|μHD¯HΞ|μ)/(2i)=δ1(Λ|μΞHγΞ|μΛ|μΞHγΞ|μ+ΞH|μD¯HΞ|μ)A.
(B10)

As a reminder, the subscript A denotes the anti-Hermitian part, and δ1 is a unit matrix which only raises the mode index. Equation (B10) is similar to the corresponding result derived in Refs. 24–26. Extra terms are included in those papers in the expressions for U0, but they are of the second order in ϵ and could have been neglected.

3. Calculation of Ĝ

Consider Eq. (181) for Ĝ derived under the assumption of the quasioptical ordering. Here, we simplify that equation and derive the explicit formula (182). First, note that

Ξ+D̂H2Ξa12Ξ+p̂μΘHμνp̂νΞa12h1/4[Ξ+DH|σσ̃Ξ(h1/4a),σ̃],σ,
(B11)

where we used Eq. (129). Also, using Eq. (128) for D̂H1, one obtains

Ξ+D̂H1Ξ¯Λ¯1Ξ¯+D̂H1Ξa(Ξ+VHσΞ¯Λ¯1Ξ¯+VHσ̃Ξ)a,σσ̃=[Ξ+VH(σΞ¯Λ¯1Ξ¯+VHσ̃)Ξ]a,σσ̃h1/4[Ξ+VH(σΞ¯Λ¯1Ξ¯+VHσ̃)Ξ(h1/4a),σ̃],σ.
(B12)

Here, in the first line, we substituted Eq. (128) for D̂H1 and also Eq. (14) for p̂μ. Then, we only kept the terms involving derivatives of a along the perpendicular direction of the wave beam. The parentheses in the indices denote symmetrization, namely, for any A and B,

A(σBσ̃)=12(AσBσ̃+Aσ̃Bσ).
(B13)

The metric factor h has been added to keep the operator self-adjoint under the inner product (176). (Accounting for the inhomogeneity of h in Ĝ is beyond the accuracy of our theory. However, keeping Ĝ self-adjoint is convenient and physically meaningful, for it is known that the exact operator is not responsible for dissipation.) Hence,

Ĝ12h1/4[Φσσ̃(h1/4a),σ̃],σ,Φσσ̃=Ξ+DH|σσ̃Ξ2Ξ+DH(|σΞ¯Λ¯1Ξ¯+DH|σ̃)Ξ.
(B14)

In order to simplify the expression for Φσσ̃, note that

Ξ¯+DH|σ̃Ξ=(Ξ¯+DHΞ)|σ̃Ξ¯+|σ̃DHΞΞ¯+DHΞ|σ̃=(Ξ¯+ΞΛ)|σ̃Ξ¯+|σ̃ΞΛΛ¯Ξ¯+Ξ|σ̃Λ¯Ξ¯+|σ̃Ξ,
(B15)

where we used Ξ¯+Ξ=0 (and thus Ξ¯+Ξ|σ̃=Ξ¯+|σ̃Ξ) and neglected terms proportional to Λ=O(ϵ||). Hence,

Ξ+DH|σΞ¯Λ¯1Ξ¯+DH|σ̃Ξ=Ξ+Ξ¯|σΛ¯Λ¯1Λ¯Ξ¯+|σ̃Ξ=Ξ+Ξ¯|σΛ¯Ξ¯+|σ̃Ξ.
(B16)

Using Eqs. (152) for Ξ and Ξ¯ like we did above, Eq. (156) for DH, and the fact that Λ=O(ϵ), we finally obtain

Φσσ̃=Ξ+(ΞΛΞ++Ξ¯Λ¯Ξ¯+)|σσ̃Ξ2Ξ+Ξ¯(|σΛ¯Ξ¯+|σ̃)Ξ=Ξ+(Ξ|σΛΞ++ΞΛ|σΞ++ΞΛΞ+|σ+Ξ¯|σΛ¯Ξ¯++Ξ¯Λ¯|σΞ¯++Ξ¯Λ¯Ξ¯+|σ)|σ̃Ξ2Ξ+Ξ¯(|σΛ¯Ξ¯+|σ̃)ΞΞ+(Ξ|σΛ|σ̃Ξ++Ξ|σ̃Λ|σΞ++ΞΛ|σσ̃Ξ++ΞΛ|σΞ+|σ̃+ΞΛ|σ̃Ξ+|σ+Ξ¯|σΛ¯Ξ¯+|σ̃+Ξ¯|σ̃Λ¯Ξ¯+|σ)Ξ2Ξ+Ξ¯(|σΛ¯Ξ¯+|σ̃)Ξ=Λ|σσ̃+Ξ+Ξ|σΛ|σ̃+Ξ+Ξ|σ̃Λ|σΛ|σΞ+Ξ|σ̃Λ|σ̃Ξ+Ξ|σΛ|σσ̃+[Ξ+Ξ(|σ,Vσ̃)],
(B17)

