Physics Letters A
Volume 366, Issues 4-5, 2 July 2007, Pages 289-297




Tachyonic spectral map of a binary pulsar

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 7 January 2007; 
accepted 29 January 2007. 
Communicated by V.M. Agranovich. 
Available online 21 February 2007.

Abstract

We study tachyonic synchrotron densities of ultra-relativistic electrons in helical motion. There is a longitudinally polarized spectral component due to the negative mass square of the superluminal quanta. The helical pitch-angle scaling of the transversal and longitudinal spectral densities is investigated, in particular the transition from circular to rectilinear motion. The magnetic field induces oscillations along the power-law slopes of the superluminal radiation densities, whose amplitude depends on the pitch angle. At moderate field strength and ultra-relativistic orbital speed, the modulations are tiny, but they become quite pronounced in the surface fields of γ-ray pulsars, resulting in cockscomb distributions. A tachyonic spectral fit to the γ-ray flux of the binary pulsar PSR B1259–63 is performed. The γ-ray wideband of this millisecond pulsar is reproduced by a tachyonic cascade spectrum, capable of generating the spectral curvature in double-logarithmic plots as well as the extended spectral plateau defined by ISGRI, OSSE, COMPTEL, and EGRET flux points.

Keywords: Tachyonic synchrotron radiation; Superluminal curvature radiation; Longitudinal polarization; Cascade spectra; Pulsar magnetospheres

PACS classification codes: 41.60.Ap; 03.50.Kk; 95.30.Gv; 97.60.Gb

Article Outline

1. Introduction
2. Superluminal radiation densities
2.1. High-frequency regime
2.2. Analytic continuation into the low-frequency band
3. Pitch-angle averaging
4. Tachyonic spectral fit to the γ-ray broadband of the binary pulsar PSR B1259–63
5. Conclusion
Acknowledgements
References

1. Introduction

We investigate tachyonic synchrotron emission from electrons in pulsar magnetospheres. The goal is to provide quantitative evidence for this radiation by working out a specific example, a tachyonic spectral fit to the γ-ray spectrum of the binary pulsar PSR B1259–63 orbiting around the massive Be star SS 2883. When tachyonic spectral densities are averaged with thermal or power-law electron densities, this leads to cascade spectra [1] and [2], extended spectral plateaus followed by power-law decay terminating in exponential decay. A glance at the spectral map of PSR B1259–63 suggests that it can readily be fitted by a cascade spectrum.

In Section 2, the tachyonic spectral densities of electrons in helical orbits and their pitch-angle scaling are derived. In Section 3, we discuss the pitch-angle averaging and its effect on modulations along the spectral slopes induced by the negative mass-square of the radiation field. In particular, we illustrate the helical cross-over from circular orbits to rectilinear motion in the surface magnetic field of a γ-ray pulsar. In Section 4, we average the tachyonic spectral densities with electronic source populations and perform a wideband fit to the PSR B1259–63/SS 2883 binary system, ranging from high-energy X-rays to TeV γ-rays. In Section 5, we present our conclusions.

We consider electrons in a constant magnetic field B=(0,0,B), B>0. The equations of motion read d(γmv)/dt=ev×B. The helical orbits can be parametrized as

(1)
View the MathML source
where α is the pitch angle between B and v, 0less-than-or-equals, slantαless-than-or-equals, slantπ. The speed υ along the orbit is constant. A pitch angle of π/2 gives circular orbits in the (x,y)-plane, and α=0 or π means straight uniform motion. We put planck constant over two pi=c=1, and use the Heaviside–Lorentz system. The electric charge e is defined negative for electrons, γ is the Lorentz factor, and m the electron mass. We also admit sign(e)=0 in (1), that is, planar transversal oscillations, realizable by undulators in storage rings [3], [4], [5], [6], [7] and [8].

