Next Article in Journal
Heating and Cooling in Transversely Oscillating Coronal Loops Powered by Broadband, Multi-Directional Wave Drivers
Next Article in Special Issue
Physical Mechanisms Underpinning the Vacuum Permittivity
Previous Article in Journal
The Impact of Radio Frequency Waves on the Plasma Density in the Tokamak Edge
Previous Article in Special Issue
New Insights into the Lamb Shift: The Spectral Density of the Shift
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Electron as a Tiny Mirror: Radiation from a Worldline with Asymptotic Inertia

by
Michael R. R. Good
1,2,* and
Yen Chin Ong
3,4
1
Department of Physics & Energetic Cosmos Laboratory, Nazarbaev University, Astana 010000, Qazaqstan
2
Leung Center for Cosmology & Particle Astrophysics, National Taiwan University, Taipei 10617, Taiwan
3
Center for Gravitation and Cosmology, College of Physical Science and Technology, Yangzhou University, Yangzhou 225002, China
4
Shanghai Frontier Science Center for Gravitational Wave Detection, School of Aeronautics and Astronautics, Shanghai Jiao Tong University, Shanghai 200240, China
*
Author to whom correspondence should be addressed.
Physics 2023, 5(1), 131-139; https://doi.org/10.3390/physics5010010
Submission received: 21 November 2022 / Revised: 17 December 2022 / Accepted: 3 January 2023 / Published: 28 January 2023
(This article belongs to the Special Issue Vacuum Fluctuations)

Abstract

:
We present a moving mirror analog of the electron, whose worldline possesses asymptotic constant velocity with corresponding Bogoliubov β coefficients that are consistent with finite total emitted energy. Furthermore, the quantum analog model is in agreement with the total energy obtained by integrating the classical Larmor power.

1. Introduction: Fixed Radiation

Uniform acceleration, while attractive and simple enough is not globally physical. Consider the problem of infinite radiation energy from an eternal uniformly accelerated charge. The physics of eternal unlimited motions is not only the cause of misunderstandings, but also the starting point of incorrect physical interpretations, especially when considering global calculable quantities, such as the total radiation emitted of a moving charge. Infinite radiation energy afflicts the accurate scrutiny of physical connections between acceleration, temperature, and particle creation.
More collective consideration should be prescribed to straighten out the issue. One path forward is the use of limited non-uniform accelerated trajectories that are capable of rendering finite global total radiation energy. The trade-off with these trajectories is usually the lack of simplicity or tractability in determining the radiation spectrum in the first place.
In this short paper, we present a solution for finite radiation energy and its corresponding spectrum. Limited solutions of this type are rare and can be employed to investigate the physics associated with contexts where a globally continuous equation of motion is desired. For instance, the solution is suited for applications such as the harvesting of entropy from a globally defined trajectory of an Unruh-DeWitt detector, the non-equilibrium thermodynamics of the non-uniform Davies-Fulling-Unruh effect, or the dynamical Casimir effect [1], and the particle production of the moving mirror model [2,3,4].
Providing a straightforward conceptual and quantitative analog application to understanding the radiation emitted by an electron, we demonstrate the existence of a correspondence (see similar correspondences in Refs. [5,6,7,8,9,10,11]) between the electron and the moving mirror. At the very least, this functional coincidence is general enough to be applied to any tractably integrable rectilinear classical trajectory that emits finite radiation energy. Here, we analytically compute the relevant integrable quantities for the specific solution and demonstrate full consistency. The analog approach treats the electron as a tiny moving mirror, somewhat similar to the Schwarzschild [12], Reissner–Nordström [13], and Kerr [14] black mirror analogies, but with the asymptotic inertia of a limited acceleration trajectory. Interestingly, the analog reveals previously unknown electron acceleration radiation spectra, thus helping to develop general but precise links between acceleration, gravity, and thermodynamics.

