Physica B: Condensed Matter
Volume 404, Issues 8-11, 1 May 2009, Pages 1383-1393




Refraction angles and transmission rates of polarized superluminal radiation

Roman TomaschitzCorresponding Author Contact Information, a, E-mail The Corresponding Author

aDepartment of Physics, Hiroshima University, 1-3-1 Kagami-yama, Higashi-Hiroshima 739-8526, Japan

Received 29 October 2008; 
revised 9 December 2008; 
accepted 22 December 2008. 
Available online 30 December 2008.

Abstract

The propagation of superluminal waves in dispersive media is investigated, in particular the refraction at surfaces of discontinuity in layered dielectrics. The negative mass-square of the tachyonic modes is manifested in the transmission and reflection coefficients, and the polarization of the incident waves (TE, TM, or longitudinal) can be determined from the refraction angles. The conditions for total internal reflection in terms of polarization and tachyon mass are derived. Brewster angles can be used to discriminate longitudinal from transversal incidence. Superluminal transmission through dielectric boundary layers is studied, and the dependence of the intensity maxima on the transversal and longitudinal refractive indices of the layer is analyzed. Estimates of the tachyonic plasma frequency and permittivity of metals are given. The integral version of the tachyonic Maxwell equations is stated, boundary conditions at the surfaces of discontinuity are derived for transversal and longitudinal wave propagation, and singular surface fields and currents are pointed out. The spectral maps of three TeV γ-ray sources associated with supernova remnants, which have recently been obtained with imaging atmospheric Cherenkov detectors, are fitted with tachyonic cascade spectra. The transversal and longitudinal polarization components are disentangled in the spectral maps, and the thermodynamic parameters of the shock-heated ultra-relativistic electron plasma generating the tachyon flux are extracted from the cascade fits.

Keywords: Superluminal wave propagation; Tachyonic Maxwell equations; Boundary conditions for radiation modes with negative mass-square; Transversal and longitudinal refraction; Polarization of tachyonic γ-rays; Tachyonic plasma frequency; Nonthermal cascade spectra

PACS classification codes: 42.25.Bs; 42.25.Gy; 52.25.Mq; 95.30.Gv

Article Outline

1. Introduction
2. Tachyonic Maxwell equations and integral field equations in a permeable medium
3. Superluminal wave fields at the surface of discontinuity of dielectrics and conductors
3.1. Boundary conditions for transversal tachyons
3.2. Longitudinal modes generating singular magnetic field strengths at the interface
3.3. Singular boundary currents and charge densities
4. Tachyon refraction at plane interfaces of dispersive media
4.1. Refraction of superluminal TE and TM waves
4.2. Manifestations of the negative mass-square: longitudinal refraction, Brewster angles, and total internal reflection
4.3. Normal incidence: reflection and transmission of tachyons at a boundary layer
4.3.1. Transversally polarized superluminal modes
4.3.2. Longitudinal reflection and transmission coefficients
5. Polarization components of tachyonic cascades radiated by a shock-heated electron plasma
6. Conclusion
Acknowledgements
References

1. Introduction

We investigate the refraction of superluminal wave modes in dispersive media, at dielectric interfaces and boundary layers. The formalism is developed in analogy to electromagnetic theory [1], even though there are substantial differences due to the negative mass-square of tachyons [2], [3], [4] and [5] and the occurrence of longitudinally polarized modes [6] and [7]. The tachyon mass shows in deflection angles, transmission coefficients, and in longitudinal refraction. The superluminal radiation field is a real Proca field with negative mass-square [8]. In 3D, the wave modes can be written in terms of field strengths and inductions, suggesting a counterpart to Maxwell's equations. The tachyonic Maxwell equations explicitly depend on the scalar and vector potentials, so that the gauge invariance is broken. We derive differential and integral field equations, including the material equations in a dispersive medium relating the tachyonic field strengths and potentials to inductions via frequency-dependent permeabilities.

The field equations, the polarization of superluminal modes, and the transversal and longitudinal Poynting vectors of the tachyon flux in a dispersive medium are discussed in Section 2. The transversal and longitudinal refractive indices for tachyonic wave propagation in dielectrics are introduced, and estimates of the tachyonic conductivity of metals are given. Refraction properties such as deflection and reflection angles are determined by boundary conditions on the tachyon potential, the field strengths, and inductions at the surface of discontinuity. There are two types of boundary conditions depending on the polarization of the wave fields, and singular magnetic field strengths and boundary currents can emerge. This is explained in Section 3.

In Section 4, we study superluminal refraction at a plane interface generated by a discontinuity in the permeabilities. The reflection and refraction angles depend on the polarization of the incident modes, and so do the transmission rates. The Brewster angles for transversal and longitudinal incidence are derived, as well as the conditions for total internal reflection of polarized superluminal radiation. The tachyon flux through a dielectric boundary layer is investigated, in particular the intensity peaks at normal incidence. We calculate the transmission and reflection coefficients, and discuss their frequency dependence and the effect of polarization.

In Section 5, we point out evidence for superluminal γ-rays in the spectral maps of three Galactic TeV sources obtained with the imaging air Cherenkov detectors HESS and MAGIC [9], [10] and [11]. The spectra are fitted with nonthermal tachyonic cascades generated by the shocked electron plasma of the remnants. The transversal and longitudinal polarization components of the cascades are resolved, exhibiting extended spectral plateaus in the high GeV region typical for tachyonic γ-ray emitters, followed by nonthermal power–law slopes at low TeV energies. The spectral break terminating the GeV plateau is compared to the break energies of the cosmic-ray spectrum. In Section 6, we present our conclusions.

2. Tachyonic Maxwell equations and integral field equations in a permeable medium

The tachyonic radiation field in vacuum is a real vector field with negative mass-square, satisfying the Proca equation View the MathML source, subject to the Lorentz condition View the MathML source. mt is the mass of the superluminal Proca field Aμ, and q the tachyonic charge carried by the subluminal electron current jμ=(ρ,j). The mass term is added with a positive sign, and the sign convention for the metric defining the d’Alembertian νν is diag (−1,1,1,1), so that View the MathML source is the negative mass-square of the radiation field [6] and [12]. The above wave equation in conjunction with the Lorentz condition is equivalent to the tachyonic Maxwell equations

(2.1)
View the MathML source
where the field strengths are related to the potentials by E=backward differenceA0-∂A/∂t and View the MathML source. In contrast to electromagnetic theory, the Lorentz condition View the MathML source follows from the field equations and current conservation, View the MathML source, cf. after Eq. (2.3).

