Skip to main content
Log in

Connecting Blackbody Radiation, Relativity, and Discrete Charge in Classical Electrodynamics

  • Published:
Foundations of Physics Aims and scope Submit manuscript

It is suggested that an understanding of blackbody radiation within classical physics requires the presence of classical electromagnetic zero-point radiation, the restriction to relativistic (Coulomb) scattering systems, and the use of discrete charge. The contrasting scaling properties of nonrelativistic classical mechanics and classical electrodynamics are noted, and it is emphasized that the solutions of classical electrodynamics found in nature involve constants which connect together the scales of length, time, and energy. Indeed, there are analogies between the electrostatic forces for groups of particles of discrete charge and the van der Waals forces in equilibrium thermal radiation. The differing Lorentz- or Galilean-transformation properties of the zero-point radiation spectrum and the Rayleigh-Jeans spectrum are noted in connection with their scaling properties. Also, the thermal effects of acceleration within classical electromagnetism are related to the existence of thermal equilibrium within a gravitational field. The unique scaling and phase-space properties of a discrete charge in the Coulomb potential suggest the possibility of an equilibrium between the zero-point radiation spectrum and matter which is universal (independent of the particle mass), and an equilibrium between a universal thermal radiation spectrum and matter where the matter phase space depends only upon the ratio mc 2/k B T. The observations and qualitative suggestions made here run counter to the ideas of currently accepted quantum physics.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Thermal effects of acceleration appeared first in work on quantum field theory by P. C. Davies, J. Phys. A 8, 609–616 (1975) and by W. G. Unruh, Phys. Rev. D 14, 870–892 (1976).

  2. T. H. Boyer, Phys. Rev. D 21, 2137–2148 (1980); Phys. Rev. D 29, 1089–1095 (1984).

    Google Scholar 

  3. T. H. Boyer, Phys. Rev. D 29, 1096–1098 (1984); Phys. Rev. D 30, 1228–1232 (1984).

    Google Scholar 

  4. D. C. Cole, Phys. Rev. D 31, 1972–1981 (1985).

    Google Scholar 

  5. Some aspects of the symmetries have been given by J. Haantjes, Koninkl. Ned. Akad. Wetenschap. Proc. 43, 1288–1299 (1940) and by T. H. Boyer, J. Phys. A 38, 1807–18221 (2005).

  6. See, for example, R. Eisberg and R. Resnick, Quantum Physics of Atoms, Molecules, Solids, Nuclei, and Particles, 2nd ed, (Wiley, New York, 1985), p. 12. In 1905, James Jeans was so convinced of the applicability to radiation of the energy equipartition law of nonrelativistic classical statistical mechanics that he looked to a failure of the high frequency modes to reach thermal equilibrium rather than to a failure of equipartition. See, J. Jeans, Proc. Roy. Soc (London) A 76, 545–552 (1905).

  7. See, for example, B. H. Lavenda, Statistical Physics: A Probabilistic Approach, (Wiley, New York, 1991), pp. 73–74, and F. Reif, Fundamentals of Statistical and Thermal Physics, (McGraw-Hill, New York, 1965), pp. 251–252.

  8. J. H. Van Vleck, Phys. Rev. 24, 347–365 (1924). T. H. Boyer, Phys. Rev. D 13, 2832–2845 (1976). T. H. Boyer, Phys. Rev. A 18, 1228–1237 (1978).

  9. T. S. Kuhn, Black-Body Theory and the Quantum Discontinuity 1894–1912, (Oxford U. Press, New York, 1978), Chapter VIII.