2. Superluminal radiation densities

2.1. High-frequency regime

The tachyonic radiation theory pertinent to a charge in arbitrary motion has been given in Ref. [1]. The multipole expansion of the superluminal Poynting vectors is summarized in Eqs. (2), (3) and (4). We introduce polar coordinates with B as polar axis and polar angle θ, and consider the tachyonic wave vector in the (y,z)-plane, so that k:=k(ω)n, n=(0,sinθ,cosθ). Wave number and frequency are related by the superluminal dispersion relation, View the MathML source, where mt is the mass of the tachyonic quanta. The sign convention is View the MathML source, so that the plus-sign in the dispersion relation implies a negative mass square. We define two transversal polarization vectors, εshort parallel:=(0,−cosθ,sinθ), and εperpendicular:=εshort parallel×n, so that n, εshort parallel and εperpendicular constitute an orthonormal triad, εshort parallel=−εperpendicular×n. We study the ultra-relativistic limit, γmuch greater-than1.

The angular-integrated power radiated in the nth harmonic can readily be assembled from Eqs. (2.27), (2.28) and (2.31) of Ref. [1],

(2)


View the MathML source


View the MathML source
Here, αq=q2/(4πplanck constant over two pic)≈1.0×10−13 is the tachyonic fine structure constant. The superluminal radiation field, a Proca field with negative mass square, is coupled by minimal substitution to the subluminal current, and q is the tachyonic charge of the source particle. The tachyon mass is mt≈2.15 keV/c2. These estimates are obtained from hydrogenic Lamb shifts [9]. The tachyonic frequencies and wave numbers in (2) read

(3)


View the MathML source


View the MathML source

The squared multipole coefficients can be calculated in closed form [1],

(4)


View the MathML source


View the MathML source
We refrain from writing sign2(e)=1, since in this way the contributions of the two linear transversal polarizations (defined by εshort parallel,perpendicular) to the radiated power can easily be distinguished; the term stemming from the εperpendicular-polarized component of the radiation field has the sign2(e)-factor attached. Moreover, by putting sign2(e)=0 in (4), we find the transversal power radiated by oscillating charges in undulator fields. The electromagnetic counterpart of (2), (3) and (4) is recovered in the limit of vanishing tachyon mass [10].

The integrals View the MathML source in (2) give the power transversally or longitudinally radiated in the harmonic ωn. We are interested in the large-n asymptotics,

(5)
View the MathML source
the expansion parameter is View the MathML source. The 1/n2-expansion of the wave numbers reads

(6)
View the MathML source
and the argument zn of the Bessel functions in (4) is expanded as

(7)
View the MathML source

In (2), we expand the integrands in θθc, at the stationary phase, dzn(θc)/dθ=0, which reads

(8)
View the MathML source
Hence,

(9)
View the MathML source
We introduce a new integration variable ψ in (2),

(10)
View the MathML source
where γ=(1−υ2)−1/2much greater-than1, and approximate

(11)
View the MathML source
In the vicinity of θc, the View the MathML source-factors in (4) can be replaced by

(12)


View the MathML source


View the MathML source
and the Bessel functions in (12) by their Nicholson approximation. The latter applies for large positive n and z, so that z/n≈1 and z<n,

(13)


View the MathML source


View the MathML source
In the preceding notation, this means

(14)


View the MathML source


View the MathML source

The integration boundaries in (2) can be extended to infinity, given the decay of the integrands. Assembling the enumerated approximations, we arrive at the transversal power coefficients,

(15)
View the MathML source
and the longitudinal coefficients read

(16)
View the MathML source
Here, we defined

(17)
View the MathML source
with ε in (10). The transversal power radiated in εperpendicular-polarization is given by the sign2(e) term, and the εshort parallel-polarized transversal radiation by the View the MathML source-integration in (15). The integrals in (15) and (16) can be simplified by the identities [10] and [11]

(18)


View the MathML source


View the MathML source


View the MathML source
where the spectral functions on the right-hand side read

The easiest way to derive these identities is to write K1/3 and K2/3 in terms of Ai(x) and its derivative, and then to use the standard integral representation for Ai(x), cf. Section 2.2.

