Fine-structure constant

In physics, the fine-structure constant, also known as Sommerfeld's constant, commonly denoted α (the Greek letter alpha), is a fundamental physical constant characterizing the strength of the electromagnetic interaction between elementary charged particles. It is related to the elementary charge (the electromagnetic coupling constant) e, which characterizes the strength of the coupling of an elementary charged particle with the electromagnetic field, by the formula ε0ħcα = e2. Being a dimensionless quantity, it has the same numerical value in all systems of units. Arnold Sommerfeld introduced the fine-structure constant in 1916.

Definition

Some equivalent definitions of α in terms of other fundamental physical constants are:

where:

The definition reflects the relationship between α and the electromagnetic coupling constant e, which equals αε0ħc.

In non-SI units

In electrostatic cgs units, the unit of electric charge, the statcoulomb, is defined so that the Coulomb constant, ke, or the permittivity factor, 4πε0, is 1 and dimensionless. Then the expression of the fine-structure constant, as commonly found in older physics literature, becomes

In natural units, commonly used in high energy physics, where ε0 = c = ħ = 1, the value of the fine-structure constant is[1]

As such, the fine-structure constant is just another, albeit dimensionless, quantity determining (or determined by) the elementary charge: e = α ≈ 0.30282212 in terms of such a natural unit of charge.

Measurement

The 2014 CODATA recommended value of α is[2]

0.0072973525664(17).

This has a relative standard uncertainty of 0.32 parts per billion.[2] For reasons of convenience, historically the value of the reciprocal of the fine-structure constant is often specified. The 2014 CODATA recommended value is given by[2]

137.035999139(31).

While the value of α can be estimated from the values of the constants appearing in any of its definitions, the theory of quantum electrodynamics (QED) provides a way to measure α directly using the quantum Hall effect or the anomalous magnetic moment of the electron. The theory of QED predicts a relationship between the dimensionless magnetic moment of the electron and the fine-structure constant α (the magnetic moment of the electron is also referred to as "Landé g-factor" and symbolized as g). The most precise value of α obtained experimentally (as of 2012) is based on a measurement of g using a one-electron so-called "quantum cyclotron" apparatus, together with a calculation via the theory of QED that involved 12,672 tenth-order Feynman diagrams:[3]

137.035999173(35).

This measurement of α has a precision of 0.25 parts per billion. This value and uncertainty are about the same as the latest experimental results.[4]

Physical interpretations

The fine-structure constant, α, has several physical interpretations. α is:

.
The optical conductivity of graphene for visible frequencies is theoretically given by πG0/4, and as a result its light absorption and transmission properties can be expressed in terms of the fine structure constant alone.[6] The absorption value for normal-incident light on graphene in vacuum would then be given by πα/(1 + πα/2)2 or 2.24%, and the transmission by 1/(1 + πα/2)2 or 97.75% (experimentally observed to be between 97.6% and 97.8%).

When perturbation theory is applied to quantum electrodynamics, the resulting perturbative expansions for physical results are expressed as sets of power series in α. Because α is much less than one, higher powers of α are soon unimportant, making the perturbation theory practical in this case. On the other hand, the large value of the corresponding factors in quantum chromodynamics makes calculations involving the strong nuclear force extremely difficult.

Variation with energy scale

According to the theory of the renormalization group, the value of the fine-structure constant (the strength of the electromagnetic interaction) grows logarithmically as the energy scale is increased. The observed value of α is associated with the energy scale of the electron mass; the electron is a lower bound for this energy scale because it (and the positron) is the lightest charged object whose quantum loops can contribute to the running. Therefore, 1/137.036 is the value of the fine-structure constant at zero energy. Moreover, as the energy scale increases, the strength of the electromagnetic interaction approaches that of the other two fundamental interactions, a fact important for grand unification theories. If quantum electrodynamics were an exact theory, the fine-structure constant would actually diverge at an energy known as the Landau pole. This fact makes quantum electrodynamics inconsistent beyond the perturbative expansions.