In a permeable medium, the potentials and field strengths in the inhomogeneous vacuum equations (2.1) are replaced by inductions, (A0,A)→(C0,C), (E,B)→(D,H), defined by material equations [13] and [14]. We will mostly consider monochromatic waves, View the MathML source, and analogously for the scalar potential, current, charge density, field strengths, and inductions. Fourier amplitudes are denoted by a hat. Maxwell's equations in Fourier space read

(2.2)
View the MathML source
supplemented by material equations

(2.3)
View the MathML source
The inductive potentials View the MathML source as well as the inductions View the MathML source and View the MathML source are related to the primary fields by frequency-dependent dielectric (ε0,ε) and magnetic (μ0,μ) permeabilities. The Fourier amplitudes of the field strengths and potentials are connected by View the MathML source and View the MathML source. We take the divergence of the third equation in Eq. (2.2), and substitute the fourth, to obtain

(2.4)
View the MathML source
Current conservation, View the MathML source, implies the Lorentz condition View the MathML source, or equivalently, View the MathML source in terms of the inductive potentials.

The sign conventions for the coupling of an electron to the Proca field are L=-m/γ+q(A0+A·v) and md(γv)/dt=q(E+v×B), where γ=(1-υ2)-1/2 is the electronic Lorentz factor and q the tachyonic charge carried by the electron, cf. after Eq. (2.14). The primary fields rather than the inductions define the Lorentz force. The potentials are unambiguously determined by the field strengths and the dielectric and magnetic permeabilities. We define polarizations View the MathML source, View the MathML source, and View the MathML source, relating the inductions and field strengths additively as View the MathML source and View the MathML source, and analogously the potentials View the MathML source and View the MathML source. Accordingly, View the MathML source and View the MathML source, with electric susceptibilities κ=ε-1 and κ0=ε0-1. Analogously, View the MathML source and View the MathML source, with magnetic susceptibilities χ=μ-1 and χ0=μ0-1. The permeabilities ε0 and μ0 define the inductive potentials in Eq. (2.3), and are not to be confused with vacuum permeabilities; we use the Heaviside–Lorentz system, so that ε=ε0=1 and μ=μ0=1 in vacuum.

Applying the Gauss theorem to the divergence equations in Eq. (2.2), as well as to the Lorentz condition and the potential definition of View the MathML source, cf. after Eq. (2.3), we find the integral field equations

(2.5)
View the MathML source
Here, dS=ndS is the surface element of a closed surface S, the boundary of a domain V with volume element dV, and n is the surface normal pointing into the interior of the domain of integration V. Current conservation gives

(2.6)
View the MathML source
Applying the Stokes theorem to the rotor equations in Eq. (2.2) and the potential definition of View the MathML source, we obtain

(2.7)
View the MathML source
Here, dS=ndS is the surface element of an oriented open surface SL bounded by the closed loop L, and ds=n×nL ds is the oriented tangent element of the loop. nL is the unit vector tangent to the surface and orthogonal to the loop, pointing to the exterior. n is the unit normal of the surface at the loop, with the same orientation as the surface element dS.

To find the transversal and longitudinal dispersion relations, we use a plane–wave ansatz in the Maxwell equations (2.2) with vanishing charge and current densities: View the MathML source, and analogously for the scalar potential and the field strengths. Here, k=k(ω)k0, where k is the wave number to be determined from the field equations, and k0 is a constant unit vector. k(ω) as well as k0 can be complex. The transversality condition is View the MathML source, and the set of transversal modes is View the MathML source, with amplitude View the MathML source, and similarly for the field strengths and the scalar potential. The dispersion relation determining the transversal wave number is

(2.8)
View the MathML source
and the amplitudes of the transversal field strengths follow from View the MathML source:

(2.9)
View the MathML source
If the product View the MathML source does not vanish, the modes must be longitudinal, View the MathML source, with dispersion relation

(2.10)
View the MathML source
so that View the MathML source. The amplitudes of the longitudinal scalar potential and the field strengths read

(2.11)
View the MathML source
The amplitudes AT,L(k) only need to satisfy the transversality/longitudinality condition, that is, the first equation in Eqs. (2.9) and (2.11), respectively.

The time-averaged tachyonic energy flux carried by homogeneous modes (with real wave vector k0) in a dispersive and non-absorptive medium is

(2.12)
View the MathML source
where View the MathML source stands for View the MathML source. As above, the superscripts T and L refer to the transversal and longitudinal flux components. The frequency-dependent permeabilities of a dispersive transparent medium are real in the absence of energy dissipation. The flux vectors (2.12) apply even for negative permeabilities [15], [16] and [17], provided that we restrict to a frequency range where the squared wave numbers (2.8) and (2.10) are positive. They are obtained by substituting the polarized plane waves (2.9) and (2.11) into the Poynting vector View the MathML source, and by performing a time average [18].

The transversal and longitudinal phase velocities read View the MathML source, and the tachyonic group velocities are View the MathML source, so that View the MathML source and View the MathML source. In the case of vacuum permeabilities, ε(0)=μ(0)=1, the transversal and longitudinal velocities coincide, and we find ω=mtγt, where View the MathML source denotes the tachyonic Lorentz factor and υgr>1 is the superluminal group velocity. The refractive index nT,L=kT,L/ω, defined as dimensionless inverse phase velocity, differs for transversal and longitudinal modes:

(2.13)
View the MathML source
At high frequencies ωnot double greater-than signmt (with View the MathML source [6] and [18] and planck constant over two pi=c=1), we can approximate View the MathML source. The longitudinal index, View the MathML source, has no electromagnetic analog. In the low-frequency regime ωdouble less-than signmt, the refractive indices are frequency dependent even if the permeabilities stay constant, View the MathML source and View the MathML source.