2. Elements of Electrodynamics: Energy from Moving Electrons

In electrodynamics [15,16,17], the relativistically covariant Larmor formula (the speed of light, c, the electron charge, q e , and vacuum permittivity, ϵ 0 , are set to unity),
P = α 2 6 π ,
is used to calculate the total power radiated by an accelerating point charge [15]. The Larmor formula’s usefulness is due in part to Lorentz invariance and proper acceleration, α , is intuitive, being what an accelerometer measures in the accelerometer’s own instantaneous rest frame [18].
When any charged particle accelerates, energy is radiated in the form of electromagnetic waves, and the total energy of these waves is found by integrating over coordinate time. That is, the time-integral,
E = P   d t ,
demonstrates that the Larmor power (1) immediately tells an observer the total energy emitted by a point charge along the point’s entire time-like worldline. This includes trajectories that lack horizons; see, e.g., [19]. This result is finite only when the proper acceleration is asymptotically zero, i.e., the worldline must be asymptotically inertial.
The force of radiation resistance, whose magnitude is given relativistically as the proper time, τ , derivative (notified by the prime) of the proper acceleration,
F = α ( τ ) 6 π ,
is known as the magnitude of the Lorentz–Abraham–Dirac (LAD) force, see, e.g., [20]. The power, F · v , associated with this force can be called the Feynman power [21]; here, v denotes the speed of the point. The total energy emitted is also consistent with the Feynman power, where one checks:
E = F · v   d t .
The negative sign demonstrates that the total work against the LAD force represents the total energy loss. That is, the total energy loss from radiation resistance due to Feynman power must equal the total energy radiated by the Larmor power (1). Larmor and Feynman powers are not the same, but the magnitude of the total energy from both are identical, at least for rectilinear trajectories that are asymptotically inertial.
Interestingly, the above results also hold in a quantum analog model of a moving mirror. A central novelty of this paper is to explicitly connect the quantum moving mirror radiation spectra with classical moving point charge radiation spectra. Traditionally (e.g., [2,3,4,22,23,24]) and recently (e.g., [25,26,27]), moving mirror models in ( 1 + 1 ) dimensions are employed to study the properties of Hawking radiation for black holes. Here, we show that it is also useful to model the spectral finite energy of electron radiation. In particular, a suitably constructed mirror trajectory (which is quite natural) can produce the same total energy consistent with Equation (4) via the Bogoliubov energy, (see, e.g., [22])
E = 0 0 ω | β ω ω | 2 d ω   d ω .
Here, β are the Bogoliubov β coefficients and ω and ω are the two sets of incoming and outgoing mode frequencies, respectively.
The final drifting speed, s, of the mirror or electron will be less than the speed of light: 0 < s < 1 . We also denote a : = ω ( 1 + s ) + ω ( 1 s ) , b : = ω ( 1 s ) + ω ( 1 + s ) , c : = a + b , and d : = a b ; note that then c = 2 ( ω + ω ) .

