Physica A: Statistical Mechanics and its Applications
Volume 387, Issue 14, 1 June 2008, Pages 3480-3494




Tachyonic γ -ray wideband of an ultra-relativistic electron plasma: Spectral fitting with Fermi power-law densities

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 28 September 2007; 
revised 14 January 2008. 
Available online 15 February 2008.

Abstract

Fermionic power-law distributions are derived by the second quantization of classical power-law ensembles, and applied to ultra-relativistic electron populations in the Galactic center. The γ-ray flux from the direction of the compact central source Sagittarius A* is fitted with a superluminal cascade spectrum. In this way, estimates of the radiating electron plasma in the Galactic center region are obtained, such as the power-law index, temperature, particle number, and internal energy. The spectral averaging of the tachyonic radiation densities with Fermi power-laws is explained. Fugacity expansions of the thermodynamic variables (thermal equation of state, entropy, isochoric heat capacity, and isothermal compressibility) are obtained in the quasiclassical high-temperature/low-density regime, where the spectral fit is carried out. The leading quantum correction to these variables is calculated, and its dependence on the electronic power-law index and the thermal wavelength is discussed. Excess counts of cosmic rays from the Galactic center region are related to the plasma temperature inferred from the cascade fit.

Keywords: Superluminal radiation; Tachyonic cascade spectra; Fermi power-law ensembles; Ultra-relativistic electron plasma; Quasiclassical fugacity expansion; Spectral averaging

PACS classification codes: 05.30.Fk; 05.70.Ce; 52.25.Kn; 95.30.Tg

Article Outline

1. Introduction
2. Thermodynamic variables of Fermi power-law ensembles
3. Equations of state, entropy, and heat capacity of fermionic power-law densities in the quasiclassical regime
3.1. Low-temperature asymptotics
3.2. Fugacity expansion and thermal wavelength in the high-temperature limit
4. Superluminal γ-rays from the Galactic center: Tachyonic spectral maps and electronic source distributions
5. Conclusion
Acknowledgements
References

1. Introduction

Electronic power-law distributions are commonly used in electromagnetic spectral averages to model the synchrotron emission of astrophysical sources, such as the X-ray spectra of supernova remnants [1]. In this article, we quantize Boltzmann power-law densities, exponentially cut power-law distributions View the MathML source, where H=mc2γ is the free electronic Hamiltonian. In Fermi–Dirac statistics, we arrive at densities

(1.1)
View the MathML source
where the momentum integration has been parametrized with the electronic Lorentz factor γ. δ is the power-law index, and we use the shortcuts View the MathML source and α=−βμ/(mc2), where μ is the chemical potential. Electronic and protonic power-law densities have been invoked to perform electromagnetic and hadronic spectral fits to various TeV γ-ray sources recently discovered with imaging air Cherenkov detectors [2]. Tachyonic spectral fits to the cascade spectra of γ-ray pulsars and microquasars are based on electronic power-law averages as well [3] and [4]. Power-law indices inferred from these spectral fits or, for that matter, from the magnetospheric radio emission of planets [5] and [6], usually range in the interval 0less-than-or-equals, slantδless-than-or-equals, slant4.

Here, we give evidence for superluminal radiation from the Galactic center by fitting recently obtained spectral maps [7], [8] and [9] with tachyonic cascade spectra. The tachyonic γ-ray wideband consists of two cascades generated by ultra-relativistic electron populations, and the spectral fit allows us to infer the thermodynamic parameters. The observed spectra are clearly distinguishable from electromagnetic synchrotron and inverse-Compton fits, due to the emergence of extended spectral plateaus. We show that the TeV spectral map of the Galactic center admits a tachyonic extension into the GeV region, providing an excellent fit to the spectrum of the unidentified γ-ray source 3EG J1746 − 2851 [10] associated with the Galactic central source Sagittarius A*. The superluminal cascades are generated by an ultra-relativistic electron gas at high temperature and low density, so that we can use distribution (1.1) in the quasiclassical regime to calculate the thermodynamic variables and the spectral averages.

In Section 2, we set up the thermodynamic formalism of fermionic power-law densities (1.1). Starting with the grand partition function, we derive the quasiclassical fugacity expansion of the thermodynamic variables, such as the caloric and thermal equations of state, entropy, specific heat, and compressibility. In Section 3, we discuss the leading quantum correction to the classical thermodynamic functions. In the high-temperature regime, the quantum corrections become more pronounced with increasing electronic power-law index, exhibiting the same temperature scaling as the classical limit usually dominant in the high-temperature/low-density regime. This is illustrated by calculating the mentioned variables for an increasing sequence of electron indices. We study the thermal wavelength of fermionic power-law ensembles in the low- and high-temperature regime, and discuss the range of applicability of the quasiclassical fugacity expansion.