  10. T. W. Marshall, Proc. Camb. Phil. Soc. 61, 537–546 (1965); T. H. Boyer, Phys. Rev. 182, 1374–1383 (1969).

  11. Boyer T.H. (1975). Phys. Rev. D 11:790–808

    Article  ADS  Google Scholar 

  12. M. J. Sparnaay, Physica 24, 751–764 (1958). S. K. Lamoreaux, Phys. Rev. Lett. 78, 5–8 (1997); 81, 5475–5476 (1998). U. Mohideen, Phys. Rev. Lett. 81, 4549–4552 (1998). H. B. Chan, V. A. Aksyuk, R. N. Kleiman, and F. Capasso, Science 291, 1941–1944 (2001). G. Bressi, G. Carugno, R. Onofrio, and G. Ruoso, Phys. Rev. Lett. 88, 041804(4) (2002).

  13. The analogy between the electrostatic scaling behavior and the electromagnetic scaling behavior provides the basis for Casimir’s model of the electron; H. B. G. Casimir, Physica 19, 846–849 (1956). However, Casimir does not seem to have realized that the scaling analogy holds more generally.

  14. The parallel-plate calculation was first made by H. B. G. Casimir, Koninkl. Ned. Akad. Wetenschap, Proc. 51, 793–795 (1948).

  15. The spherical case was first calculated by T. H. Boyer. Phys. Rev. 174, 1764–1776 (1968).

    Google Scholar 

  16. Because of historical accident, Stefan’s constant is usually rewritten using \(a_{s}=k_{B}^{4}\pi^{2}/(15\hbar^{3}c^{3})\). See, for example, P. H. Morse, Thermal Physics, 2nd ed., (Benjamin/Cummings Publishing, Reading, Mass. 1969), p. 339. Then the dimensionless universal constant is given as \(e^{2}(a_{S}/k_{B}^{4})^{1/3}=e^{2}\{[\pi^{2}k_{B}^{4}/(15\hbar^{3}c^{3})]/k_{B}^{4}\}^{1/3}=(\pi^{2}/15)^{1/3}[e^{2}/(\hbar c)]=0.0063487.\) Thus we recognize the dimensionless constant as related to Sommerfeld’s fine-structure constant. However, for many physicists (including both the physicists of the early twentieth century and also recent referees), the use of Planck’s constant \(\hbar\) rather than Stefan’s constant a s serves as an overwhelming distraction. The appearance of Planck’s constant seems to cause an immediate fixation on ideas of energy quanta. There are no ideas of energy quanta in the present analysis, and such ideas have no place in classical physics. Actually, Planck’s constant does not necessarily have any connection to energy quanta; it can serve as the scale factor for classical electromagnetic zero point radiation without in any way implying energy quanta. See, for example, ref. 11.

  17. Readers will recognize b s in terms of currently familiar constants as \(b_{s}=(1/2)\hbar c\). We emphasize that no energy quanta whatsoever are used in our classical analysis.

  18. See, for example, L. L. Henry and T. W. Marshall, Nuovo Cimento 41, 188–197 (1966); T. H. Boyer, Phys. Rev. A 9, 2078–2084 (1974).

  19. S. O. Rice, in Selected Papers on Noise and Stochastic Processes, N. Wax ed. (Dover, New York, 1954), p. 133.

  20. See, for example, E. A. Power, Introductory Quantum Electrodynamics (American Elsevier, New York, 1964), pp. 18–22.

  21. Boyer T.H. (1975). Phys. Rev. A 18:1238–1245

    Article  ADS  Google Scholar 

  22. The existence of energy fluctuations which involve no entropy requires ideas of special relativity and is foreign to nonrelativistic classical statistical mechanics.

  23. E. Cunningham, Proc. London Math. Soc. 8, 77–98 (1910); H. Bateman, “The transformation of the electrodynamical equations,” Proc. London Math. Soc. 8, 223–264 (1910).