To estimate the dielectric permeability for tachyonic wave propagation, we start with a monochromatic superluminal mode, View the MathML source, in a dispersive and possibly dissipative medium. This mode generates the current View the MathML source, View the MathML source, where σ(ω) is the tachyonic conductivity of the medium, cf. Eq. (2.14). The charge distribution View the MathML source is found by means of current conservation, cf. after Eq. (2.4). We substitute this current into the field equations (2.2) with vacuum permeabilities ε(0)=μ(0)=1, and absorb current and charge density by introducing the permittivity εσ(ω)=1+iσ(ω)/ω. In this way, we can write the inhomogeneous field equations (2.2) as View the MathML source and View the MathML source; the homogeneous Maxwell equations remain unchanged. The polarization vector is View the MathML source, cf. after Eq. (2.4), and the London equation View the MathML source applies. The dispersion relations for the transversal and longitudinal modes read as in Eqs. (2.8) and (2.10), with ε replaced by εσ and ε0=μ(0)=1.

We consider a tachyonic conductivity

(2.14)
View the MathML source
where View the MathML source is the tachyonic plasma frequency, and ne the electron density of the medium. This is based on Drude's damped oscillator model, View the MathML source, with View the MathML source, and E as above [14]. ω0 is the characteristic binding frequency of the electronic oscillators, q the tachyonic charge, and γ0 is the damping constant related to the tachyonic resistivity ρt by View the MathML source. We solve this equation in dipole approximation, neglecting the spatial dependence of the Fourier components View the MathML source, so that σ(ω) in Eq. (2.14) follows from View the MathML source, View the MathML source, and View the MathML source. We note View the MathML source, where View the MathML source is the electromagnetic plasma frequency, and αq/αe≈1.4×10-11 the ratio of tachyonic and electric fine structure constants. In the Heaviside–Lorentz system, αe=e2/(4πplanck constant over two pic)≈1/137 and αq=q2/(4πplanck constant over two pic)≈1.0×10-13. The tachyon–electron mass ratio is mt/m≈1/238; αq and mt are estimated from hydrogenic Lamb shifts [6]. We may thus conclude that the tachyonic conductivity View the MathML source is by a factor of 10−11 smaller than the electric counterpart. However, εσ(ω)-1 can still be of order one.

To demonstrate this, we consider a free electron gas, ω0=γ0=0, so that the tachyonic conductivity (2.14) simplifies to View the MathML source, and the induced permittivity is View the MathML source. We may write the electromagnetic plasma frequency as View the MathML source, where small lambda with strokee is the reduced electronic Compton wavelength, αe the electric fine structure constant, and ne the electron density as above. We thus find View the MathML source, and the tachyonic counterpart ωp≈3.74×10-6ωp,em. Metallic electron densities of 1022−1023 cm−3 result in a tachyonic plasma frequency ωp in the 10−5 eV range. Thus, even though the tachyonic conductivity is much smaller than the electromagnetic one at the same frequency, the tachyonic permittivity εσ(ω) becomes noticeably different from 1 for frequencies comparable to ωp. Finally, small frequencies of the order of 10−5 eV do not imply large wavelengths. The tachyonic wavelength is λT,L=2π/kT,L, with wave numbers defined in Eqs. (2.8) and (2.10). The maximal transversal/longitudinal wavelength in the medium is thus View the MathML source and View the MathML source, attained in the zero frequency limit, where the permeabilities are of order one and View the MathML source is the tachyonic Compton wavelength.

3. Superluminal wave fields at the surface of discontinuity of dielectrics and conductors

To study tachyon refraction at the interface of two media with different permeabilities, we have to specify the boundary conditions. To this end, we split the wave fields into regular and singular parts, writing the scalar induction and the magnetic field strength as

(3.1)
View the MathML source
The same decomposition is used for the scalar and vector potentials View the MathML source, the electric field strength View the MathML source, and the vectorial inductions View the MathML source, View the MathML source, and View the MathML source. Here, θ(S) is the Heaviside step function, and S(x)=0 is the surface separating the media defined by permeabilities (ε0,1,ε1,μ0,1,μ1) and (ε0,2,ε2,μ0,2,μ2), cf. Eq. (2.3), which can be frequency dependent (dispersive) and complex (dissipative). The subscripts 1 and 2 on the fields and permeabilities refer to the respective media; in the decomposition (3.1) of the regular field components, it is understood that domain S(x)>0 is occupied by medium 2, so that View the MathML source for S>0, and View the MathML source applies in domain S<0 where medium 1 is located. The normal vector, n:=backward differenceS/|backward differenceS|, thus points into the interior of medium 2. The singular part of the field, if any, is a distribution supported on the boundary surface S=0, typically containing a Dirac function δ(S) as factor. In contrast to the regular components in Eq. (3.1), the singular field strengths and inductions do not show in the material equations (2.3). In Sections 3.1 and 3.2, we assume vanishing charge and current densities in the field equations, as well as the absence of a singular surface current at the interface. In Section 3.3, we consider inhomogeneous boundary conditions consistent with currents.

3.1. Boundary conditions for transversal tachyons

We start by substituting ansatz (3.1) into the field equations (2.2) (with zero current and charge density) and the Lorentz condition, and assume that the fields subjected to rotors and divergences have no singular part. That is, we try to solve under the condition that View the MathML source, View the MathML source, View the MathML source, View the MathML source, and View the MathML source are regular at the interface, obtaining five relations to be satisfied at the boundary S=0,

(3.2)
View the MathML source


(3.3)
View the MathML source


(3.4)
View the MathML source
where Eqs. (3.2) and (3.4) can be combined to give

(3.5)
View the MathML source
Assuming the potentials View the MathML source and View the MathML source to be non-singular, we find, from the potential representation of View the MathML source and View the MathML source, cf. after Eq. (2.3),

(3.6)
View the MathML source
All fields are regular at the boundary surface, with exception of View the MathML source, whose singular part is obtained from (3.2) or (3.4). View the MathML source does not vanish except for a special polarization (TE waves), but no singular contribution can occur in the transversal flux vector (2.12), as View the MathML source is identically zero in both media, cf. Eq. (2.9). The boundary conditions on transversal modes are defined by Eqs. (3.3), (3.5) and (3.6).

3.2. Longitudinal modes generating singular magnetic field strengths at the interface

We proceed analogously to the transversal case, but now assuming all fields to be regular at the boundary except for the magnetic field View the MathML source. On substituting ansatz (3.1) into the field equations (2.3) and (2.4) and the Lorentz condition, cf. after Eq. (2.4), we find

(3.7)
View the MathML source


(3.8)
View the MathML source
The potential representations of View the MathML source and View the MathML source give

(3.9)
View the MathML source


(3.10)
View the MathML source
By combining Eqs. (3.8) and (3.9), we obtain

(3.11)
View the MathML source
Finally, the divergence equation, View the MathML source, results in

(3.12)
View the MathML source
where we used (3.9). Eqs. (3.7), (3.10), (3.11) and (3.12) constitute the boundary conditions for longitudinal modes. The regular longitudinal View the MathML source field vanishes identically in both media, cf. Eq. (2.11), but a singular surface field View the MathML source emerges, which, however, does not affect the longitudinal flux vector (2.12), since all other field strengths, potentials, and inductions are regular.