In Section 4, we average the quantized superluminal radiation densities with the power-law distributions (1.1), and derive the fugacity expansion of the spectral averages. We perform a tachyonic cascade fit to the γ-ray spectrum of the Galactic center, and obtain estimates of the thermodynamic parameters of the electron plasma generating the superluminal radiation. The cutoff energy of the high-energy cascade fitting the TeV spectrum can be related to anisotropies in the cosmic ray spectrum detected by the AGASA and SUGAR air shower arrays [11] and [12]. If the TeV cascade is generated by a protonic source population, this requires a thermal proton density at View the MathML source. This cutoff temperature very closely matches the upper energy edge of both the AGASA and SUGAR excess counts from the Galactic center region, suggesting that the compact radio source Sagittarius A* is capable of accelerating protons into the 1018 eV region. In Section 5, we present our conclusions.

2. Thermodynamic variables of Fermi power-law ensembles

We start with the fermionic partition function,

(2.1)
View the MathML source
where ρ(H) stands for density

(2.2)
View the MathML source
and H=mγ is the free Hamiltonian. The electronic Lorentz factors γ=(1−υ2)−1/2 range in an interval γ1less-than-or-equals, slantγ<. H1=mγ1 is the lower threshold energy, and γ1≥1 the lower edge of Lorentz factors of the electron distribution. The momentum parametrization is View the MathML source, so that we can substitute View the MathML source, where the angular integration has already been performed, as there is no angular dependence in the integrand. The Heaviside step function θ restricts the View the MathML source integration in (2.1) to γγ1. The exponent α defines the fugacity View the MathML source, and is related to the chemical potential by μ=−mα/β, cf. after (2.26). δ is the electronic power-law exponent, and View the MathML source the cutoff parameter in the Boltzmann factor, so that the relativistic Fermi–Dirac distribution is recovered with δ=0 and γ1=1. Here, we study power-law ensembles of arbitrary real power-law index δ. We use c=ħ=1 for most of this article; the units can easily be restored, e.g., View the MathML source, where the ratio of electron rest energy and Boltzmann constant is View the MathML source.

The grand partition function (2.1) is obtained via a standard trace calculation in fermionic occupation number representation,

(2.3)
View the MathML source
where View the MathML source refers to the trace over multi-particle states with occupation energies ω exceeding the lower threshold ω1colon, equalsH1. We briefly sketch the derivation of partition function (2.1) from the trace (2.3). A basis |nright-pointing angle bracket for the occupation number representation of the fermionic creation/annihilation operators View the MathML source (labeled by the wave vector View the MathML source) is explicitly defined in Eq. (4.11) in Ref. [13]. In (2.3), we substitute the particle number operator View the MathML source, where View the MathML source, as well as the energy operator View the MathML source, where View the MathML source. In occupation number representation, the Hermitian number operators are diagonal, View the MathML source, and the fermionic occupation numbers View the MathML source attached to a wave vector View the MathML source can only take the values zero and one. We also note the dispersion relation View the MathML source, relating electron energy and wave number View the MathML source. Finally, View the MathML source is the wave number corresponding to the lower threshold energy ω1=mγ1.

We employ box quantization, discretizing the wave vector as View the MathML source. The following summations are taken over integer lattice points View the MathML source in R3, corresponding to periodic boundary conditions on a box of size L3. By making use of View the MathML source, we may write trace (2.3) over the basis states |nright-pointing angle bracket as

(2.4)
View the MathML source
We thus find

(2.5)
View the MathML source
The continuum limit L amounts to replacing the summation over the lattice wave vectors by the integration View the MathML source [14]; the factor of two accounts for the spin degeneracy. As we have put ħ=c=1, we can identify View the MathML source, so that View the MathML source. Performing these substitutions in (2.5), we arrive at integral representation (2.1) of the partition function, since the Heaviside function in the integrand (2.2) is equivalent to the restriction pk1 of the momentum integration (2.1).

The internal energy,

(2.6)
View the MathML source
and the particle number,

(2.7)
View the MathML source
are obtained from the integral representation (2.1),

(2.8)
View the MathML source
Here we have reparametrized the momentum integration View the MathML source in (2.1) with the Lorentz factor as explained after (2.2). The integral representations (2.6), (2.7) and (2.8) are the key ingredients of the Legendre formalism employed below; in particular, logZ(δ,β,α,V) in (2.8) is the starting point for the quasiclassical fugacity expansion (2.10) (at high temperature and low density) of the thermodynamic functions. The opposite asymptotic limit, the nearly degenerate quantum regime (low temperature, high density), is based on the representation

(2.9)
View the MathML source
where the logarithm in the integrand of (2.8) has been removed by a partial integration. That is, we write the integrand in (2.8) as f(γ)g(γ)/3, with View the MathML source and g=(γ2−1)3/2, and integrate by parts to arrive at (2.9).