where we substituted Λ|σVσ [Eq. (B2)]. The quasioptical approximation implies that Vσ̃ is close to a scalar matrix, and so the commutator [Ξ+Ξ(|σ,Vσ̃)] can be neglected. Hence, Φσσ̃Λ|σσ̃.

1.
E. R.
Tracy
,
A. J.
Brizard
,
A. S.
Richardson
, and
A. N.
Kaufman
,
Ray Tracing and Beyond: Phase Space Methods in Plasma Wave Theory
(
Cambridge University Press
,
New York
,
2014
).
2.
Yu. A.
Kravtsov
and
Yu. I.
Orlov
,
Geometrical Optics of Inhomogeneous Media
(
Springer-Verlag
,
New York
,
1990
).
3.
G. B.
Whitham
,
Linear and Nonlinear Waves
(
Wiley
,
New York
,
1974
).
4.
T. H.
Stix
,
Waves in Plasmas
(
AIP
,
New York
,
1992
).
5.
P. T.
Bonoli
, “
Review of recent experimental and modeling progress in the lower hybrid range of frequencies at ITER relevant parameters
,”
Phys. Plasmas
21
,
061508
(
2014
).
6.
See http://www.compxco.com/genray.html for the GENRAY code by A. P. Smirnov and R. W. Harvey.
7.
N. B.
Marushchenko
,
Y.
Turkin
, and
H.
Maassberg
, “
Ray-tracing code TRAVIS for ECR heating, EC current drive and ECE diagnostic
,”
Comput. Phys. Commun.
185
,
165
(
2014
).
8.
T. I.
Tsujimura
,
S.
Kubo
,
H.
Takahashi
,
R.
Makino
,
R.
Seki
,
Y.
Yoshimura
,
H.
Igami
,
T.
Shimozuma
,
K.
Ida
,
C.
Suzuki
,
M.
Emoto
,
M.
Yokoyama
,
T.
Kobayashi
,
C.
Moon
,
K.
Nagaoka
,
M.
Osakabe
,
S.
Kobayashi
,
S.
Ito
,
Y.
Mizuno
,
K.
Okada
,
A.
Ejiri
,
T.
Mutoh
, and
LHD Experiment Group,
Development and application of a ray-tracing code integrating with 3D equilibrium mapping in LHD ECH experiments
,”
Nucl. Fusion
55
,
123019
(
2015
).
9.
E.
Mazzucato
, “
Propagation of a Gaussian beam in a nonhomogeneous plasma
,”
Phys. Fluids B
1
,
1855
(
1989
).
10.
D.
Farina
, “
A quasi-optical beam-tracing code for electron cyclotron absorption and current drive: GRAY
,”
Fusion Sci. Technol.
52
,
154
(
2007
).
11.
G. V.
Pereverzev
, “
Use of the multidimensional WKB method to describe propagation of lower hybrid waves in tokamak plasma
,”
Nucl. Fusion
32
,
1091
(
1992
).
12.
G. V.
Pereverzev
, “
Beam tracing in inhomogeneous anisotropic plasmas
,”
Phys. Plasmas
5
,
3529
(
1998
).
13.
E.
Poli
,
A. G.
Peeters
, and
G. V.
Pereverzev
, “
TORBEAM, a beam tracing code for electron-cyclotron waves in tokamak plasmas
,”
Comput. Phys. Commun.
136
,
90
(
2001
).
14.
E.
Poli
,
A.
Bock
,
M.
Lochbrunner
,
O.
Maj
,
M.
Reich
,
A.
Snicker
,
A.
Stegmeir
,
F.
Volpe
,
N.
Bertelli
,
R.
Bilato
,
G. D.
Conway
,
D.
Farina
,
F.
Felici
,
L.
Figini
,
R.
Fischer
,
C.
Galperti
,
T.
Happel
,
Y. R.
Lin-Liu
,
N. B.
Marushchenko
,
U.
Mszanowski
,
F. M.
Poli
,
J.
Stober
,
E.
Westerhof
,
R.
Zille
,
A. G.
Peeters
, and
G. V.
Pereverzev
, “
TORBEAM 2.0, a paraxial beam tracing code for electron-cyclotron beams in fusion plasmas for extended physics applications
,”
Comput. Phys. Commun.
225
,
36
(
2018
).
15.