The transversal and longitudinal boundary conditions derived in Sections 3.1 and 3.2 apply at the interface of two media with different permeabilities discontinuous at the interface. They unambiguously determine the refractive properties of tachyons, such as deflection and reflection angles, and assure transmission and reflection ratios consistent with energy conservation in non-absorptive media, cf. Section 4.

3.3. Singular boundary currents and charge densities

In the case of non-vanishing charge and current densities in field equations (2.2), the transversal and longitudinal boundary conditions derived in Sections 3.1 and 3.2 become inhomogeneous due to singular surface currents. As in (3.1), we split current and charge density into a regular and singular part,

View the MathML source


(3.13)
View the MathML source
The singular surface charge and current, View the MathML source and View the MathML source, are supported at the interface S=0. The subscripts 1 and 2 refer to the respective medium as defined after Eq. (3.1). We substitute ansatz (3.13) into the continuity equation, cf. after Eq. (2.4), to find the boundary condition required by current conservation,

(3.14)
View the MathML source
Three of the transversal boundary conditions in Section 3.1 have to be modified in the presence of currents. Condition (3.2) is replaced by

(3.15)
View the MathML source
the third condition in Eq. (3.3) reads

(3.16)
View the MathML source
and condition (3.5) becomes inhomogeneous as well:

(3.17)
View the MathML source
As for the longitudinal boundary conditions in Section 3.2, there are two changes. The first condition in Eq. (3.7) is replaced by Eq. (3.15) with View the MathML source, and the third condition in Eq. (3.7) by Eq. (3.16). It is easy to check that boundary condition (3.16) is consistent with current conservation (3.14).

4. Tachyon refraction at plane interfaces of dispersive media

We take the z=0 plane as the interface S separating medium 2 in the upper half-space z>0 from medium 1 in the lower half-space. The media are defined by frequency-dependent permeabilities (ε0,1,ε1,μ0,1,μ1) and (ε0,2,ε2,μ0,2,μ2), respectively, cf. Eq. (2.3) and after Eq. (3.1). We consider a tachyonic plane wave, cf. after Eq. (2.7), incident from the lower half-space upon the interface, the e1,2 plane. The wave number kin of this incoming transversal or longitudinal wave is defined by dispersion relation (2.8) or (2.10), with permeabilities carrying subscript 1. Part of the wave is reflected back into medium 1, and the wave number kre of the reflected wave coincides with kin. The wave number ktr of the wave transmitted into the upper half-space is determined by the permeabilities of medium 2. The wave vectors of the respective modes are denoted by kin=kink0,in, kre=krek0,re, and ktr=ktrk0,tr, where the zero subscript indicates unit vectors. We assume the incoming plane wave to be homogeneous, so that k0,in is a real unit vector; the wave numbers kin,tr can be complex. Since k0,in is real, the unit wave vector k0,re of the reflected wave is real too, as shown below. The transmitted wave is in general inhomogeneous if the wave number in medium 1 or 2 is complex, so that k0,tr is a complex unit vector, View the MathML source. We adopt the convention Re(kin,tr)>0, since wave numbers are only defined as squares by the dispersion relations. If medium 1 in the lower half-space has real permeabilities, the vacuum for instance, then the incident wave number kin is real.

We choose the incoming real unit wave vector k0,in in the e1,3 plane. (The ei are coordinate unit vectors.) The normal vector of the interface is e3, pointing into medium 2. A convenient angular parametrization of the wave vectors is

(4.1)
View the MathML source
The incoming wave moves through the lower half-space towards the interface at z=0, so that View the MathML source and View the MathML source are positive, kin=kre, View the MathML source, and View the MathML source. The angles View the MathML source and View the MathML source in Eq. (4.1) are in general complex, and θre=π-θin. The boundary conditions at z=0 can only be satisfied if the phase factors eik·x of the three waves coincide at the boundary. This requires kin·e1=kre·e1=ktr·e1, and the same for e2 at the interface. (The wave vectors are orthogonal to e2, as k0,in is by definition.) Hence, View the MathML source, which is the tachyonic counterpart to Snell's reflection law [19] and [20].

The refractive indices of medium 1 and 2 are denoted by n1,2=kin,tr/ω, cf. Eq. (2.13), and their ratio by View the MathML source. This applies to transversal as well as longitudinal indices, e.g., View the MathML source. The transversal/longitudinal refraction law can thus be written as View the MathML source. In the high-frequency regime ωnot double greater-than signmt, the transversal refractive index ratio simplifies to View the MathML source, and the longitudinal one to View the MathML source, cf. after Eq. (2.13). In the low-frequency limit, ωdouble less-than signmt, the refraction angle θtr is determined by View the MathML source or View the MathML source. If we consider dielectrics with μ(0)=ε0=1, the longitudinal refractive index ratio View the MathML source at low frequencies is just the inverse of View the MathML source at high frequency. The refraction law can thus be used to discriminate between transversal and longitudinal polarization.

4.1. Refraction of superluminal TE and TM waves

We first consider tachyonic TE waves, so that the incoming mode View the MathML source is linearly polarized, with amplitude orthogonal to the plane of incidence generated by the normal vector e3 of the boundary and the wave vector k0,in in the e1,3 plane. Thus, View the MathML source, cf. Eq. (2.9). The boundary conditions stated in Section 3.1 are satisfied by the reflected and transmitted waves, which are likewise polarized in the e2 direction, so that the respective modes are Eree2eikre·x and Etre2eiktr·x. The wave numbers of the incident and reflected waves are defined by the transversal dispersion relation (2.8) of medium 1 with permeabilities (ε0,1,ε1,μ0,1,μ1), cf. Eq. (2.3). The wave number of the transmitted wave is calculated with the permeabilities (ε0,2,ε2,μ0,2,μ2) of medium 2 in the upper half-space. The boundary conditions (3.3), (3.5) and (3.6) give, if combined with Snell's law as stated above, two independent relations for the three amplitudes:

(4.2)
View the MathML source
obtained via View the MathML source and View the MathML source. We write this in amplitude ratios by means of Eq. (4.1):

(4.3)
View the MathML source
Using Snell's law, we parametrize by the incidence angle, substituting View the MathML source. Even though we employ here and in the following only two boundary conditions, the remaining ones are satisfied as well, by virtue of the above refraction laws.