The total power transversally and longitudinally radiated is obtained by summing over the individual modes, View the MathML source. We pass to continuous frequencies by identifying ω=ωnωBn/sin2α, so that the continuous spectral densities read View the MathML source, with n=(ω/ωB)sin2α. We find ξ in (17) as

(20)
View the MathML source
This is ξ(ω) as in Ref. [1], rescaled by1/sinα.

The restriction z<n on the Nicholson asymptotics (13) translates into ω>ωb. In this frequency band, we find the transversal density, cf. (15) and (18),

(21)
View the MathML source


(22)
View the MathML source
where ξ and ωb are defined in (20). The lower plus-sign in (22) refers to the perpendicular-polarization. pT(ω) in (21) stands for the total transversal radiation; we have sign2(e)=1. The longitudinal density is assembled with (16) and (18),

(23)
View the MathML source

The spectral functions F(ξ), G(ξ), and L(ξ) are defined in (19). The pitch angle enters in pT,L(ω) via the rescaled gyrofrequency, ωBωBsinα, cf. (20).

2.2. Analytic continuation into the low-frequency band

Densities (21), (22) and (23) can be continued into the low-frequency regime, ωless-than-or-equals, slantωb, by means of Airy functions. The Airy representation of the spectral functions (19) reads

(24)
View the MathML source
where z=(3ξ/2)2/3. We define the shortcut

(25)
View the MathML source
so that ξ=(2/3)|η|3/2, cf. (20), and find

(26)


View the MathML source


View the MathML source



(27)
View the MathML source
These densities coincide with (22) and (23) for ωgreater-or-equal, slantedωb, and they also constitute the analytic continuation into the lower frequency band ωless-than-or-equals, slantωb, where η(ω,α) is negative. The lower plus-sign in (26) refers to the perpendicular-polarization. The pitch angle enters only via η(ω,α), cf. (25).

3. Pitch-angle averaging

We perform a pitch-angle average with density [11]

(28)
View the MathML source
where λ>−1. This density is normalized to one over the interval [0,π/2]. The averaging is done by replacing View the MathML source in (26) and (27) with

(29)
View the MathML source
and Ai(η)/η in (26) is replaced by

(30)
View the MathML source
Here, η0(ω):=η(ω,α=π/2), cf. (25), so that η=η0/sin2/3α. The averages (29) and (30) remain unaltered, if we replace the integration over View the MathML source by View the MathML source. This also holds for the normalization of density (28), View the MathML source. More generally, we may average over the full pitch-angle range [0,π] with density

(31)
View the MathML source
where λ1,2>−1 and 0less-than-or-equals, slantAless-than-or-equals, slant1. dρλ is defined in (28). The distribution (31) is normalized to one, View the MathML source. Thus the averaging of the spectral densities (26) and (27) is effected by the substitutions

(32)
View the MathML source


(33)
View the MathML source

We denote the averaged densities by left angle bracketpT,L(ω)right-pointing angle bracketσ.

The averages left angle bracket∫Airight-pointing angle bracketλ and left angle bracketAiright-pointing angle bracketλ in (29) and (30) depend on ω only via η0(ω), cf. after (30). We briefly discuss their |η0|much greater-than1 asymptotics, starting with the low-frequency band, ωless-than-or-equals, slantωb, cf. after (27), where η0 is negative. In leading order,

(34)
View the MathML source


(35)
View the MathML source
where βgreater-or-equal, slanted0. The systematic asymptotic expansion of the integral (35) can be obtained by substituting the asymptotic series of the Airy function and applying term-by-term integration, which results in a series of incomplete Γ-functions. The latter are then replaced by their asymptotic series. The leading order stated in (35) is independent of the exponent β, which enters linearly in the next-to-leading order; if β is large, we have to require β/|η0|3/2much less-than1 for the O-term to be small. In (29), we also need

(36)
View the MathML source
Substituting the leading-order asymptotics (34), (35) and (36) into the averages (29) and (30), we find

(37)
View the MathML source
Thus the averaged spectral densities (as defined by (26) and (27) with substitutions (32) and (33)) converge for large |η0| to the densities generated by a charge in uniform motion, cf. Section 4, irrespectively of the choice of parameters in the pitch-angle distribution (31).