    Google Scholar 

  24. T. H. Boyer, Found. Phys. 19, 349–365 (1989), Eq. (40).

  25. The correlation function can be obtained by following the same pattern as given in ref. 24 except that we replace (1/2)\(\hbar c|k|\) by k B T in Eq. (39). The integral then involves \(\int {(d^3 k/|k|\cos [{\text{k}} \cdot {\text{(r - r')}} - c|k|(t - t')] = \int\limits_0^\infty {dk} \int\limits_0^\pi {d\theta } \sin \theta \int\limits_0^{2\pi } {d\phi } \cos [k(R\cos \theta - c\tau )] = \int\limits_0^\infty {(dk} /k)\{ {\text{sin[}}k(R - c\tau )] + } {\text{sin[}}k(R + c\tau )]\} .\) Finally the definite integral gives \(\int\limits_0^\infty {dx} (\sin mx)/x = \pi /2 {\text{if }} m > 0, 0 {\text{if}} m = 0\) and π/2 if m < 0

  26. Boyer T.H. (2003). J Phys A 36:7425–7440

    Article  MATH  ADS  MathSciNet  Google Scholar 

  27. As far as the quantum theorists are concerned, the introduction of quanta solved the blackbody problem in conjunction with statistical mechanics. Thus they are surprised when the Planck spectrum appears without the use of quantum statistics. The point of view presented here is entirely classical and is quite different. In the classical perspective present here, classical statistical mechanics plays no role but rather is a derived concept, derived from the randomness of random phases of waves, in the sense of ref. 19. The connection between conformal motions and gravity is crucial in connecting the Planck spectrum to thermal radiation. Thus if thermal radiation exists in equilibrium in a gravitational field, it must correspond to the radiation associated with uniform acceleration through zero-point radiation plus perhaps additional radiation. The finite temperature radiation must fit on top of the spectrum from acceleration through zero-point radiation. In order to satisfy Wien’s displacement theorem, the additional radiation must follow the same pattern which already appeared from the acceleration through classical zero-point radiation.

  28. W. Rindler, Essential Relativity: Special, General, and Cosmological, 2nd ed. (Springer-Verlag, New York, 1977), pp. 49–51.

  29. During the early years of the 20th century, the nonrelativistic harmonic oscillator was often connected to radiation by a very small charge q. See, for example, M. Planck, The Theory of Heat Radiation (Dover, New York, 1959). The electric dipole oscillator was said to come to equilibrium with an energy U equal to the energy \(U_{\omega_{0}}\) of the radiation mode at the same frequency as the natural frequency of the oscillator, \(U=U_{\omega_{0}}\). This equality held for fixed \(\omega_{0}\) whether the mass m and spring constant K of the oscillator were both large and the velocity and amplitude were both small, or whether m and K were both small with the velocity and amplitude large, perhaps the velocity even exceeding c. However, classical electromagnetic radiation does not treat these two possible radiating systems alike because finite velocity and amplitude means that radiation will be emitted into the higher harmonics of the fundamental frequency. The traditional classical calculations are accurate only in the limit \(m\rightarrow\infty\) where the velocity and amplitude vanish; in this limit, the oscillator interacts with radiation only at its fundamental frequency and so does not scatter radiation toward equilibrium. This corresponds to the electric dipole limit.

  30. H. A. Lorentz, The Theory of Electrons and its Application to the Phenomena of Light and Radiation Heat, 2nd ed. (Dover, New York, 1952), p. 96. This is a republication of the 2nd edition of 1915.

  31. T. H. Boyer, Found. Phys. 19, 1371–1383 (1989). D. C. Cole, Found. Phys. 20, 225–240 (1989).

  32. See, for example, H. Goldstein, Classical Mechanics, 2nd ed., (Addison-Wesley Publishing, Reading, Mass. 1981), p. 498. We are using action variables which are 1/(2π) times those in Goldstein’s text.