We turn to superluminal TM waves, where the electric field is linearly polarized parallel to the plane of incidence. It is convenient to write the boundary conditions in terms of the magnetic field, by way of View the MathML source, where View the MathML source is orthogonal to the plane of incidence, cf. Eq. (2.9). Accordingly, View the MathML source, and analogously for the reflected and transmitted modes. The boundary conditions for transversal modes again give two independent relations among the three amplitudes,

(4.4)
View the MathML source
Here, we used the same two boundary conditions as for TE modes. The amplitude ratios read

(4.5)
View the MathML source
which differ from the TE ratios just by an interchange of μ1ktr and μ2kin, and we substitute View the MathML source stated after Eq. (4.3) to parametrize by the incidence angle.

The energy flux (i.e., the incident, reflected, or transmitted energy per unit time and unit surface area) carried by a superluminal mode with real unit wave vector k0 is FT,L:=|left angle bracketST,Lright-pointing angle bracket||k0·n|. The transversal and longitudinal flux vectors left angle bracketST,Lright-pointing angle bracket are defined in Eq. (2.12), with the respective incident, reflected, or transmitted wave substituted. As for the transmitted wave, we assume the two media to be non-dissipative, so that k0,tr is real. The superscripts T and L denote transversal and longitudinal waves, the latter are studied in Section 4.2. The transversal and longitudinal reflection and transmission coefficients are defined by the flux ratios

(4.6)
View the MathML source
where the subscripts indicate the respective fields (incident, reflected, or transmitted) to be substituted into the flux vector. Energy conservation requires RT,L+TT,L=1. (We do not define a transmission coefficient for dissipative media.) By making use of the transversal flux vector (2.12), we find the reflection and transmission ratios of tachyonic TE and TM modes as

(4.7)
View the MathML source
with the amplitude ratios (4.3) and (4.5) substituted. It is easy to check that energy is conserved, which suggests that we have got the boundary conditions in Section 3.1 right.

At normal incidence, θin=θtr=0, θre=π, the amplitude ratios (4.3) and (4.5) reduce to

(4.8)
View the MathML source
where View the MathML source is the ratio of the transversal refractive indices of medium 1 and 2, and View the MathML source, cf. after Eq. (4.1). In the high-frequency regime ωnot double greater-than signmt, we approximate View the MathML source, cf. after Eq. (2.13), and find the reflection coefficients for TE and TM waves as

(4.9)
View the MathML source
At low frequency ωdouble less-than signmt, we substitute View the MathML source in Eq. (4.8):

(4.10)
View the MathML source
In this limit, the reflected fraction of the transversal tachyon flux is determined by the magnetic permeabilities only. Transversal refraction will further be discussed after Eq. (4.16), together with the longitudinal reflection coefficients derived in Section 4.2.

4.2. Manifestations of the negative mass-square: longitudinal refraction, Brewster angles, and total internal reflection

We start with a longitudinal incident mode, View the MathML source, and use analogous notation for the reflected and transmitted fields. All wave vectors lie in the e1,3 plane. The permeabilities of media 1 and 2 are labeled as indicated before Eq. (4.2). The boundary conditions (3.7), (3.10), (3.11) and (3.12) give two independent relations for the amplitudes,

(4.11)
View the MathML source
derived from View the MathML source and View the MathML source. The amplitude ratios read accordingly

(4.12)
View the MathML source
with View the MathML source defined after Eq. (4.3). The singular surface magnetic field View the MathML source is calculated via Eq. (3.9):

(4.13)
View the MathML source
The longitudinal reflection and transmission coefficients defined in Eqs. (2.12) and (4.6) are

(4.14)
View the MathML source
where we substitute the ratios (4.12). The transmission coefficient applies for real permeabilities, energy being conserved in non-absorptive media, RL+TL=1.

At normal incidence, θin=θtr=0, the ratios (4.12) simplify,

(4.15)
View the MathML source
where View the MathML source, View the MathML source, and View the MathML source is the quotient nL,2/nL,1 of the longitudinal refractive indices, cf. Eq. (2.13). At high frequency ωnot double greater-than signmt, we find View the MathML source, and for ωdouble less-than signmt the refractive index becomes View the MathML source, so that the longitudinal reflection coefficients read in the respective limit

(4.16)
View the MathML source
If we consider vacuum permeabilities in medium 1, ε(0),1=μ(0),1=1, and a dielectric permeability ε2 different from one in medium 2 (with ε0,2=μ(0),2=1), we find View the MathML source and a finite reflectivity, View the MathML source, for longitudinal modes in the low-frequency regime. The same finite reflectivity applies for tachyonic TE and TM modes, but in the opposite limit, ωnot double greater-than signmt, with different ε2(ω), cf. Eq. (4.9). The transversal reflection coefficients Eq. (4.10) valid for ωdouble less-than signmt vanish, as the indicated leading order of the frequency expansion is independent of ε2.

We consider two other special cases. First, the case where the refracted wave vector is orthogonal to the reflected wave, θre-θtr=π/2, so that θtr+θin=π/2, and thus View the MathML source and View the MathML source. This Brewster incidence angle, View the MathML source, follows from the refraction law stated after Eq. (4.1); View the MathML source denotes the transversal or longitudinal refractive index ratio, cf. after Eqs. (4.8) and (4.15). The refractive indices are assumed to be real in both media. As for TM waves, we find Hre=0 in Eq. (4.5), provided that the magnetic permeabilities μ1 and μ2 of the two media coincide. Similarly for longitudinally polarized waves, Ere=0 in Eq. (4.12), provided that ε0,1=ε0,2. At this incidence angle, the energy of a tachyonic TM wave or a longitudinal wave is fully transmitted. If the incident transversal wave is elliptically polarized (being a complex linear combination of TE and TM waves), the reflected wave is a TE wave linearly polarized orthogonal to the plane of incidence. If we do not require μ1=μ2, and define the incidence angle by Hre=0, we find

(4.17)
View the MathML source
The longitudinal incidence angle defined by Ere=0 is likewise given by Eq. (4.17), with View the MathML source replaced by View the MathML source, and View the MathML source by View the MathML source, cf. Eqs. (4.8) and (4.15).