3. GO Trajectory for Finite Energy Emission

We consider a globally defined, continuous worldline, which is rectilinear, time-like, and possesses asymptotic zero velocity in the far past, while travelling to an asymptotically constant velocity in the far future (but asymptotically inertial both to the past and the future). It radiates a finite amount of positive energy and has Bogoliubov β coefficients that are analytically tractable. The Good-Ong (GO) trajectory, as defined by the authors [28], proceeds as follows:
z ( t ) = s 2 κ ln ( e 2 κ t + 1 ) ,
where κ is an acceleration parameter. The GO trajectory’s total power, applying the Larmor formula (1), is
P = 2 κ 2 3 π s 2 e 4 κ t 1 + e 2 κ t 2 1 + e 2 κ t 2 s 2 3 .
Let us note that the power is always positive and it asymptotically drops to zero, as Figure 1 illustrates.
The Feynman power, F · v , associated with the self-force (3), is
F · v = 2 κ 2 s 2 e 4 κ t j 1 e 6 κ t + j 2 e 4 κ t + e 2 κ t + 1 3 π j 1 e 4 κ t + 2 e 2 κ t + 1 3
where j 1 = s 2 1 and j 2 = 2 s 2 1 . Similar to the Larmor power (7), the Feynman power (8) asymptotically dies off, but unlike the Larmor power, the Feynman power has a period of negative radiation reaction, as illustrated in Figure 2.
Let us now compute the total energy using either the Larmor power or Feynman power, and integrating over time. In terms of the rapidity, η = tanh 1 v , and Lorentz factor, γ , the total energy is given by
E = κ 24 π γ 2 1 + η s 1 .
The second term, ( η s 1 ) = 1 2 s ln 1 + s 1 s 1 , is proportional to the lowest-order soft energy of inner bremsstrahlung in the case of beta decay (see Equation (3) in Ref. [11]), which is the deep infra-red contribution. One can see that Equation (9) is finite for all 0 < s < 1 and consistent with both the Larmor power and Feynman power. Below, after we compute the Bogoliubov β -spectrum (and plot it in Figure 3), we compute the Bogoliubov total energy (and plot it in Figure 4). We call Equation (9) the Larmor energy to differentiate it from the Bogoliubov energy (5), while substituting Equation (13) below. The energy is a function of the final constant speed, s.
Finally, the spectrum given by the Bogoliubov coefficients is best found by first considering the presence of a mirror in vacuum, e.g., [29,30]. The mode functions that correspond to the in-vacuum state,
ϕ ω in = 1 4 π ω e i ω v e i ω p ( u ) ,
and mode functions that correspond to the out-vacuum state,
ϕ ω out = 1 4 π ω e i ω f ( v ) e i ω u ,
comprise the two sets of incoming and outgoing modes needed for the Bogoliubov coefficients. The f ( v ) and p ( u ) functions express the trajectory (6) of the mirror, but in null coordinates, u = t z and v = t + z . In spacetime coordinates congruent with Equation (6), one form of the β -integral [19] for one side of the mirror is
β ω ω = d z e i ω n z i ω p t ( z ) 4 π ω ω ω p ω n t ( z ) ,
where ω p = ω + ω and ω n = ω ω . Combining the results for each side of the mirror [31] by adding the squares of the Bogoliubov β coefficients ensures that one accounts for all of the radiation emitted by the mirror [32]. The overall count per double-mode is
| β ω ω | 2 = s 2 ω ω Z 2 π a b c d κ e π d 4 κ 1 e π c 4 κ 1 e π b 4 κ ,
where Z = b csc h π a 4 κ + a csc h π b 4 κ . Equation (13) combines the squares, | β R | 2 + | β L | 2 , of the coefficients for left (L) and right (R) sides of mirror [28]. Figure 3 shows a plot of the symmetry between the modes ω and ω .
It is then straightforward to verify that the total energy obtained by integrating the power is the same as by using the Bogoliubov β integral (5). We cannot prove this analytically, but a numerical integral is quite convincing; see Figure 4 for a plot of the Larmor and Bogoliubov energies as a function of final constant speed.