It is not a purpose of this paper to provide a comprehensive list of existing approaches to reduced modeling of EM waves in plasmas, so the above reference list is not intended as exhaustive.

16.
A. A.
Balakin
,
M. A.
Balakina
,
G. V.
Permitin
, and
A. I.
Smirnov
, “
Quasi-optical description of wave beams in smoothly inhomogeneous anisotropic media
,”
J. Phys. D: Appl. Phys.
40
,
4285
(
2007
).
17.
A. A.
Balakin
,
M. A.
Balakina
,
G. V.
Permitin
, and
A. I.
Smirnov
, “
Scalar equation for wave beams in a magnetized plasma
,”
Plasma Phys. Rep.
33
,
302
(
2007
).
18.
For applications and comparison with other codes, see also
A. A.
Balakin
,
M. A.
Balakina
, and
E.
Westerhof
, “
ECRH power deposition from a quasioptical point of view
,”
Nucl. Fusion
48
,
065003
(
2008
);
O.
Maj
,
A. A.
Balakin
, and
E.
Poli
, “
Effects of aberration on paraxial wave beams beam tracing versus quasioptical solutions
,”
Plasma Phys. Controlled Fusion
52
,
085006
(
2010
).
19.
There have been works on quasioptical modeling for systems where wave beams are naturally aligned with an axis of symmetry. For example, see
C. K.
Phillips
,
F. W.
Perkins
, and
D. Q.
Hwang
, “
Parabolic approximation method for fast magnetosonic wave propagation in tokamaks
,”
Phys. Fluids
29
,
1608
(
1986
). However, the methods used there do not easily extend to beam modeling in more general settings.
20.

This does not extend to ray-tracing codes. For example, a ray-tracing code raycon1 captures mode conversion by using an approximate yet somewhat generic analytic solution that applies in the limit of a narrow region of resonant coupling. A related ray-tracing model with mode conversion was also proposed in Refs. 24 and 25.

21.

Mode conversion is understood here as the exchange of action between near-resonant but nonintersecting branches of the dispersion relation.1 Mathematically, this process is similar to quantum tunneling. Adiabatic transformations of a wave mode along continuous dispersion curves, such as the X–B transformation,4 are not mode conversion in the true sense of the term and can also be simulated using already existing reduced models.

22.
K. Y.
Bliokh
,
F. J.
Rodríguez-Fortuño
,
F.
Nori
, and
A. V.
Zayats
, “
Spin-orbit interactions of light
,”
Nat. Photonics
9
,
796
(
2015
).
23.
M. A.
Oancea
,
C. F.
Paganini
,
J.
Joudioux
, and
L.
Andersson
, “
An overview of the gravitational spin Hall effect
,” e-print arXiv:1904.09963.
24.
D. E.
Ruiz
, “
Geometric theory of waves and its applications to plasma physics
,” Ph.D. thesis (Princeton University,
2017
);
D. E.
Ruiz
, e-print arXiv:1708.05423.
25.
D. E.
Ruiz
and
I. Y.
Dodin
, “
Extending geometrical optics: A Lagrangian theory for vector waves
,”
Phys. Plasmas
24
,
055704
(
2017
).
26.
D. E.
Ruiz
and
I. Y.
Dodin
, “
Lagrangian geometrical optics of nonadiabatic vector waves and spin particles
,”
Phys. Lett. A
379
,
2337
(
2015
).
27.
D. E.
Ruiz
and
I. Y.
Dodin
, “
First-principles variational formulation of polarization effects in geometrical optics
,”
Phys. Rev. A
92
,
043805
(
2015
).
28.
D. E.
Ruiz
,
C. L.
Ellison
, and
I. Y.
Dodin
, “
Relativistic ponderomotive Hamiltonian of a Dirac particle in a vacuum laser field
,”
Phys. Rev. A
92
,
062124
(
2015
).
29.
I. Y.
Dodin
,
D. E.
Ruiz
, and
S.
Kubo
, “
Mode conversion in cold low-density plasma with a sheared magnetic field
,”
Phys. Plasmas
24
,
122116
(
2017
).
30.