For the remainder of this section, we derive the quasiclassical fugacity expansions of the thermodynamic variables. In Section 3, we will discuss the low- and high-temperature limits of these expansions and conditions for their applicability. We start by expanding partition function (2.8) in ascending powers of View the MathML source,



(2.10)
View the MathML source


(2.11)
View the MathML source
and introduce the shortcuts

(2.12)
View the MathML source
so that K1=K and Kn=Kn,0. Differentiation with respect to β is denoted by a prime, View the MathML source. The fugacity expansions of the partition function, internal energy, and particle number thus read



(2.13)
View the MathML source


(2.14)
View the MathML source


(2.15)
View the MathML source
We eliminate the fugacity in the partition function and the internal energy by inverting View the MathML source in (2.15):

(2.16)
View the MathML source
with expansion parameter

(2.17)
View the MathML source
The following expansions are ascending series in View the MathML source. On substituting (2.16) into (2.13) and (2.14), we find the fugacity expansion of the partition function in Helmholtz parametrization

(2.18)
View the MathML source
as well as the internal energy

(2.19)
View the MathML source
Ucl is the classical limit, the internal energy of a Boltzmann power-law density [4], and the first two series coefficients in (2.19) read

(2.20)
View the MathML source
The fugacity expansion of the entropy function

(2.21)
View the MathML source
is obtained by substituting the ascending View the MathML source-series (2.18), (2.19), and, cf. (2.16),

(2.22)
View the MathML source
We find

(2.23)
View the MathML source
where the coefficients u1,2 are defined in (2.20), and

(2.24)
View the MathML source
is the entropy of a classical Boltzmann power-law density, cf. Ref. [4] and (4.13), with a term View the MathML source added owing to the multiplicity factor in (2.1).

The thermal equation of state is derived from the Helmholtz free energy,

(2.25)
View the MathML source
as P=−F/V. By making use of the View the MathML source-series (2.18) and (2.22), we obtain

(2.26)
View the MathML source
The fugacity expansion of the chemical potential μ=F/N=−mα/β is found by substituting the series expansion (2.22) of α.

The isochoric specific heat and the isothermal compressibility are

(2.27)
View the MathML source
where V(δ,β,P,N) is obtained by solving the thermal equation (2.26). As for the heat capacity, we find the fugacity expansion by substitution of (2.23) and (2.24)

(2.28)
View the MathML source
where CV,cl denotes the classical limit

(2.29)
View the MathML source
To obtain the quantum correction of the isothermal compressibility, we have to iteratively solve (2.26) for V (since View the MathML source, cf. (2.17)),

(2.30)
View the MathML source
where κT,cl=1/P is the classical limit. Thermodynamic stability requires CV≥0 and κT≥0. The leading order of the fugacity expansion of CV in (2.28) and κT in (2.30) is in either case positive, as it coincides with the classical limit based on a Boltzmann power-law density [4].

3. Equations of state, entropy, and heat capacity of fermionic power-law densities in the quasiclassical regime

3.1. Low-temperature asymptotics

In the quasiclassical regime covered by the fugacity expansions in Section 2, the low-temperature limit of the thermodynamic variables is determined by the βmuch greater-than1 asymptotics of integral K(δ,β,γ1) in (2.11). At γ1=1, the asymptotic 1/β-series reads

(3.1)
View the MathML source
to be substituted into the fugacity expansions of the respective variables, cf. (2.19), (2.23), (2.26), (2.28) and (2.30). In the following, we put γ1=1 in partition function (2.8). (An ultra-relativistic lower threshold energy can be treated in like manner; the low-temperature expansion of K(δ,β,γ1much greater-than1) is given in Ref. [15].) The series coefficients in the fugacity expansions are composed of certain K-ratios. Using the notation (2.12), we find

(3.2)
View the MathML source
and the same for K2,1/(K1,1K1), with 45/16 replaced by 57/16. As for coefficient u1 in (2.20), we note View the MathML source, up to terms of O(1/β). Finally, K1,1/K1not, vert, similar1 in leading order. Since View the MathML source, cf. (2.17), these ratios already suffice to obtain the temperature and density scaling of the leading quantum correction to the thermodynamic variables. The low-temperature limit of the internal energy (2.19) is found as

(3.3)
View the MathML source
where the ellipsis indicates terms of View the MathML source, that is, higher orders in N/V. The leading factor Ucl is the low-temperature expansion of the internal energy of a classical power-law density [4]. The thermal equation of state (2.26) reads

(3.4)
View the MathML source
where the quantum correction N/V stems from the linear View the MathML source-term in (2.26). We may replace β by the thermal wavelength, View the MathML source, to find condition View the MathML source for the quasiclassical fugacity expansion to apply; at low temperatures, the gas has to be sufficiently dilute. The mean kinetic energy per particle in the low-temperature regime is Eavcolon, equalsU/Nm, so that βm/Eav in leading order. We thus recover the nonrelativistic wave-mechanical scaling View the MathML source.