The second special case is total internal reflection, which requires real wave numbers and incidence angles satisfying View the MathML source, so that View the MathML source defined after Eq. (4.3) is zero or imaginary with View the MathML source to ensure damping. (More generally, the damping condition for a transmitted wave in medium 2 is View the MathML source.) Thus k0,tr is a complex unit vector, even though the permeabilities in both media are real; its e1 component View the MathML source is found via Snell's law, cf. after Eq. (4.1). The reflection coefficients RT,L in Eqs. (4.7) and (4.14) are equal to 1, so that the incident flux is totally reflected. The refracted wave in medium 2 is exponentially damped along the z axis, and no energy is transmitted. Internal reflection can only occur if View the MathML source, that is, medium 1 must be optically thicker than medium 2 for transversal or longitudinal modes. A third special case, normal incidence on a boundary layer of finite thickness separating two dielectric media, is discussed in the next subsection.

4.3. Normal incidence: reflection and transmission of tachyons at a boundary layer

We consider three dispersive media separated by parallel boundary planes z=0 and z=h. Medium 1 lies in the lower half-space, medium 2 is a layer of thickness h located in 0<z<h, and medium 3 fills the half-space z>h. The respective permeabilities and refractive indices are denoted by subscripts, ε1,2,3, etc., cf. after Eq. (3.1). The layer is hit by a tachyonic plane wave propagating in the lower half-space orthogonally incident upon the z=0 plane. To satisfy the boundary conditions at the two interfaces, we start with the ansatz

(4.18)
View the MathML source
where the Ei are fields in the respective media. The notation is explained at the beginning of Section 4 and in the previous two subsections. The transversal field strengths View the MathML source are complex linear combinations of the linearly polarized fields Eie1 and Eie2, and the longitudinal ones read Eie3. Thus, E1ek is a superposition of the incoming and reflected waves in medium 1, the wave number in this medium being kin. Similarly, E2ek is composed of the wave transmitted into medium 2 and a second wave arising by reflection at the second interface z=h. Finally, E3ek is the outgoing wave in medium 3. ktr and kout are the wave numbers in medium 2 and 3, respectively. All wave numbers have a positive real part, so that the negative sign in the exponents of the reflected waves implies the unit wave vector −e3. We take the amplitude of the incident wave Ein as input parameter; the remaining four amplitudes are obtained from the boundary conditions at the two interfaces.

4.3.1. Transversally polarized superluminal modes

The boundary conditions for transversal tachyons read, cf. Sections 3.1 and 4.1,

View the MathML source


(4.19)
View the MathML source
The third and fourth of these equations lead to the amplitude ratios

(4.20)
View the MathML source
On substituting αT into the first and second equations in Eq. (4.19), we find

View the MathML source


(4.21)
View the MathML source
Reflection and transmission coefficients are defined as in Eqs. (4.6) and (4.7), View the MathML source and View the MathML source, so that the transversal ratios read

(4.22)
View the MathML source
As for the transmission coefficient, we assume real permeabilities and wave numbers in media 1 and 3, so that damping can only occur in the boundary layer, that is medium 2. If all three media are dielectrics, energy conservation applies, RT+TT=1. In this case, the coefficients are periodic in h, the thickness of the boundary layer, and the extrema of RT and TT are determined by sin(2hktr(ω))=0, the variation being with respect to the layer thickness, RT/∂h=0. Thus the intensity minima and maxima occur at frequencies where the transversal wave number ktr=ωnT,2 is an integer multiple of π/(2h); nT,2(ω) is the transversal refractive index (2.13) of the layer. We find, for cos(2hktr)=±1,

(4.23)
View the MathML source
where μi denotes the magnetic permeability and nT,i the transversal refractive index of the respective medium, cf. after Eq. (2.13). In the high-frequency limit ωnot double greater-than signmt, we substitute View the MathML source in Eq. (4.23),

(4.24)
View the MathML source
and in the low-frequency band ωdouble less-than signmt, we use View the MathML source to obtain

(4.25)
View the MathML source
The meaning of these extremal transversal reflection coefficients and their longitudinal counterpart is discussed after Eq. (4.32).

4.3.2. Longitudinal reflection and transmission coefficients

We proceed as in the previous case of transversal normal incidence, starting with the longitudinal boundary conditions, cf. Sections 3.2 and 4.2,

(4.26)
View the MathML source
The third and fourth of these equations give

(4.27)
View the MathML source
We substitute αL into the first and second equations in Eq. (4.26),

View the MathML source


(4.28)
View the MathML source
to find the longitudinal reflection and transmission coefficients, cf. Eq. (4.14),

(4.29)
View the MathML source
If all permeabilities are real, energy is conserved, RL+TL=1, and these coefficients are periodic in the layer thickness h. The intensity extrema of RL and TL are defined by the longitudinal wave number ktr=ωnL,2(ω) in the layer, cf. Eqs. (2.10) and (2.13), occurring at frequencies solving cos(2hktr(ω))=±1, analogously to Eq. (4.23):

(4.30)
View the MathML source
where nL,i is the longitudinal refractive index of the respective medium. At high frequencies, ωnot double greater-than signmt, we substitute View the MathML source to find the extremal reflection coefficients for longitudinal tachyons,

(4.31)
View the MathML source
In the low-frequency regime, we approximate View the MathML source so that Eq. (4.30) simplifies to

(4.32)
View the MathML source
For example, we may set all permeabilities equal to one apart from the permittivity ε2 of the layer. At high frequency, the longitudinal flux is almost totally transmitted since the leading order of the reflection coefficient vanishes, View the MathML source. At low frequency, we still have View the MathML source (that is, for wave numbers ktr=(l+1/2)π/h with integer l), but there is a non-vanishing fraction View the MathML source of the incident flux reflected at frequencies satisfying ktr(ω)=lπ/h. This is just the opposite of the transversal case in Eqs. (4.24) and (4.25), where View the MathML source at high frequencies, whereas View the MathML source for ωdouble less-than signmt. More generally, there is a symmetry in the reflection coefficients (4.23) and (4.30) with regard to the interchange ε0μ, εμ0, which is apparent in the asymptotic limits (4.24) and (4.31) as well as Eqs. (4.25) and (4.32). However, the extremal frequencies defined by the zeros of sin(2hktr(ω)) differ for transversal and longitudinal modes, unless the wave numbers (2.8) and (2.10) coincide in the layer.