  33. Boyer T.H. (2004). Am. J. Phys. 72:992–997

    Article  ADS  MathSciNet  Google Scholar 

  34. J. D. Jackson, Classical Electrodynamics, 2nd. ed., (Wiley, New York, 1975), p. 664.

  35. See ref. 34, p. 695.

  36. Burko L.M. (2000). Am. J. Phys. 68:456–468

    Article  ADS  Google Scholar 

  37. A complete classical electromagnetic calculation of radiation equilibrium for a relativistic particle in the Coulomb potential at zero temperature should give the value of the fine structure constant. The probability distribution for the action variables J i involves two different scales; e 2/c and b s /c. The scale e 2/c is related to the relativistic mechanical behavior and provides a limit at small J 2. See ref. 33. The scale b s /c involves the energy balance with the random zero-point radiation. Roughly, when the action variables J i are larger than b s /c, then the hydrogen system is losing more energy by radiation emission than it is picking up from the zero-point radiation; when the J i are smaller than b s /c, the reverse is true. See section D2 of ref. 11. We note that if b s /c were too small, \(b_{s}/c < e^{2}/c\), then no hydrogen atom would exist because the energy-conserving orbits at this angular momentum would spiral into the nucleus. See ref. 33. In nature, the value is \(b_{s}/c\approx68e^{2}/c.\)

  38. See ref. 30, note 6, p. 240, which gives Lorentz’s explicit assumption on the boundary condition.

  39. A review of most of the work on classical zero-point radiation is presented by L. de la Pena and A. M. Cetto, The Quantum Dice - An Introduction to Stochastic Electrodynamics (Kluwer Academic, Dordrecht, 1996).

  40. D. C. Cole and Y. Zou, Phys. Lett. A 317, 14–20 (2003). Cole and Zou’s numerical simulation calculations suggest that classical zero-point radiation acting on the classical model for hydrogen (a point charge in a Coulomb potential) produces a steady-state probability distribution which is finite and which seems to approach the ground state obtained from solution of the Schroedinger equation for a point charge in the Coulomb potential. Cole and Zou have no adjustable parameters in their calculations. See also, T. H. Boyer, Found. Phys. Lett. 16, 613–617 (2003).

  41. See ref. 9, p. 132.

  42. On page 78 of ref. 30, Lorentz writes, “Now, if the state of radiation is produced by a ponderable body, the values of the two constants [in the blackbody spectrum] must be determined by something in the constitution of this body, and these values can only have the universal meaning of which we have spoken, if all ponderable bodies have something in common.”

  43. T. H. Boyer, Phys. Rev. 182, 1374–1383 (1969); T. W. Marshall, Phys. Rev. D 24, 1509–1515 (1981).

  44. Boyer T.H. (1969). Phys. Rev. 186:1304–1318

    Article  ADS  Google Scholar 

  45. T. H. Boyer, Phys. Rev. D 27, 2906–2911 (1983); “Reply to” Phys. Rev. D 29, 2418–2419 (1984).

  46. See R. Blanco, L. Pesquera, and E. Santos Phys. Rev. D 27, 1254–1287 (1983); Phys. Rev. D 29, 2240–2254 (1984). These articles suggest that a relativistic classical scatterer again leads to the Rayleigh–Jeans spectrum. The calculations involve a general class of potentials. The mechanical momentum of the particle is calculated relativistically but the class of potentials excludes the Coulomb potential and all purely electromagnetic scattering systems. Many physicists do not realize that such mechanical systems do not satisfy Lorentz invariance. It is the Coulomb potential and only the Coulomb potential which has been incorporated into a fully relativistic classical electron theory. A general potential (unextended to involve new acceleration-dependent forces and new radiation specific to the potential) violates the center-of-energy conservation laws which are directly related to the generator of Lorentz transformations. See T. H. Boyer, Am. J. Phys. 73, 953–961 (2005).

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Timothy H. Boyer.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Boyer, T.H. Connecting Blackbody Radiation, Relativity, and Discrete Charge in Classical Electrodynamics. Found Phys 37, 999–1026 (2007). https://doi.org/10.1007/s10701-007-9139-3

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10701-007-9139-3

Keywords

Navigation