The leading quantum correction to entropy, heat capacity, and compressibility scales linearly with density N/V. As for the entropy, cf. (2.23) and (2.24), we find



View the MathML source


(3.5)
View the MathML source
and the specific heat at constant volume reads, cf. (2.28) and (2.29),

(3.6)
View the MathML source
The classical isothermal compressibility is κT,cl=1/P, and the quantum correction is found as, cf. (2.30),

(3.7)
View the MathML source
The series expansions (3.3)(3.7) are in ascending powers of N/V or P. We have only stated the leading quantum correction, that is, the term linear in N/V or P. The indicated temperature scaling of this term is meant in leading order as well, the next-to-leading order being smaller by a factor of O(1/β), cf. (3.3).

3.2. Fugacity expansion and thermal wavelength in the high-temperature limit

Like in Section 3.1, the quasiclassical high-temperature asymptotics of the thermodynamic variables are based on the ascending View the MathML source-series View the MathML source derived in Section 2, namely U in (2.19), S in (2.23), the thermal equation of state (2.26), and CV in (2.28). As for the compressibility κT, we will use the ascending P-series in (2.30). The leading quantum correction to these variables is the term linear in View the MathML source or P, which is in all cases composed of the K-ratios, cf. (2.11) and (2.12),

(3.8)
View the MathML source
The high-temperature expansion (βmuch less-than1) of these ratios is calculated from the ascending β-series of integral K(δ,β,1) in (2.11), cf. Ref. [15]. The structure of the high-temperature expansion of K(δ,β,1) depends on the power-law index δ. In the following, we list the thermodynamic functions at integer power-law indices δ=0,1,…,4. These indices exhaust all qualitatively different cases; non-integer power-law indices can be dealt with in the same way, based on the β-expansions of K(δ,β,γ1) in Ref. [15], which cover real δ as well as ultra-relativistic threshold Lorentz factors γ1much greater-than1. In this subsection, we study γ1=1 and integer power-law indices 0less-than-or-equals, slantδless-than-or-equals, slant4.

At δ=0, a thermal Fermi–Dirac distribution is recovered, admitting the quasiclassical high-temperature expansions [14]



View the MathML source


View the MathML source


(3.9)
View the MathML source
We write here κTP for κT/κT,cl, cf. (2.30). The quantum corrections are the terms linear in N/V or P; the logarithms in the expansion of S stem from the classical term Scl in (2.24). At index δ=1, we find



View the MathML source


View the MathML source


(3.10)
View the MathML source
There is no qualitative difference to the Fermi–Dirac variables (3.9) as yet. The entropy has the usual logarithmic divergence, the heat capacity approaches a finite limit, and the internal energy has a 1/β divergence. All quantum corrections vanish for β→0. At δ=2, we obtain



View the MathML source


View the MathML source


(3.11)
View the MathML source
In the thermal equation of state and the compressibility, the quantum correction is augmented by a factor of 1/β as compared to the previous cases δ=0,1, which has implications for the thermal wavelength, cf. after (3.16). The quantum correction to the compressibility is of the same order as in the thermal equation, since Pnot, vert, similarmN/(βV).

For index δ=3, there emerges a logarithmic temperature dependence already in the classical internal energy and heat capacity [4]. Introducing the shortcut

where γE≈0.5772 is Euler’s constant, we find



View the MathML source


View the MathML source


View the MathML source


(3.13)
View the MathML source
The entropy diverges very slowly for β→0, owing to a double-logarithmic divergence. The heat capacity approaches zero logarithmically, but stays positive at finite temperature. The quantum corrections vanish logarithmically, except for κT. At δ=4, the expansions read



View the MathML source


View the MathML source


View the MathML source


(3.14)
View the MathML source
with lE(β) defined in (3.12). In this case, the internal energy diverges logarithmically, and the entropy approaches a finite limit at β=0. (For power-law indices δ>4, the internal energy reaches saturation, attaining a finite limit at β=0 like the entropy, since integral (2.6) stays finite without exponential cutoff.) In (3.14), the quantum correction to the thermal equation is in leading order independent of β, approaching the indicated finite limit linear in N/V. The quantum correction to the specific heat has the same linear β-dependence as the classical term, in leading order that is, and both terms are positive. (At δ=5, the classical term as well as the quantum correction scale β2lE(β) in leading order, with positive proportionality constants.)