5. Polarization components of tachyonic cascades radiated by a shock-heated electron plasma

Fig. 1, Fig. 2 and Fig. 3 depict tachyonic cascade spectra of the TeV γ-ray sources HESS J1837−069, HESS J1834−087, and HESS J1813−178. The cascades are plots of the E2-rescaled flux densities,

(5.1)
View the MathML source
where d is the distance to the source, and left angle bracketpT,L(ω=E/planck constant over two pi)right-pointing angle bracket the transversal/longitudinal tachyonic spectral density averaged over a nonthermal electronic power-law distribution, View the MathML source [21]. As for the latter, α is the electronic power-law index, the ultra-relativistic electronic Lorentz factors range in an interval γ1less-than-or-equals, slantγ<∞, γ1not double greater-than sign1, and the exponential cutoff is related to the electron temperature by β=mc2/(kT). A thermal Maxwell–Boltzmann distribution corresponds to α=−2 and γ1=1. The least-squares fit is performed with the unpolarized flux density dNT+L=dNT+dNL, and then split into transversal and longitudinal radiation components. The details of the spectral fitting have been explained in Ref. [22]. The parameters of the electron distributions dρα,β generating the tachyonic cascades are listed in Table 1. The cutoff parameter β in the Boltzmann factor could not be extracted from the presently available flux points, in contrast to the power–law index α and the lower edge γ1 of Lorentz factors. The power–law slope ultimately terminates in exponential decay, cf. the spectral map of HESS J1825−137 in Fig. 5 of Ref. [7]. As for the spectral fits in Fig. 1, Fig. 2 and Fig. 3, there is no downward bend yet in the presently accessible TeV range.



Full-size image

Fig. 1. Spectral map of the TeV γ-ray source HESS J1837−069 associated with the pulsar wind nebula AX J1838.0−0655. Flux points from Ref. [9]. The solid line depicts the unpolarized differential tachyon flux dNT+L/dE rescaled with E2 tachyon flux dNT+L/dE rescaled with E2, cf. (5.1). The transversal (dot-dashed) and longitudinal (double-dot-dashed) flux densities dNT,L/dE add up to the total unpolarized flux cascade ρ1=T+L generated by a nonthermal electron population. The cascade admits a power–law slope E1-α with electron index α≈1.4. A spectral break at View the MathML source is visible as edge in the longitudinal component, where View the MathML source is the tachyon mass [6] and [18]. The least-squares fit is based on the unpolarized tachyon flux T+L, cf. Table 1.


Full-size image

Fig. 2. Spectral map of the extended TeV source HESS J1834−087 in supernova remnant W41. HESS data points from Ref. [9], MAGIC points from Ref. [10]. Notation as in Fig. 1. The parameters of the shock-heated electron plasma are listed in Table 1. The spectral break in the longitudinal flux component (L) of the cascade occurs at 0.80 TeV. The distance estimate of this source is 4 kpc, and its electron index is 1.9, quite similar to the TeV source in Fig. 3 at a comparable distance. The power-law slope is steeper than of HESS J1837−069 at 6.6 kpc, cf. Fig. 1; there is no interstellar absorption of the tachyon flux [22].


Full-size image

Fig. 3. Spectral map of the TeV γ-ray source HESS J1813−178 coincident with supernova remnant W33. HESS flux points from Ref. [9], MAGIC points from Ref. [11]. Notation as in Fig. 1. The tachyonic cascade ρ1=T+L is generated by the ultra-relativistic electron plasma of the remnant, cf. Table 1. The spectral break at View the MathML source separates the power-law slope from the extended GeV plateau distinctive of tachyonic cascade spectra [21], [23] and [24].


Table 1.

Parameters of the nonthermal electron plasma generating the tachyonic cascade spectra of the TeV γ-ray sources in Fig. 1, Fig. 2 and Fig. 3.

α γ1 View the MathML source d (kpc) ne
HESS J1837−069 1.4 3.3×108 7.5×10−4 6.6 1.9×1048
HESS J1834−087 1.9 3.7×108 6.0×10−4 4 5.5×1047
HESS J1813−178 1.9 3.0×109 3.1×10−4 4.5 3.6×1047

α is the electron index, and γ1 the lower threshold Lorentz factor of the ultra-relativistic electron populations, cf. after Eq. (5.1). View the MathML source determines the amplitude of the tachyon flux, from which the electron count ne is inferred at the indicated distance d, cf. Refs. [[27], [28] and [29]]. The parameters α, γ1, and View the MathML source are extracted from the χ2-fit T+L in the figures. The amplitude View the MathML source is related to the electron number by View the MathML source, cf. Ref. [7].


Fig. 1 shows the TeV spectral map of the unidentified TeV source HESS J1837−069 [9], coincident with the pulsar wind nebula AX J1838.0−0655. The distance estimate to this X-ray nebula is 6.6 kpc, by association with a nearby cluster of red supergiants. Fig. 2 depicts the spectral fit to the extended TeV source HESS J1834−087, associated with the shell-type supernova remnant W41, cf. Refs. [9] and [10]. The kinematic distance estimate of W41 is 4 kpc. Spectral plateaus in the MeV to GeV range occur frequently in spectral maps of both thermal and nonthermal TeV sources, and can easily be fitted with tachyonic cascade spectra, in contrast to electromagnetic or hadronic radiation models. Thermal spectra of γ-ray binaries such as binary pulsars and microquasars are studied in Refs. [12] and [23], and a thermal cascade fit of a γ-ray quasar is performed in Ref. [24]. The shocked electron plasma of supernova remnants requires nonthermal electron densities. Fig. 3 shows a nonthermal cascade fit to the TeV source HESS J1813−178, located in the vicinity of the H II region W33 [9] and [11] at a distance of 4.5 kpc. The lower edge of Lorentz factors of the electron plasma is γ1≈3.0×109, inferred from the cascade fit. The corresponding electron and proton energies are View the MathML source and View the MathML source. These lower bounds on the energy of the radiating source particles are to be compared to the spectral breaks in the cosmic-ray spectrum at 1015.5 and 1017.8 eV, dubbed knee and second knee, respectively [25] and [26]. The bounds are one order lower for the sources in Fig. 1 and Fig. 2, cf. Table 1, but in all three remnants the lower bound on the proton energy is close to the second knee.