We turn to the upper frequency range, ωgreater-or-equal, slantedωb, where η0(ω) is positive. The asymptotics of left angle bracket∫Airight-pointing angle bracketλ and left angle bracketAiright-pointing angle bracketλ is obtained from

(38)


View the MathML source


View the MathML source


View the MathML source
The next-to-leading orders are by a factor of View the MathML source smaller, like in (34), (35) and (36). We thus find, above the break frequency,

(39)


View the MathML source


View the MathML source
The averaged spectral densities vanish for large (and even moderate) positive η0(ω), reflecting the fact that the spectral densities of a charge in uniform motion vanish identically for ωgreater-or-equal, slantedωb, cf. Section 4.

The pitch-angle distribution dσA,λ1,λ2(α) is peaked in the interval [0,π/2], at View the MathML source, provided that λ1>0. If −1<λ1less-than-or-equals, slant0, it decreases monotonically to zero. The peak in the interval [π/2,π] occurs at View the MathML source. If −1<λ2<0, the density increases in this interval from zero to infinity. For instance, λ1,2=1 gives αp1=π/4; λ1,2=1/3 gives αp1=π/6, and λ1,2=3 amounts to a peak at αp1=π/3. The respective second peak in [π/2,π] is obtained as αp2=παp1. If A=1/2 and λ1=λ2, the density is symmetric around α=π/2. If λ1,2>0, it vanishes at the interval boundaries and π/2. These three pitch angles correspond to the limit cases of rectilinear motion parallel and antiparallel to the magnetic field, and to circular orbits orthogonal to B. If λ1,2 is negative or close to zero, the average is dominated by pitch angles close to zero and π, so that the radiation density generated in uniform motion is prevalent. The latter is obtained by performing the zero-pitch-angle limit sinα→0 in (26) and (27), cf. Ref. [12] and Fig. 1 and Fig. 2. If λ1,2much greater-than1, the distribution (31) has two peaks close to π/2, even though it vanishes at π/2, so that the synchrotron densities (26) and (27) at sinα≈1 dominate the average.



Full-size image

Fig. 1. Transversal tachyonic spectral density pT(ω), cf. (26), in the polar surface field, B≈3.1×1013 G, of the γ-ray pulsar PSR B1509–58, cf. Refs. [13], [14] and [15]. The electronic orbital energy is 10 GeV. The density is shown for three pitch angles, α=π/2 (circular orbit), α=0.1, and α=0. The latter corresponds to a charge in straight uniform motion, cf. (40). The radiation densities are rescaled with the tachyonic fine structure constant αq, cf. after (2). The gyrofrequency is ωB≈18 eV, the helical curvature radius R≈1.1×10−6 cm. The break frequency is ωb≈42 MeV. The spectral peak is located at about the tachyon mass, 2 keV, followed by an extended X-ray tail with power-law decay, in contrast to the exponential cutoff in the upper frequency band, cf. (19).


Full-size image

Fig. 2. Longitudinal spectral density pL(ω), cf. (27). Magnetic field, electron energy and pitch angles as in Fig. 1. The oscillations are superimposed by the magnetic field and vanish at zero pitch angle, for rectilinear motion along the magnetic field lines. The transversal and longitudinal spectral densities of a uniformly moving ultra-relativistic electron (dotted curve in Fig. 1 and Fig. 2) coalesce in the depicted frequency range, ωmuch less-thanωb.


The units planck constant over two pi=c=1 can easily be restored. Gyroradius (magnetic bending radius, helical curvature radius) and gyrofrequency are connected by R=c/ωB in the ultra-relativistic regime. The length of one helical turn is 2πR, and Rsinα is the radius of the helical base circle. Electronic energy and Lorentz factor are related as γ≈1957E [GeV], so that View the MathML source and View the MathML source, and the tachyonic break frequency scales as View the MathML source. Fig. 1, Fig. 2 and Fig. 3 illustrate the pitch-angle scaling and averaging of the superluminal radiation densities of ultra-relativistic electrons orbiting in the polar field of a γ-ray pulsar.