In the fugacity expansions (3.9), (3.10), (3.11), (3.12), (3.13) and (3.14), the quantum correction is the term linear in N/V (or P in the case of κTP). Only the leading order in β is indicated, in the classical term as well as the quantum correction, except for the entropy function in (3.14), where the next-to-leading order in β is included in the quantum correction, so that the heat capacity can be recovered by differentiation, cf. (2.27). Otherwise, the omitted β-terms are by at least a factor of O(βlogβ) smaller than the indicated ones. (In the expansion procedure, the logs are treated as constants.) The ellipses in expansions (3.9), (3.10), (3.11), (3.12), (3.13) and (3.14) indicate higher-order quantum corrections in powers of N/V. The units are restored by replacing m by mc2 on the left-hand side of the above equations as well as in β, cf. after (2.2); on the right-hand side, m is replaced by mc.

As mentioned after (3.8), high-temperature expansions at a non-integer power-law exponent δ are calculated analogously, by making use of ratios (3.8) and the expansions of K(δ,β,1) in Ref. [15]. High-temperature expansions of thermodynamic variables are differently structured in different δ-intervals, n−1<δ<n, joining the expansions (3.9), (3.10), (3.11), (3.12), (3.13) and (3.14) at integer δ. Ultra-relativistic high-temperature expansions of K(δ,β,γ1much greater-than1) have been obtained in Ref. [15] as well, to be substituted into the ratios (3.8) in the case of power-law distributions with Lorentz factors exceeding a high-energy threshold, cf. (2.3).

Returning to the thermal equation of state (2.26), we define the thermal wavelength λT by writing the leading quantum correction in (2.26) as

(3.15)
View the MathML source
The numerical factor in λT is chosen in a way to recover the usual definition, View the MathML source, in the low-temperature limit: View the MathML source for βmuch greater-than1. The numerical proportionality factors in View the MathML source and λT are a mere convention; λT provides a second length scale to be compared to (V/N)1/3, which sets the scale for the quantum correction, the quasiclassical regime being defined by View the MathML source, cf. after (3.4). The high-temperature scaling of λT at integer power-law indices δ=0,1,…,4 is



View the MathML source


(3.16)
View the MathML source
where lE(β) denotes the logarithmic temperature dependence (3.12). If δless-than-or-equals, slant3, we invert the caloric equation of state to find in leading order, cf. (3.9), (3.10), (3.11), (3.12) and (3.13),

(3.17)
View the MathML source
where Eav stands for the mean particle energy U/N, and βmuch less-than1 is implied. On combining (3.16) and (3.17), we find the high-temperature dispersion relation λT(Eav). At δ=0,1, we recover the ultra-relativistic wave-mechanical scaling λT∝1/Eav. For δ≥4, the thermal wavelength is proportional to the Compton wavelength, λT∝1/m. For intermediate power-law indices, the thermal wavelength is a hybrid of ultra-relativistic and Compton wavelength, e.g., View the MathML source.

4. Superluminal γ-rays from the Galactic center: Tachyonic spectral maps and electronic source distributions

In this section, we average tachyonic radiation densities [15] with fermionic power-law distributions, and use the spectral averages to perform a cascade fit to the γ-ray broadband of the Galactic central source Sagittarius A*, cf. Fig. 1 and Fig. 2. The thermodynamic parameters of the electron plasma in the Galactic center generating the superluminal cascades are extracted from the spectral fit, cf. Table 1. First we briefly summarize the tachyonic radiation densities, cf. (4.1) and (4.2), then we explain the spectral averaging, in particular the fugacity expansion of the spectral functions and their low- and high-temperature asymptotics, cf. (4.3), (4.4), (4.5), (4.6), (4.7), (4.8), (4.9), (4.10), (4.11), (4.12), (4.13), (4.14), (4.15), (4.16), (4.17), (4.18) and (4.19). The spectral fit in Fig. 1 and Fig. 2 is discussed after (4.20).



Full-size image

Fig. 1. γ-ray wideband of the Galactic center. EGRET data points from Ref. [2], HESS points from Refs. [7] and [8], MAGIC points from Ref. [9]. EGRET points refer to the source 3EG J1746 − 2851, HESS and MAGIC points to HESS J1745 − 290. The solid line T+L depicts the unpolarized differential tachyon flux View the MathML source, obtained by adding the flux densities ρ1,2 of two electron populations and rescaled with E2 for better visibility of the spectral curvature, cf. (4.20). The transversal (T, dot-dashed) and longitudinal (L, double-dot-dashed) flux densities View the MathML source add up to the total flux T+L. The χ2-fit is done with the unpolarized tachyon flux T+L, and subsequently split into transversal and longitudinal components. The exponential decay of the cascades ρ1,2 sets in at about View the MathML source, implying cutoffs at 5.5 TeV for the ρ1 cascade and at 38 GeV for ρ2, which terminate the spectral plateaus. The unpolarized flux T+L is the actual spectral fit, the parameters of the electron densities are recorded in Table 1.


Table 1.