6. Conclusion

We have studied the refraction of superluminal radiation at dielectric interfaces, in particular the refraction angles for transversal and longitudinal incidence, and the dependence of the transmission and reflection coefficients on the polarization of the incident radiation modes. Speed and energy of tachyonic quanta are related by View the MathML source in vacuum, cf. after Eq. (2.12). At γ-ray energies, their speed is close to the speed of light, the basic difference to electromagnetic radiation being the longitudinally polarized flux component. The polarization of tachyons can be determined from the refraction angles at dielectric interfaces, cf. after Eq. (4.1), or from the reflection coefficients, which greatly differ for transversal and longitudinal modes, cf. after Eq. (4.32). We performed tachyonic cascade fits to the γ-ray spectra of the TeV sources in Fig. 1, Fig. 2 and Fig. 3, and disentangled the transversal and longitudinal flux components.

Shocked electron plasmas generate nonthermal γ-ray cascades typical for supernova remnants and pulsar wind nebulae. The characteristic feature is the extended spectral plateau at GeV energies, followed by a steep but barely bent spectral slope in the low TeV range, assuming a double-logarithmic and E2-rescaled flux representation as in Fig. 1, Fig. 2 and Fig. 3. The spectral maps discussed here are to be compared to the unpulsed γ-ray spectrum of the Crab Nebula, cf. Fig. 1 in Ref. [18], the spectral map of supernova remnant RX J1713.7−3946 in Fig. 2 of Ref. [18], the spectra of HESS J1825−137 and TeV J2032+4130 in Figs. 5 and 6 of Ref. [7], and the extended γ-ray cascade of supernova remnant W28 in Fig. 4 of Ref. [22]. All these spectra show GeV plateaus followed by straight or slightly curved power–law slopes. Traditional radiation mechanisms such as inverse Compton scattering or proton–proton scattering followed by pion decay fail to reproduce the extended plateaus in the spectral maps, a fact often concealed by compression in broadband maps. By contrast, tachyonic cascades provide excellent fits to the GeV plateaus and power–law slopes. The latter are a signature of shock-heated electron plasmas, and absent in the spectra of thermal γ-ray sources like TeV blazars [30] and [31], where the plateaus terminate in exponential decay without power–law transition. The spectral breaks at the join of the spectral plateaus and the power–law slopes are determined by the lower threshold Lorentz factors of the nonthermal source populations in the remnants. These Lorentz factors can be extracted from the spectral fits [32], and suggest that TeV γ-ray sources in Galactic supernova remnants are capable of accelerating protons to energies above the spectral break at 1017.8 eV in the cosmic-ray spectrum.

Acknowledgments

The author acknowledges the support of the Japan Society for the Promotion of Science. The hospitality and stimulating atmosphere of the Centre for Nonlinear Dynamics, Bharathidasan University, Trichy, and the Institute of Mathematical Sciences, Chennai, are likewise gratefully acknowledged. I also thank the referee for useful suggestions regarding substance as well as readability, which greatly helped to improve the initial draft.

References

[1] J.A. Wheeler and R.P. Feynman, Rev. Mod. Phys. 17 (1945), p. 157.

[2] S. Tanaka, Prog. Theor. Phys. 24 (1960), p. 171. 

[3] Ya.P. Terletsky, Sov. Phys. Dokl. 5 (1961), p. 782.

[4] R. Newton, Science 167 (1970), p. 1569. 

[5] K. Kamoi and S. Kamefuchi, Prog. Theor. Phys. 45 (1971), p. 1646. 

[6] R. Tomaschitz, European Phys. J. B 17 (2000), p. 523. 

[7] R. Tomaschitz, Ann. Phys. 322 (2007), p. 677. 

[8] R. Tomaschitz, European Phys. J. D 32 (2005), p. 241. 

[9] F. Aharonian et al., Astrophys. J. 636 (2006), p. 777. 

[10] J. Albert et al., Astrophys. J. 643 (2006), p. L53.

[11] J. Albert et al., Astrophys. J. 637 (2006), p. L41. 

[12] R. Tomaschitz, Phys. Lett. A 366 (2007), p. 289. 

[13] L.D. Landau and E.M. Lifshitz, Electrodynamics of Continuous Media, Pergamon Press, Oxford (1984).

[14] M. Born and E. Wolf, Principles of Optics, Cambridge University Press, Cambridge (2003).

[15] A. Lagarkov and V. Kisel, Phys. B 394 (2007), p. 163. 

[16] G. Xu et al., Phys. B 403 (2008), p. 3417. 

[17] L.-Y. Wu, L.-W. Chen and R.C.-C. Wang, Phys. B 403 (2008), p. 3599. 

[18] R. Tomaschitz, European Phys. J. C 49 (2007), p. 815.

[19] J.A. Adam, Phys. Rep. 356 (2002), p. 229. 

[20] J.A. Stratton, Electromagnetic Theory, Wiley–IEEE Press, New York (2007).

[21] R. Tomaschitz, Phys. A 387 (2008), p. 3480. 

[22] R. Tomaschitz, Phys. Lett. A 372 (2008), p. 4344. 

[23] R. Tomaschitz, Phys. A 385 (2007), p. 558. 

[24] R. Tomaschitz, EPL 84 (2008), p. 19001. 

[25] J.W. Cronin, Nucl. Phys. B (Proc. Suppl.) 138 (2005), p. 465. 

[26] M. Nagano and A.A. Watson, Rev. Mod. Phys. 72 (2000), p. 689. 

[27] B. Davies et al., Astrophys. J. 676 (2008), p. 1016. 

[28] W.W. Tian et al., Astrophys. J. 657 (2007), p. L25. 

[29] D.J. Helfand et al., Astrophys. J. 665 (2007), p. 1297. 

[30] R. Tomaschitz, EPL 85 (2009), p. 29001. 

[31] R. Tomaschitz, Opt. Commun. (2009), doi:10.1016/j.optcom.2009.01.024.

[32] R. Tomaschitz, Astropart. Phys. 27 (2007), p. 92. 


Corresponding Author Contact InformationTel.: +81 824 247361; fax: +81 824 240717.