4. Discussions: Mirrors, Electrons, and Black Holes

Prior studies of accelerated electrons and their relationship to mirrors are few; however, several papers, e.g., [33,34], connect electrons to the general Davies-Fulling-Unruh effect (for example, perhaps the most known one is by Bell and Leinaas [35], which considered the possibility of using accelerated electrons as thermometers to demonstrate the relationship between acceleration and temperature). Nevertheless, perhaps an early clue a that functional identity existed was made in 1982 by Ford and Vilenkin [24], who found that the LAD self-force was the same form for both mirrors and electrons. In 1995, Nikishov and Ritus [5] asserted the spectral symmetry and found that the LAD radiation reaction has a term that corresponds to the negative energy flux (NEF) from moving mirrors. Ritus examined [6,7,8] the correspondence connecting the radiation from both the electron and mirror systems, claiming not only a deep symmetry between the two, but a fundamental identity related to bare charge quantization [10]. Recently, the duality was extended to Larmor power [32] and deep infrared radiation [11]. The approach has pedagogical application; for instance, it was used to demonstrate the physical difference between radiation power loss and kinetic power loss [9].
The GO moving mirror was initially constructed to model the evaporation of black holes that exhibit a “death gasp” [36,37,38]—an emission of NEF due to unitarity being preserved. Therefore, it is a sturdy result that the total finite energy emitted from the double-sided mirror matches the result from the Larmor formula for an electron. In this sense, the GO mirror trajectory represents a crude but functional analog of a drifting electron that starts at zero velocity and speeds away to some constant velocity. A single-sided moving mirror does not account for all of the radiation emitted by an electron, and as such, the single-sided mirror radiation spectrum differs from the electron spectrum. A notable difference: there is no known NEF radiated from an electron.
In the literature, one finds some properties that are shared by black holes and electrons. For example, the ratio of the magnetic moment of an electron (of mass m and charge e) to its spin angular momentum is g e / 2 m with g = 2 , which is twice the value of the gyromagnetic ratio for classical rotating charged bodies ( g = 1 ). Curiously, as Carter has shown [39], a Kerr–Newmann black hole also has g = 2 . This has led to some speculations as to whether the electron is a Kerr–Newman singularity (the angular momentum and charge of the electron are too large for a black hole of the electron’s mass, so there is no horizon) [40] (see also [41,42]). The no-hair property of black holes is also similar to elementary particle indistinguishably: all electrons look the same. Certainly, the mirror model considered here looks too simple to seek certain further connections between particle physics and black holes; in particular, it does not involve any charge or angular momentum. Nevertheless, precisely because of this, it is surprising the total emitted energy in the model should be given by the integral of the Larmor formula.
Near-term possible theoretical applications of electron-mirror correspondence include extension to non-rectilinear trajectories; notably, the uniform accelerated worldlines of Letaw [43], which have Unruh-like temperatures [44] and power distributions [45]. Applying the general study of Kothawala and Padmanabhan [46] to electrons moving along time-dependent accelerations, and comparing the effect to an Unruh-DeWitt detector may prove to be fruitful for understanding the thermal response. Moreover, moving mirror models can be useful in cosmology [47], in particular, in modeling particle production due to the expansion of space [48]. This expansion is accelerated due to an unknown dark energy, which may not be a cosmological constant and thus can decay [49,50]. If dark energy is some kind of vacuum energy, it might be subject to further study from mirror analogs just like Casimir energy. (Actually, dark energy could be Casimir-like [51].)
Near-term possible experimental applications of electron-mirror holography include leveraging the correspondence to disentangle effects in experiments like in the Analog Black Hole Evaporation via Lasers (AnaBHEL) [52] and the RDK II [53,54] experiments (see also [55]). The former exploits the accelerating relativistic moving mirror as a probe of the spectrum of quantum vacuum radiation [56,57], and the latter measures the photon spectrum with high precision as the electron-mirror is subjected to extreme accelerations during the process of radiative neutron beta decay.

Author Contributions

Both authors have contributed equally to the work. All authors have read and agreed to the published version of the manuscript.