Electronic source densities ρ1,2 generating the γ-ray broadband of the TeV source HESS J1745 − 290 and the associated EGRET source 3EG J1746 − 2851 in the Galactic center region

β View the MathML source View the MathML source View the MathML source (TeV) U (erg)
ρ1 3.91×10−10 2.5×10−4 9.2×10 47 1310 5.8×10 51
ρ2 5.66×10−8 9.0×10−3 3.3×10 49 9.0 1.4×10 51

Each ρi stands for a Maxwell–Boltzmann density View the MathML source, cf. (4.13). β is the cutoff parameter in the Boltzmann factor. View the MathML source determines the amplitude of the tachyon flux generated by ρi, from which the electron count View the MathML source is inferred at the indicated distance of 8 kpc, cf. after (4.21). (The subscript 1 in View the MathML source and View the MathML source has been dropped.) View the MathML source is the electron temperature, and U (erg) the internal energy of the thermal densities ρi. The cascades labeled ρ1,2 in Fig. 1 and Fig. 2 are obtained by averaging the tachyonic radiation densities (4.1) with the electron densities ρi, cf. (4.14), (4.15) and (4.16). The parameters β and View the MathML source are extracted from the least-squares fit T+L in Fig. 1.


Full-size image

Fig. 2. Close-up of the HESS spectrum in Fig. 1. The TeV spectral map coincides with the ρ1 cascade, since the ρ2 flux is exponentially cut at 38 GeV. T and L stand for the transversal and longitudinal flux components, and T+L labels the unpolarized flux. The HESS points define a spectral plateau in the high GeV range typical for tachyonic cascade spectra [15] and [16]. The spectral curvature is intrinsic, being generated by the Boltzmann factor in the electron densities.


The quantized tachyonic radiation densities of a uniformly moving spinning charge read [16]

(4.1)
View the MathML source
where the superscripts T and L refer to the transversal/longitudinal polarization components defined by View the MathML source and View the MathML source. γ is the electronic Lorentz factor, αq the tachyonic fine structure constant, and mt the tachyon mass. A spectral cutoff occurs at

(4.2)
View the MathML source
Only frequencies in the range 0less-than-or-equals, slantωless-than-or-equals, slantωmax(γ) can be radiated by a uniformly moving charge, the tachyonic spectral densities View the MathML source being cut off at the break frequency ωmax. The units ħ=c=1 can easily be restored. We use the Heaviside–Lorentz system, so that αq=q2/(4πħc)≈1.0×10−13. The tachyon mass is View the MathML source, and the tachyon–electron mass ratio mt/m≈1/238. These estimates are obtained from hydrogenic Lamb shifts [17]. A positive ωmax(γ) requires Lorentz factors exceeding the threshold μt in (4.2), since ωmax(μt)=0. The lower threshold on the speed of the electron for radiation to occur is thus υmincolon, equalsmt/(2mμt). The tachyon–electron mass ratio gives υmin/c≈2.1×10−3, cf. Ref. [18].

The radiation densities (4.1) refer to a single charge with Lorentz factor γ. We average them with a Fermi power-law distribution,

(4.3)
View the MathML source
with normalization AFcolon, equalsm3V/π2, cf. (2.6), (2.7) and (2.8). The particle number is View the MathML source, where γ1 is the lower edge of Lorentz factors in the source population. The exponential cutoff in (4.3) is related to the electron temperature by View the MathML source. In this section, we write the exponent of the fugacity View the MathML source with a hat, cf. after (2.2), to distinguish it from the electron index customarily defined as α=δ−2, cf. (4.13). The normalization factor AF is dimensionless via mmc, cf. (1.1); the volume factor in the quantum corrections discussed in Section 3 is thus found as View the MathML source, where View the MathML source is the reduced electronic Compton wavelength.

The spectral average of the radiation densities (4.1) is carried out as

(4.4)
View the MathML source
where θ is the Heaviside step function. The spectral range of densities (4.1) is bounded by ωmax, so that the solution of View the MathML source defines the minimal electronic Lorentz factor for radiation at this frequency,

(4.5)
View the MathML source
The average (4.4) can be reduced to the fermionic spectral functions

(4.6)
View the MathML source
with lower integration boundary γ1μt, cf. after (4.2). The threshold Lorentz factor γ1 defines the break frequency [19]

(4.7)
View the MathML source
which separates the spectrum into a low- and high-frequency band. In particular, View the MathML source, and View the MathML source if ω>ω1. By making use of the spectral functions (4.6), we can write the averaged radiation densities (4.4) as

(4.8)
View the MathML source
with View the MathML source in (4.5) and ω1 in (4.7). The superscripts T and L denote the transversal and longitudinal radiation components, cf. (4.1). The spectral functions View the MathML source in (4.8) are assembled by substituting the radiation densities (4.1) into the integral representation (4.6),

(4.9)
View the MathML source
where the coefficients fk read

(4.10)
View the MathML source
with density View the MathML source in (4.3).