This also facilitates the aforementioned applications to related problems in general relativity,23 as will be discussed in future publications.

31.
K.
Yanagihara
,
I. Y.
Dodin
, and
S.
Kubo
, “
Quasioptical modeling of wave beams with and without mode conversion: II. Numerical simulations of single-mode beams
,”
Phys. Plasmas
26
,
072111
(
2019
).
32.
K.
Yanagihara
,
I. Y.
Dodin
, and
S.
Kubo
, “
Quasioptical modeling of wave beams with and without mode conversion: III. Numerical simulations of mode-converting beams
,”
Phys. Plasmas
26
,
072112
(
2019
).
33.
T.
Notake
,
S.
Kubo
,
T.
Shimozuma
,
H.
Idei
,
Y.
Yoshimura
,
S.
Inagaki
,
K.
Ohkubo
,
S.
Kobayashi
,
Y.
Mizuno
,
S.
Ito
,
Y.
Takita
,
T.
Watari
,
K.
Narihara
,
T.
Morisaki
,
I.
Yamada
,
Y.
Nagayama
,
K.
Tanaka
,
S.
Sakakibara
,
R.
Kumazawa
,
T.
Seki
,
K.
Saito
,
T.
Mutoh
,
A.
Shimizu
,
A.
Komori
, and
LHD Experiment Group,
Optimization of incident wave polarization for ECRH in LHD
,”
Plasma Phys. Controlled Fusion
47
,
531
(
2005
).
34.
S.
Kubo
,
H.
Igami
,
T. I.
Tsujimura
,
T.
Shimozuma
,
H.
Takahashi
,
Y.
Yoshimura
,
M.
Nishiura
,
R.
Makino
, and
T.
Mutoh
, “
Plasma interface of the EC waves to the LHD peripheral region
,”
AIP Conf. Proc.
1689
,
090006
(
2015
).
35.
A.
Iiyoshi
,
A.
Komori
,
A.
Ejiri
,
M.
Emoto
,
H.
Funaba
,
M.
Goto
,
K.
Ida
,
H.
Idei
,
S.
Inagaki
,
S.
Kado
,
O.
Kaneko
,
K.
Kawahata
,
T.
Kobuchi
,
S.
Kubo
,
R.
Kumazawa
,
S.
Masuzaki
,
T.
Minami
,
J.
Miyazawa
,
T.
Morisaki
,
S.
Morita
,
S.
Murakami
,
S.
Muto
,
T.
Mutoh
,
Y.
Nagayama
,
Y.
Nakamura
,
H.
Nakanishi
,
K.
Narihara
,
K.
Nishimura
,
N.
Noda
,
S.
Ohdachi
,
N.
Ohyabu
,
Y.
Oka
,
M.
Osakabe
,
T.
Ozaki
,
B. J.
Peterson
,
A.
Sagara
,
S.
Sakakibara
,
R.
Sakamoto
,
H.
Sasao
,
M.
Sasao
,
K.
Sato
,
M.
Sato
,
T.
Seki
,
T.
Shimozuma
,
M.
Shoji
,
H.
Suzuki
,
Y.
Takeiri
,
K.
Tanaka
,
K.
Toi
,
T.
Tokuzawa
,
K.
Tsumori
,
K.
Tsuzuki
,
K. Y.
Watanabe
,
T.
Watari
,
H.
Yamada
,
I.
Yamada
,
S.
Yamaguchi
,
M.
Yokoyama
,
R.
Akiyama
,
H.
Chikaraishi
,
K.
Haba
,
S.
Hamaguchi
,
M.
Iima
,
S.
Imagawa
,
N.
Inoue
,
K.
Iwamoto
,
S.
Kitagawa
,
J.
Kodaira
,
Y.
Kubota
,
R.
Maekawa
,
T.
Mito
,
T.
Nagasaka
,
A.
Nishimura
,
C.
Takahashi
,
K.
Takahata
,
Y.
Takita
,
H.
Tamura
,
T.
Tsuzuki
,
S.
Yamada
,
K.
Yamauchi
,
N.
Yanagi
,
H.
Yonezu
,
Y.
Hamada
,
K.
Matsuoka
,
K.
Murai
,
K.
Ohkubo
,
I.
Ohtake
,
M.
Okamoto
,
S.
Satoh
,
T.
Satow
,
S.
Sudo
,
S.
Tanahashi
,
K.
Yamazaki
,
M.
Fujiwara
, and
O.
Motojima
, “
Overview of the Large Helical Device project
,”
Nucl. Fusion
39
,
1245
(
1999
).
36.
M.
Osakabe
,
Y.
Takeiri
,
T.
Morisaki
,
G.
Motojima
,
K.
Ogawa
,
M.
Isobe
,
M.
Tanaka
,
S.
Murakami
,
A.
Shimizu
,
K.
Nagaoka
,
H.
Takahashi
,
K.
Nagasaki
,