Full-size image

Fig. 3. Transversal and longitudinal spectral densities left angle bracketpT,L(ω)right-pointing angle bracketσ, averaged with the symmetric pitch-angle distribution dσ(A=1/2,λ1=λ2=1), cf. (31). This distribution is peaked at pitch angles of 45° and 135°, in the cross-over region between the synchrotron regime and rectilinear motion. Magnetic field and orbital energy as in Fig. 1 and Fig. 2. For comparison, we also indicate the unaveraged transversal distribution pT(ω) at these pitch angles, View the MathML source; the modulations are wiped out by the averaging.


4. Tachyonic spectral fit to the γ-ray broadband of the binary pulsar PSR B1259–63

To compare to observational spectral maps, we have to perform a further average over the electron density [2]. We demonstrated that a pitch-angle average attenuates the modulations in the spectral slope, cf. Fig. 3, so that the synchrotron densities approach the spectral densities of a uniformly moving charge,

(40)
View the MathML source
where ΔT=1 and ΔL=0 for the transversal and longitudinal radiation component, respectively. The spectral cutoff occurs at View the MathML source, to be identified with the break frequency ωb in the ultra-relativistic limit, cf. (20). Only frequencies in the range 0<ω<ωmax(γ) can be radiated by a uniformly moving charge [16].

We average with electronic power laws exponentially cut with a Boltzmann density, View the MathML source. Parametrized with the Lorentz factor, this reads

(41)
View the MathML source
The normalization Aα,β is determined by View the MathML source, where n1 is the electron count and γ1 the smallest Lorentz factor of the source population. β=mc2/(kT), and α is the electronic power-law index (not to be confused with the pitch angle). Here, we consider thermal distributions, defined by α=−2 and γ1=1. The average is carried out as [2]

(42)
View the MathML source
The spectral fit is based on the E2-rescaled flux density dNT,L/dE, which is related to the averaged energy density by

(43)
View the MathML source
where d is the distance to the pulsar. In Fig. 4, Fig. 5 and Fig. 6, we plot the flux density of two thermal electron populations ρ1,2 specified in Table 1. Each electron density generates a cascade ρi, the wideband spectral map is obtained by adding the cascade spectra.



Full-size image

Fig. 4. γ-ray wideband of the binary pulsar PSR B1259–63. RXTE and ISGRI data points from [18], OSSE points from [19], COMPTEL and EGRET upper limits from [20], HESS points from [21]. The solid line T+L is the unpolarized differential tachyon flux dNT+L/dE, obtained by adding the flux densities ρ1,2 of two thermal electron populations. The χ2-fit is based on the first three ISGRI points, the first four OSSE points, and the HESS points. The fourth ISGRI point, the remaining five OSSE points, as well as the COMPTEL and EGRET points are upper limits only. The four depicted RXTE points are at the upper edge of the hard X-ray spectrum detectable with the RXTE proportional counter [18], they suggest a seamless transition into the X-ray band, but have not been included in the least-squares fit.


Full-size image

Fig. 5. EGRET plateau and HESS slope of PSR B1259–63. The solid line T+L is the unpolarized tachyon flux dNT+L/dE rescaled with E2 to render the spectral curvature visible. The latter is caused by the Boltzmann factor in the averaged spectral densities (42); the exponential decay of the cascades ρ1,2 sets in at about Ecut≈(mt/m)kT, implying cutoffs at 0.8 TeV for the ρ1-cascade and 80 GeV for ρ2. The unpolarized flux T+L can be split into transversal (dot-dashed) and longitudinal (dot-dot-dashed) components dNT,L/dE, cf. (43). The spectral plateau extends from the ISGRI and OSSE points in the keV range to the first spectral cut at 80 GeV, cf. Fig. 4.


Full-size image

Fig. 6. Close-up of the HESS spectrum of PSR B1259–63 in Fig. 5. The tachyon flux T+L is obtained by adding the cascades ρ1,2. The TeV spectral map coincides with the ρ1-cascade, since the ρ2-flux is exponentially cut at 80 GeV, cf. Fig. 5. Temperature and source number of the electron populations generating the cascades are listed in Table 1.