Funding

M.R.R.G. thanks the FY2021-SGP-1-STMM Faculty Development Competitive Research Grant No. 021220FD3951 at Nazarbayev University (Astana, Qazaqstan). Y.C.O. thanks the National Natural Science Foundation of China (No. 11922508) for funding support.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Moore, G.T. Quantum theory of the electromagnetic field in a variable-length one-dimensional cavity. J. Math. Phys. 1970, 11, 2679–2691. [Google Scholar] [CrossRef]
  2. DeWitt, B.S. Quantum field theory in curved space-time. Phys. Rep. 1975, 19, 295–357. [Google Scholar] [CrossRef]
  3. Fulling, S.A.; Davies, P.C.W. Radiation from a moving mirror in two dimensional space-time: Conformal anomaly. Proc. R. Soc. Lond. A 1976, 348, 393–414. [Google Scholar] [CrossRef]
  4. Davies, P.C.W.; Fulling, S.A. Radiation from moving mirrors and from black holes. Proc. R. Soc. Lond. A 1977, 356, 237–257. [Google Scholar] [CrossRef]
  5. Nikishov, A.; Ritus, V. Emission of scalar photons by an accelerated mirror in (1+1) space and its relation to the radiation from an electrical charge in classical electrodynamics. J. Exp. Theor. Phys. 1995, 81, 615–624. Available online: http://jetp.ras.ru/cgi-bin/e/index/e/81/4/p615?a=list (accessed on 30 December 2022).
  6. Ritus, V. The Symmetry, inferable from Bogoliubov transformation, between the processes induced by the mirror in two-dimensional and the charge in four-dimensional space-time. J. Exp. Theor. Phys. 2003, 97, 10–23. [Google Scholar] [CrossRef] [Green Version]
  7. Ritus, V. Vacuum-vacuum amplitudes in the theory of quantum radiation by mirrors in 1+1-space and charges in 3+1-space. Int. J. Mod. Phys. A 2002, 17, 1033–1040. [Google Scholar] [CrossRef]
  8. Ritus, V. Symmetries and causes of the coincidence of the radiation spectra of mirrors and charges in (1+1) and (3+1) spaces. J. Exp. Theor. Phys. 1998, 87, 25–34. [Google Scholar] [CrossRef]
  9. Good, M.R.R.; Singha, C.; Zarikas, V. Extreme electron acceleration with fixed radiation energy. Entropy 2022, 24, 1570. [Google Scholar] [CrossRef]
  10. Ritus, V.I. Finite value of the bare charge and the relation of the fine structure constant ratio for physical and bare charges to zero-point oscillations of the electromagnetic field in the vacuum. Phys.-Usp. 2022, 65, 468–486. [Google Scholar] [CrossRef]
  11. Good, M.R.R.; Davies, P.C.W. Infrared acceleration radiation. arXiv 2022, arXiv:2206.07291. [Google Scholar] [CrossRef]
  12. Good, M.R.R.; Anderson, P.R.; Evans, C.R. Mirror reflections of a black hole. Phys. Rev. D 2016, 94, 065010. [Google Scholar] [CrossRef] [Green Version]
  13. Good, M.R.R.; Ong, Y.C. Particle spectrum of the Reissner–Nordström black hole. Eur. Phys. J. C 2020, 80, 1169. [Google Scholar] [CrossRef]
  14. Good, M.R.R.; Foo, J.; Linder, E.V. Accelerating boundary analog of a Kerr black hole. Class. Quantum Grav. 2021, 38, 085011. [Google Scholar] [CrossRef]
  15. Jackson, J.D. Classical Electrodynamics; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 1999; Available online: https://www.scribd.com/doc/200995304/Classical-Electrodynamics-3rd-Ed-1999-John-David-Jackson (accessed on 30 December 2022).
  16. Zangwill, A. Modern Electrodynamics; Cambridge University Press: Cambridge, UK, 2012; Available online: https://faculty.kashanu.ac.ir/file/download/page/1604994434-modern-electrodynamics.pdf (accessed on 30 December 2022).