The quasiclassical fugacity expansion of the spectral functions (4.9) is found by expanding density (4.3) in ascending powers of View the MathML source, cf. (2.10),

(4.11)
View the MathML source
so that the asymptotic series of the reduced spectral functions fk(γ1) in (4.10) reads

(4.12)
View the MathML source
The classical limit of the fermionic density View the MathML source is defined by the leading term in series (4.11), the Boltzmann power-law density

(4.13)
View the MathML source
with normalization View the MathML source. (As mentioned after (4.3), exponent View the MathML source in the normalization factor defines the fugacity, and is not to be confused with the electron index α=δ−2.) The electron count based on the classical density (4.13) is View the MathML source, to be identified with the renormalized electron count View the MathML source obtained from the spectral fit, cf. after (4.21).

The classical limit of the fermionic spectral functions View the MathML source in (4.6) is the Boltzmann average [15]

(4.14)
View the MathML source
Here, View the MathML source is the classical spectral density, obtained by dropping all terms containing mt/m factors in (4.1) and (4.2). In particular, View the MathML source and View the MathML source, cf. (4.1). Carrying out the integration in (4.14), we find the classical spectral functions

(4.15)
View the MathML source
which can also be obtained from the Fermi functions View the MathML source in (4.9) by dropping the mt/m terms and substituting the leading order of the fugacity expansion of fk(γ1), cf. (4.12). The classical limit of the fermionic spectral average View the MathML source in (4.8) thus reads

(4.16)
View the MathML source
where View the MathML source is the classical limit of the break frequency (4.7).

In the low-temperature limit, βγ1much greater-than1, the incomplete View the MathML source-functions occurring in the fugacity expansion of fk(γ1) in (4.12) can be replaced by the asymptotic series

(4.17)
View the MathML source
At low temperatures, the quasiclassical fugacity expansion can only be used at sufficiently low densities, cf. the discussion of thermal wavelength following (3.4) and (3.15). Spectral averaging at low temperature and high density will be discussed elsewhere. The spectral fit discussed below is carried out in the high-temperature regime, βγ1much less-than1. In the fugacity expansion (4.12) of the reduced spectral functions, we can therefore substitute the ascending series

(4.18)
View the MathML source
If kδn is zero or a negative integer, kδn=−m, m≥0, we use instead of (4.18) the ascending series of the exponential integral Em+1(βγ1),

(4.19)
View the MathML source
The spectral fit in Fig. 1 and Fig. 2 is based on the E2-rescaled flux densities [20]

(4.20)
View the MathML source
where d is the distance to the source and View the MathML source the spectral average (4.16) (with ω=E). The fit is done with the unpolarized flux density View the MathML source of two electron populations ρi=1,2, thermal Maxwell–Boltzmann distributions (4.13) with α=−2 and γ1=1, cf. Table 1. Each electron density generates a cascade ρi, and the wideband fit is obtained by adding two cascade spectra, labeled ρ1 and ρ2 in Fig. 1. As for the electron count, View the MathML source, we use a rescaled parameter View the MathML source for the fit,

(4.21)
View the MathML source
which is independent of the distance estimate in (4.20). Here, View the MathML source implies the tachyon mass in keV units, that is, we put mt≈2.15 in the spectral density (4.1). At γ-ray energies, only a tiny View the MathML source-fraction (the ratio of tachyonic and electric fine structure constants) of the tachyon flux is absorbed by the detector, 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 to be View the MathML source, where View the MathML source defines the tachyonic flux amplitude extracted from the spectral fit [20]. This renormalized count View the MathML source is to be identified with the particle number N in the thermodynamic variables of the electron plasma. The electron temperature and cutoff parameter in the Boltzmann factor are related by View the MathML source, and the energy estimates in Table 1 are based on View the MathML source, cf. (3.9).

Fig. 1 and Fig. 2 show the spectral map of the TeV γ-ray source HESS J1745 − 290 and the GeV source 3EG J1746 − 2851. Both sources are located in the vicinity of the unidentified compact radio source Sagittarius (Sgr) A* at the core of the Galactic center region, which comprises the supernova remnant Sgr A East and the pulsar wind nebula G359.95 − 0.04, as well as molecular cloud complexes such as Sgr B and Sgr C [21] and [22].

In Table 1, we list the thermodynamic parameters of the electron populations generating the tachyonic cascade spectra ρ1,2 depicted in the figures. In the case of protonic source densities (4.13), the cutoff energy View the MathML source in the Boltzmann factor has to be multiplied by 1.84×103, the proton/electron mass ratio, resulting in a cutoff energy of 1016.2 eV for a protonic ρ2 population. The protonic high-energy population ρ1 is cut at 1018.4 eV, which is to be compared to the excess fluxes from the Galactic center region observed by AGASA in the 1018–1018.4 eV range [11], and the SUGAR array in the 1017.9–1018.5 eV interval [12].