Table 1.

Thermal electronic source densities ρ1,2 generating the γ-ray broadband of the binary pulsar PSR B1259–63

View the MathML source β ne kT (TeV)
ρ1 1.9×10−4 2.69×10−9 2.4×1046 190
ρ2 2.3×10−3 2.69×10−8 2.95×1047 19

The cascades ρ1,2 in Fig. 4, Fig. 5 and Fig. 6 are obtained by averaging the superluminal radiation densities with electron populations, cf. (42), which are in turn inferred from the spectral fit. Each distribution is defined by electron temperature kT and source count ne, the latter based on a distance estimate of 1.5 kpc. β is the cutoff parameter in the electronic Boltzmann factor. View the MathML source is the normalized amplitude of the tachyonic flux density defined in (44). The parameters View the MathML source and β are extracted from the χ2-fit.


As for the electron count n1, it is convenient to use a rescaled parameter View the MathML source for the fit,

(44)
View the MathML source
which is independent of the distance estimate. Here, planck constant over two pi [keV s] implies the tachyon mass in keV units, that is, we put mt≈2.15 in the spectral density (40). At γ-ray energies, only a tiny αq/αe-fraction (the ratio of tachyonic and electric fine structure constants) of the tachyon flux is actually absorbed by the detector [17], which requires a rescaling of the electron count n1, so that the actual number of radiating electrons is View the MathML source. We thus find the electron count as View the MathML source [kpc], where View the MathML source is determined by the tachyonic flux amplitude obtained from the spectral fit. Finally, electron temperature and cutoff parameter in the electronic Boltzmann factor are related as View the MathML source.

Fig. 4, Fig. 5 and Fig. 6 show the least-squares fit to the γ-ray wideband of the binary system PSR B1259–63/SS 2883, cf. Refs. [18], [19], [20], [21], [22], [23] and [24]. The extended spectral plateau as well as the spectral curvature over the entire frequency range can readily be reproduced by a tachyonic cascade spectrum. (As for the high-magnetic-field pulsar PSR B1509–58, cf. Fig. 1, Fig. 2 and Fig. 3, there are currently not enough data points available for a spectral fit [14]. A tachyonic spectral fit to the pulsed Crab flux is given in Ref. [25], where the spectral asymptotics of the averaged densities and the technical details of the nonlinear χ2-fit are discussed as well.) We assume d≈1.5 kpc as distance to the pulsar, based on optical and scintillation observations [22], although it could be as high as 4.5 kpc if inferred from the dispersion measure [23]. In the latter case, the electronic source count has to be rescaled accordingly, View the MathML source; the spectral maps are not affected by the distance estimate. The total flux above 380 GeV is 5% of the unpulsed Crab flux [21]. The electronic source count for the Crab pulsar at 2 kpc is 2.6×1049, cf. Ref. [25], as compared to 3.2×1047 for PSR B1259–63, cf. Table 1.

Several electromagnetic radiation mechanisms have been invoked to model the spectra of γ-ray pulsars, most notably curvature radiation due to electric fields followed by pair creation, accompanied by synchrotron radiation and inverse Compton scattering [26]. An electromagnetic spectral fit to the γ-ray wideband of the PSR B1259–63/SS 2883 binary was performed in Ref. [20], in the context of a synchro–inverse-Compton model; the tachyonic wideband fit in Fig. 4 is directly comparable to Fig. 16 of this reference. The tachyonic TeV fit in Fig. 6 is directly comparable to the inverse-Compton fit in Fig. 15 of Ref. [24].

5. Conclusion

We have studied superluminal radiation from ultrarelativistic electrons helically orbiting in strong magnetic fields and derived the pitch-angle scaling of the radiation densities. In the zero-magnetic-field limit, the tachyonic synchrotron densities converge to the spectral densities (40) for uniform motion. Orbital curvature induces modulations along the spectral slopes. If integrated over the pitch angles, these oscillations are averaged out, cf. Fig. 1, Fig. 2 and Fig. 3.