H.
Takahashi
,
T.
Fujita
,
Y.
Oya
,
M.
Sakamoto
,
Y.
Ueda
,
T.
Akiyama
,
H.
Kasahara
,
S.
Sakakibara
,
R.
Sakamoto
,
M.
Tokitani
,
H.
Yamada
,
M.
Yokoyama
,
Y.
Yoshimura
, and
LHD Experiment Group
, “
Current status of large helical device and its prospect for deuterium experiment
,”
Fusion Sci. Technol.
72
,
199
(
2017
).
37.
For a more rigorous definition of the GO approximation, see, for instance, Ref. 2 or
I. Y.
Dodin
,
A. I.
Zhmoginov
, and
D. E.
Ruiz
, “
Variational principles for dissipative (sub)systems, with applications to the theory of linear dispersion and geometrical optics
,”
Phys. Lett. A
381
,
1411
(
2017
).
38.
D. E.
Ruiz
and
I. Y.
Dodin
, “
Ponderomotive dynamics of waves in quasiperiodically modulated media
,”
Phys. Rev. A
95
,
032114
(
2017
).
39.
H.
Weyl
,
The Theory of Groups and Quantum Mechanics
(
Dover
,
New York
,
1931
).
40.
E.
Wigner
, “
On the quantum correction for the thermodynamic equilibrium
,”
Phys. Rev.
40
,
749
(
1932
).
41.
J. E.
Moyal
, “
Quantum mechanics as a statistical theory
,”
Proc. Cambridge Philos. Soc.
45
,
99
(
1949
).
42.
R. G.
Littlejohn
, “
The semiclassical evolution of wave packets
,”
Phys. Rep.
138
,
193
(
1986
).
43.
S. W.
McDonald
, “
Phase-space representations of wave equations with applications to the eikonal approximation for short-wavelength waves
,”
Phys. Rep.
158
,
337
(
1988
).
44.
R. G.
Littlejohn
and
W. G.
Flynn
, “
Geometric phases in the asymptotic theory of coupled wave equations
,”
Phys. Rev. A
44
,
5239
(
1991
).
45.
I. B.
Bernstein
, “
Geometric optics in space and time varying plasmas. I
,”
Phys. Fluids
18
,
320
(
1975
).
46.
H. L.
Berk
and
D.
Pfirsch
, “
WKB method for systems of integral equations
,”
J. Math. Phys.
21
,
2054
(
1980
).
47.
L.
Friedland
,
G.
Goldner
, and
A. N.
Kaufman
, “
Four-dimensional eikonal theory of linear mode conversion
,”
Phys. Rev. Lett.
58
,
1392
(
1987
).
48.
For Euclidean coordinates, a detailed discussion of the Weyl calculus can be found in Refs. 1 and 24. A generalization to curved metric was considered also in
C.
Gneiting
,
T.
Fischer
, and
K.
Hornberger
, “
Quantum phase-space representation for curved configuration spaces
,”
Phys. Rev. A
88
,
062117
(
2013
) and, to some extent, Ref. 66.
49.
B. S.
DeWitt
, “
Point transformations in quantum mechanics
,”
Phys. Rev.
85
,
653
(
1952
).
50.
J. M.
Domingos
and
M. H.
Caldeira
, “
Self-adjointness of momentum operators in generalized coordinates
,”
Found. Phys.
14
,
147
(
1984
).
51.

Eventually (Sec. V E), we shall choose k(x) such that D(x,k(x)) be small. But that will characterize k(x) rather than the function D itself, which is considered order-one when evaluated on general (x,p).