The tachyonic spectral maps are further explained in the figure captions, and can be compared to electromagnetic inverse-Compton fits [23] or hadronic fits based on proton–proton scattering and pion decay [2] and [24]. The electronic source count for the Crab Nebula at View the MathML source is 1.9×1050, cf. Ref. [16], as compared to 3.4×1049 for HESS J1745 − 290 and 3EG J1746 − 2851, at a distance of 8 kpc, cf. Table 1. The internal energy of the electron gas is 7.2×1051 erg, to be compared to hadronic model estimates predicting 5×1045 erg [24], 1049–1050 erg [25], and 5.1×1050 erg [23] for a protonic source population.

5. Conclusion

We have investigated superluminal radiation from electron populations in the Galactic center region, and found the thermodynamic parameters of the source densities. We demonstrated that the γ-ray wideband of the Galactic center can be fitted with a tachyonic cascade spectrum. In particular, the extended spectral plateau in the MeV–GeV range as well as the spectral curvature apparent in double-logarithmic plots are reproduced by the cascade fit in Fig. 1 and Fig. 2. Estimates of the temperature, the source count, and the internal energy of the electron plasma generating the superluminal cascades are given in Table 1.

In Section 4, we averaged the superluminal radiation densities with fermionic power-law distributions, and derived the quantized spectral functions. The power-law densities in Sections Sections 2 and 3 and the spectral averages were studied mostly in the quasiclassical limit, as γ-ray spectral fits are done in the high-temperature regime. The opposite asymptotic limit of nearly degenerate fermionic power-law ensembles at low temperature and high density will be discussed elsewhere.

We quantized Boltzmann power-law densities in Fermi–Dirac statistics, derived the fermionic partition function, developed the Legendre formalism of fermionic power-law distributions, and calculated the fugacity expansion of the thermodynamic variables, cf. Section 2. The qualitative dependence of the variables on the electronic power-law index, the temperature scaling of the quantum corrections, and the thermal wavelength of power-law ensembles were investigated in Section 3. We have focused on γ-ray spectra, which can be fitted with ultra-relativistic electron densities at high temperature, cf. Section 4. As for tachyonic X-ray spectra obtained from diffraction gratings [21], the tachyon mass of 2 keV has to be included in the dispersion relation when parametrizing the glancing angle in the Bragg condition with energy, which affects the shape of the spectral maps in the X-ray bands; this will be discussed elsewhere. The spectral fit in Fig. 1 and Fig. 2 is performed with the unpolarized tachyon flux. At γ-ray energies, the speed of tachyons is close to the speed of light, the basic difference to electromagnetic radiation being the longitudinal flux component [26]. The polarization of tachyons can be determined from transversal and longitudinal ionization cross-sections of Rydberg atoms, which peak at different scattering angles [27].

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] S.P. Reynolds and J.W. Keohane, Astrophys. J. 525 (1999), p. 368. 

[2] K. Tsuchiya et al., Astrophys. J. 606 (2004), p. L115. 

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

[4] R. Tomaschitz, Physica A 385 (2007), p. 558.

[5] I. de Pater and B.J. Butler, Icarus 163 (2003), p. 428. 

[6] I. de Pater et al., Icarus 163 (2003), p. 434. 

[7] F. Aharonian et al., Astron. Astrophys. 425 (2004), p. L13.

[8] F. Aharonian et al., Phys. Rev. Lett. 97 (2006), p. 221102. 

[9] J. Albert et al., Astrophys. J. 638 (2006), p. L101.

[10] R.C. Hartman et al., Astrophys. J. Suppl. 123 (1999), p. 79.

[11] N. Hayashida et al., Astropart. Phys. 10 (1999), p. 303. 

[12] J.A. Bellido et al., Astropart. Phys. 15 (2001), p. 167.

[13] R. Tomaschitz, Physica A 307 (2002), p. 375. 

[14] S. Chandrasekhar, An Introduction to the Study of Stellar Structure, Dover, New York (1967).

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

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

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

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

[19] R. Tomaschitz, Eur. Phys. J. C 45 (2006), p. 493. 

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

[21] G. Bélanger et al., Astrophys. J. 636 (2006), p. 275.

[22] F. Yusef-Zadeh et al., Astrophys. J. 656 (2007), p. 847. 

[23] S. Liu et al., Astrophys. J. 647 (2006), p. 1099. 

[24] D.R. Ballantyne et al., Astrophys. J. 657 (2007), p. L13. 

[25] F. Aharonian et al., Nature 439 (2006), p. 695.

[26] R. Tomaschitz, Eur. Phys. J. D 32 (2005), p. 241. 

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

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