We demonstrated that the γ-broadband of the binary pulsar PSR B1259–63 can be reproduced by a tachyonic cascade spectrum, cf. Fig. 4, Fig. 5 and Fig. 6. The cascades are obtained by averaging the radiation densities with thermal electron distributions. At γ-ray energies, the speed of tachyons is close to the speed of light; the basic difference to electromagnetic radiation is the longitudinally polarized flux component. The polarization of tachyons can be determined from transversal and longitudinal ionization cross-sections of Rydberg atoms, which peak at different scattering angles [27].

We have focused on the γ-ray spectrum, including the high end of the X-ray tail. At X-ray energies, if a grating spectrometer is used [28], the interference peaks are determined by wavelength. In this case, the data points have to be reparametrized by wavelength, and then the tachyonic dispersion relation is substituted (which substantially differs from the photonic counterpart due to the 2 keV tachyon mass located in the X-ray regime) to obtain the tachyonic energy-flux relation. Tachyonic spectral fits in the X-ray band will be discussed elsewhere.

Acknowledgements

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.

References

[1] R. Tomaschitz, Physica A 335 (2004), p. 577.

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

[3] K. Nakagawa et al., J. Electron Spectrosc. Relat. Phenomena 144 (2005), p. 271.

[4] K. Yagi-Watanabe et al., Nucl. Instrum. Methods Phys. Res. A 553 (2005), p. 620. 

[5] I.H. Suzuki et al., Radiat. Phys. Chem. 75 (2006), p. 1778. 

[6] Y. Morishita et al., J. Phys. B 39 (2006), p. 1323. 

[7] Y. Zou et al., J. Phys. B 39 (2006), p. 4775. 

[8] M. Oura et al., Radiat. Phys. Chem. 76 (2007), p. 469. 

[9] R. Tomaschitz, Eur. Phys. J. D 32 (2005), p. 241.
R. Tomaschitz, Eur. Phys. J. B 17 (2000), p. 523. 

[10] A.A. Sokolov and I.M. Ternov, Radiation from Relativistic Electrons, Hilger, Bristol (1986).

[11] V.L. Ginzburg and S.I. Syrovatskii, Annu. Rev. Astron. Astrophys. 7 (1969), p. 375.
K.C. Westfold, Astrophys. J. 130 (1959), p. 241. 

[12] R. Tomaschitz, Eur. Phys. J. C 45 (2006), p. 493. 
R. Tomaschitz, Astropart. Phys. 23 (2005), p. 117. 

[13] S. Gunji et al., Astrophys. J. 428 (1994), p. 284. 

[14] L. Kuiper et al., Astron. Astrophys. 351 (1999), p. 119. 

[15] T. Sako et al., Astrophys. J. 537 (2000), p. 422. 

[16] R. Tomaschitz, Physica A 320 (2003), p. 329. 

[17] R. Tomaschitz, Eur. Phys. J. C 49 (2007), p. 815. 

[18] S.E. Shaw et al., Astron. Astrophys. 426 (2004), p. L33. 

[19] J.E. Grove et al., Astrophys. J. 447 (1995), p. L113. 

[20] M. Tavani and J. Arons, Astrophys. J. 477 (1997), p. 439. 

[21] F. Aharonian et al., Astron. Astrophys. 442 (2005), p. 1. 

[22] N.M. McClure-Griffiths et al., Astrophys. J. 492 (1998), p. L49. 

[23] M. Chernyakova et al., Mon. Not. R. Astron. Soc. 367 (2006), p. 1201. 
T.W. Connors et al., Mon. Not. R. Astron. Soc. 336 (2002), p. 1201. 

[24] D. Khangulyan et al. astro-ph/0605663.

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

[26] K. Hirotani, Adv. Space Res. 35 (2005), p. 1085. 

[27] R. Tomaschitz, J. Phys. A 38 (2005), p. 2201. 

[28] C.R. Canizares et al., Publ. Astron. Soc. Pac. 117 (2005), p. 1144.


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