  17. Griffiths, D.J. Introduction to Electrodynamics; Cambridge University Press: New York, NY, USA, 2017. [Google Scholar] [CrossRef]
  18. Rindler, W.A. Essential Relativity: Special, General and Cosmological; Springer-Verlag: Berlin/Heidelber, Germany, 1977. [Google Scholar] [CrossRef]
  19. Good, M.R.R.; Yelshibekov, K.; Ong, Y.C. On horizonless temperature with an accelerating mirror. J. High Energy Phys. 2017, 3, 13. [Google Scholar] [CrossRef] [Green Version]
  20. Myrzakul, A.; Xiong, C.; Good, M.R.R. CGHS black hole analog moving mirror and its relativistic quantum information as radiation reaction. Entropy 2021, 23, 1664. [Google Scholar] [CrossRef]
  21. Feynman, R.P. Feynman Lectures on Gravitation; CRC Press/Taylor & Francis Group: Boca Raton, FL, USA, 2003. [Google Scholar] [CrossRef]
  22. Walker, W.R. Particle and energy creation by moving mirrors. Phys. Rev. D 1985, 31, 767–774. [Google Scholar] [CrossRef]
  23. Carlitz, R.D.; Willey, R.S. Reflections on moving mirrors. Phys. Rev. D 1987, 36, 2327–2335. [Google Scholar] [CrossRef]
  24. Ford, L.H.; Vilenkin, A. Quantum radiation by moving mirrors. Phys. Rev. D 1982, 25, 2569–2575. [Google Scholar] [CrossRef]
  25. Good, M.R.; Linder, E.V.; Wilczek, F. Moving mirror model for quasithermal radiation fields. Phys. Rev. D 2020, 101, 025012. [Google Scholar] [CrossRef] [Green Version]
  26. Moreno-Ruiz, A.; Bermudez, D. Optical analogue of the Schwarzschild–Planck metric. Class. Quant. Grav. 2022, 39, 145001. [Google Scholar] [CrossRef]
  27. Good, M.R.R.; Linder, E.V. Modified Schwarzschild metric from a unitary accelerating mirror analog. New J. Phys. 2021, 23, 043007. [Google Scholar] [CrossRef]
  28. Good, M.R.R.; Ong, Y.C. Signatures of energy flux in particle production: A black hole birth cry and death gasp. J. High Energy Phys. 2015, 7, 145. [Google Scholar] [CrossRef]
  29. Birrell, N.D.; Davies, P.C.W. Quantum Fields in Curved Space; Cambridge Univercity Press: New York, NY, USA, 1982. [Google Scholar] [CrossRef]
  30. Good, M.R.R. Extremal Hawking radiation. Phys. Rev. D 2020, 101, 104050. [Google Scholar] [CrossRef]
  31. Good, M.R.R. Reflecting at the speed of light. In Memorial Volume for Kerson Huang; Phua, K.K., Low, H.B., Xiong, C., Eds.; World Scientific Co., Ltd.: Singapore, 2017; pp. 113–116. [Google Scholar] [CrossRef] [Green Version]
  32. Zhakenuly, A.; Temirkhan, M.; Good, M.R.R.; Chen, P. Quantum power distribution of relativistic acceleration radiation: Classical electrodynamic analogies with perfectly reflecting moving mirrors. Symmetry 2021, 13, 653. [Google Scholar] [CrossRef]
  33. Myhrvold, N.P. Thermal radiation from accelerated electrons. Ann. Phys. 1985, 160, 102–113. [Google Scholar] [CrossRef]
  34. Paithankar, K.; Kolekar, S. Role of the Unruh effect in Bremsstrahlung. Phys. Rev. D 2020, 101, 065012. [Google Scholar] [CrossRef] [Green Version]
  35. Bell, J.S.; Leinaas, J.M. Electrons as accelerated thermometers. Nucl. Phys. B 1983, 212, 131–150. [Google Scholar] [CrossRef] [Green Version]
  36. Bianchi, E.; Smerlak, M. Last gasp of a black hole: Unitary evaporation implies non-monotonic mass loss. Gen. Rel. Grav. 2014, 46, 1809. [Google Scholar] [CrossRef] [Green Version]
  37. Bianchi, E.; Smerlak, M. Entanglement entropy and negative energy in two dimensions. Phys. Rev. D 2014, 90, 041904. [Google Scholar] [CrossRef] [Green Version]
  38. Abdolrahimi, S.; Page, D.N. Hawking radiation energy and entropy from a Bianchi-Smerlak semiclassical black hole. Phys. Rev. D 2015, 92, 083005. [Google Scholar] [CrossRef] [Green Version]