52.
Reduced codes typically assume that the anti-Hermitian part of the dielectric tensor is small. That is not always the case; for example, see
M. D.
Tokman
,
E.
Westerhof
, and
M. A.
Gavrilova
, “
Wave power flux and ray-tracing in regions of resonant absorption
,”
Plasma Phys. Controlled Fusion
42
,
91
(
2000
) and references therein. However, this subject will not be discussed below.
53.

Strictly speaking, Eq. (1) can always be multiplied by a constant matrix to ensure that the new dispersion operator satisfies the ordering (63) at a given location. Hence, the only nontrivial requirement that we actually impose here is that such DA does not grow too much as the wave propagates.

54.
I. Y.
Dodin
and
N. J.
Fisch
, “
Axiomatic geometrical optics, Abraham-Minkowski controversy, and photon properties derived classically
,”
Phys. Rev. A
86
,
053834
(
2012
).
55.
We use an approach similar to the 3 + 1 approach in general relativity. For example, see
I. Y.
Dodin
and
N. J.
Fisch
, “
Vlasov equation and collisionless hydrodynamics adapted to curved spacetime
,”
Phys. Plasmas
17
,
112118
(
2010
) including the supplementary material there.
56.
For example, see
P. D.
Powell
, “
Calculating determinants of block matrices
,” e-print arXiv:1112.4379.
57.

In principle, D̂A produces effects qualitatively different from those produced by D̂H, so one might argue that the accuracy of its approximation does not necessarily need to equal that of D̂H. Besides, DA is often more abrupt than DH, so retaining VAσ in Eq. (82) may in fact lead to a better model in principle. However, in such cases, the Weyl symbol D of the original dispersion operator is typically known only to the zeroth order (Sec. VI C), and thus so is DA. Hence, retaining first-order corrections the damping coefficient does not seem warranted.

58.
For example, see the derivation of the linear part of the nonlinear Schrödinger equation for a wave in a homogeneous medium in
V. I.
Karpman
,
Non-Linear Waves in Dispersive Media
(
Pergamon Press
,
New York
,
1974
), Sec. 27.
59.
T. J.
Bridges
and
S.
Reich
, “
Multi-symplectic integrators: Numerical schemes for Hamiltonian PDEs that conserve symplecticity
,”
Phys. Lett. A
284
,
184
(
2001
).
60.
L.
Friedland
and
A. N.
Kaufman
, “
Congruent reduction in geometric optics and mode conversion
,”
Phys. Fluids
30
,
3050
(
1987
).
61.
M.
Onoda
,
S.
Murakami
, and
N.
Nagaosa
, “
Hall effect of light
,”
Phys. Rev. Lett.
93
,
083901
(
2004
).
62.
K. Y.
Bliokh
,
A.
Niv
,
V.
Kleiner
, and
E.
Hasman
, “
Geometrodynamics of spinning light
,”
Nat. Photonics
2
,
748
(
2008
).
63.
P.
Gosselin
,
A.
Bérard
, and
H.
Mohrbach
, “
Spin Hall effect of photons in a static gravitational field
,”
Phys. Rev. D
75
,
084035
(
2007
).
64.
C.
Duval
and
T.
Schücker
, “
Gravitational birefringence of light in Robertson–Walker cosmologies
,”
Phys. Rev. D
96
,
043517
(
2017
).
65.
I. Y.
Dodin
, “
Geometric view on noneikonal waves
,”
Phys. Lett. A
378
,
1598
(
2014
).
66.
L. D.
Landau
and
E. M.
Lifshitz
,
The Classical Theory of Fields
(
Pergamon Press
,
New York
,
1971
).
67.
L.
Friedland
, “
Gyroresonant absorption from congruent reduction of an anisotropic pressure fluid model
,”
Phys. Fluids
31
,
2615
(
1988
).
68.
J.
Preinhaelter
and
V.
Kopecky
, “
Penetration of high-frequency waves into a weakly inhomogeneous magnetized plasma at oblique incidence and their transformation to Bernstein modes
,”
J. Plasma Phys.
10
,
1
(
1973
).
69.
C. W.
Misner
,
K. S.
Thorne
, and
J. A.
Wheeler
,
Gravitation
(
Freeman
,
San Francisco
,
1973
).

Supplementary Material