  39. Carter, B. Global structure of the Kerr family of gravitational fields. Phys. Rev. 1968, 174, 1559–1571. [Google Scholar] [CrossRef] [Green Version]
  40. Burinskii, A. The Dirac-Kerr electron. Grav. Cosmol. 2008, 14, 109–122. [Google Scholar] [CrossRef] [Green Version]
  41. Schmekel, B.S. Quasilocal energy of a charged rotating object described by the Kerr-Newman metric. Phys. Rev. D 2019, 100, 124011. [Google Scholar] [CrossRef] [Green Version]
  42. Burinskii, A. The Dirac electron consistent with proper gravitational and electromagnetic field of the Kerr–Newman solution. Galaxies 2021, 9, 18. [Google Scholar] [CrossRef]
  43. Letaw, J.R. Vacuum excitation of noninertial detectors on stationary world lines. Phys. Rev. D 1981, 23, 1709–1714. [Google Scholar] [CrossRef]
  44. Good, M.; Juárez-Aubry, B.A.; Moustos, D.; Temirkhan, M. Unruh-like effects: Effective temperatures along stationary worldlines. J. High Energy Phys. 2020, 06, 059. [Google Scholar] [CrossRef]
  45. Good, M.R.R.; Temirkhan, M.; Oikonomou, T. Stationary worldline power distributions. Int. J. Theor. Phys. 2019, 58, 2942–2968. [Google Scholar] [CrossRef] [Green Version]
  46. Kothawala, D.; Padmanabhan, T. Response of Unruh-DeWitt detector with time-dependent acceleration. Phys. Lett. B 2010, 690, 201–206. [Google Scholar] [CrossRef] [Green Version]
  47. Good, M.R.R.; Zhakenuly, A.; Linder, E.V. Mirror at the edge of the universe: Reflections on an accelerated boundary correspondence with de Sitter cosmology. Phys. Rev. D 2020, 102, 045020. [Google Scholar] [CrossRef]
  48. Castagnino, M.; Ferraro, R. A toy cosmology: Radiation from moving mirrors, the final equilibrium state and the instantaneous model of particle. Ann. Phys. 1985, 161, 1–20. [Google Scholar] [CrossRef]
  49. Akhmedov, E.T.; Buividovich, P.V.; Singleton, D.A. De Sitter space and perpetuum mobile. Phys. Atom. Nucl. 2012, 75, 525–529. [Google Scholar] [CrossRef] [Green Version]
  50. Polyakov, A.M. Decay of vacuum energy. Nucl. Phys. B 2010, 834, 316–329. [Google Scholar] [CrossRef] [Green Version]
  51. Leonhardt, U. The case for a Casimir cosmology. Philos. Trans. R. Soc. A 2020, 378, 20190229. [Google Scholar] [CrossRef] [PubMed]
  52. Chen, P. et al. [AnaBHEL Collaboration] AnaBHEL (Analog Black Hole Evaporation via Lasers) experiment: Concept, design, and status. Photonics 2022, 9, 1003. [Google Scholar] [CrossRef]
  53. Nico, J.S.; Dewey, M.S.; Gentile, T.R.; Mumm, H.P.; Thompson, A.K.; Fisher, B.M.; Kremsky, I.; Wietfeldt, F.E.; Chupp, T.E.; Cooper, R.L.; et al. Observation of the radioactive decay mode of the free neutron. Nature 2006, 444, 1059–1062. [Google Scholar] [CrossRef] [PubMed]
  54. Bales, M.J. et al. [RDK II Collaboration] Precision measurement of the radiative β decay of the free neutron. Phys. Rev. Lett. 2016, 116, 242501. [Google Scholar] [CrossRef] [Green Version]
  55. Lynch, M.H.; Good, M.R.R. Experimental observation of a moving mirror. arXiv 2022, arXiv:2211.14774. [Google Scholar] [CrossRef]
  56. Chen, P.; Mourou, G. Trajectory of a flying plasma mirror traversing a target with density gradient. Phys. Plasmas 2020, 27, 123106. [Google Scholar] [CrossRef]
  57. Chen, P.; Mourou, G. Accelerating plasma mirrors to investigate black hole information loss paradox. Phys. Rev. Lett. 2017, 118, 045001. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The Larmor power (1) of the Good-Ong (GO) trajectory (6) as a function of time, t, and at final constant speed, s = 0.9 , i.e., Equation (7), with acceleration parameter, κ = 1 . This plot illustrates that the Larmor power never emits negative energy flux (NEF) and asymptotically dies off, consistent with a physically finite amount of total radiation energy (2).
Figure 1. The Larmor power (1) of the Good-Ong (GO) trajectory (6) as a function of time, t, and at final constant speed, s = 0.9 , i.e., Equation (7), with acceleration parameter, κ = 1 . This plot illustrates that the Larmor power never emits negative energy flux (NEF) and asymptotically dies off, consistent with a physically finite amount of total radiation energy (2).
Physics 05 00010 g001
Figure 2. The Feynman power, F · v , associated with the self-force, F (3), of the GO trajectory (6), as a function of time and at final constant speed, s = 0.9 , i.e., Equation (8), with κ = 1 . This plot illustrates the Feynman power dies off asymptotically, has a period of negative radiation reaction, and is also consistent with a physically finite amount of total radiation energy (4).
Figure 2. The Feynman power, F · v , associated with the self-force, F (3), of the GO trajectory (6), as a function of time and at final constant speed, s = 0.9 , i.e., Equation (8), with κ = 1 . This plot illustrates the Feynman power dies off asymptotically, has a period of negative radiation reaction, and is also consistent with a physically finite amount of total radiation energy (4).
Physics 05 00010 g002
Figure 3. A contour plot of the coefficients (13) as a function of in and out modes, ω and ω , where final constant speed, s = 0.444 and κ = 1 . The color gradient darkens for lower values of the count. This plot underscores the symmetry of the modes in the particle per mode squared distribution spectrum of the Bogoliubov β coefficients (13).
Figure 3. A contour plot of the coefficients (13) as a function of in and out modes, ω and ω , where final constant speed, s = 0.444 and κ = 1 . The color gradient darkens for lower values of the count. This plot underscores the symmetry of the modes in the particle per mode squared distribution spectrum of the Bogoliubov β coefficients (13).
Physics 05 00010 g003
Figure 4. The Larmor energy (9) and Bogoliubov energy (5) as a function of final constant speed, s. Here, 0 < s < 0.99 and κ = 1 . This plot confirms that Larmor and Bogoliubov energies are equivalent, substantiating the double-sided moving mirror as an analog model of the electron.
Figure 4. The Larmor energy (9) and Bogoliubov energy (5) as a function of final constant speed, s. Here, 0 < s < 0.99 and κ = 1 . This plot confirms that Larmor and Bogoliubov energies are equivalent, substantiating the double-sided moving mirror as an analog model of the electron.
Physics 05 00010 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Good, M.R.R.; Ong, Y.C. Electron as a Tiny Mirror: Radiation from a Worldline with Asymptotic Inertia. Physics 2023, 5, 131-139. https://doi.org/10.3390/physics5010010

AMA Style

Good MRR, Ong YC. Electron as a Tiny Mirror: Radiation from a Worldline with Asymptotic Inertia. Physics. 2023; 5(1):131-139. https://doi.org/10.3390/physics5010010

Chicago/Turabian Style

Good, Michael R. R., and Yen Chin Ong. 2023. "Electron as a Tiny Mirror: Radiation from a Worldline with Asymptotic Inertia" Physics 5, no. 1: 131-139. https://doi.org/10.3390/physics5010010

Article Metrics

Back to TopTop