History

Arnold Sommerfeld introduced the fine-structure constant in 1916, as part of his theory of the relativistic deviations of atomic spectral lines from the predictions of the Bohr model. The first physical interpretation of the fine-structure constant α was as the ratio of the velocity of the electron in the first circular orbit of the relativistic Bohr atom to the speed of light in the vacuum.[10] Equivalently, it was the quotient between the minimum angular momentum allowed by relativity for a closed orbit, and the minimum angular momentum allowed for it by quantum mechanics. It appears naturally in Sommerfeld's analysis, and determines the size of the splitting or fine-structure of the hydrogenic spectral lines.

Is the fine-structure constant actually constant?

Physicists have pondered whether the fine-structure constant is in fact constant, or whether its value differs by location and over time. A varying α has been proposed as a way of solving problems in cosmology and astrophysics.[11][12][13][14] String theory and other proposals for going beyond the Standard Model of particle physics have led to theoretical interest in whether the accepted physical constants (not just α) actually vary.

Past rate of change

The first experimenters to test whether the fine-structure constant might actually vary examined the spectral lines of distant astronomical objects and the products of radioactive decay in the Oklo natural nuclear fission reactor. Their findings were consistent with no variation in the fine-structure constant between these two vastly separated locations and times.[15][16][17][18][19][20]

More recently, improved technology has made it possible to probe the value of α at much larger distances and to a much greater accuracy. In 1999, a team led by John K. Webb of the University of New South Wales claimed the first detection of a variation in α.[21][22][23][24] Using the Keck telescopes and a data set of 128 quasars at redshifts 0.5 < z < 3, Webb et al. found that their spectra were consistent with a slight increase in α over the last 10–12 billion years. Specifically, they found that

In 2004, a smaller study of 23 absorption systems by Chand et al., using the Very Large Telescope, found no measureable variation:[25][26]

However, in 2007 simple flaws were identified in the analysis method of Chand et al., discrediting those results.[27][28]

King et al. have used Markov Chain Monte Carlo methods to investigate the algorithm used by the UNSW group to determine from the quasar spectra, and have found that the algorithm appears to produce correct uncertainties and maximum likelihood estimates for for particular models.[29] This suggests that the statistical uncertainties and best estimate for stated by Webb et al. and Murphy et al. are robust.

Lamoreaux and Torgerson analyzed data from the Oklo natural nuclear fission reactor in 2004, and concluded that α has changed in the past 2 billion years by 45 parts per billion. They claimed that this finding was "probably accurate to within 20%." Accuracy is dependent on estimates of impurities and temperature in the natural reactor. These conclusions have to be verified.[30][31][32][33]

In 2007, Khatri and Wandelt of the University of Illinois at Urbana-Champaign realized that the 21 cm hyperfine transition in neutral hydrogen of the early Universe leaves a unique absorption line imprint in the cosmic microwave background radiation.[34] They proposed using this effect to measure the value of α during the epoch before the formation of the first stars. In principle, this technique provides enough information to measure a variation of 1 part in 109 (4 orders of magnitude better than the current quasar constraints). However, the constraint which can be placed on α is strongly dependent upon effective integration time, going as t−1/2. The European LOFAR radio telescope would only be able to constrain Δα/α to about 0.3%.[34] The collecting area required to constrain Δα/α to the current level of quasar constraints is on the order of 100 square kilometers, which is economically impracticable at the present time.

Present rate of change

In 2008, Rosenband et al.[35] used the frequency ratio of
Al+
and
Hg+
in single-ion optical atomic clocks to place a very stringent constraint on the present time variation of α, namely Δα̇/α=(−1.6±2.3)×10−17 per year. Note that any present day null constraint on the time variation of alpha does not necessarily rule out time variation in the past. Indeed, some theories[36] that predict a variable fine-structure constant also predict that the value of the fine-structure constant should become practically fixed in its value once the universe enters its current dark energy-dominated epoch.

Spatial variation – Australian dipole

In September 2010 researchers from Australia said they had identified a dipole-like structure in the variation of the fine-structure constant across the observable universe. They used data on quasars obtained by the Very Large Telescope, combined with the previous data obtained by Webb at the Keck telescopes. The fine-structure constant appears to have been larger by one part in 100,000 in the direction of the southern hemisphere constellation Ara, 10 billion years ago. Similarly, the constant appeared to have been smaller by a similar fraction in the northern direction, 10 billion years ago.[37][38][39]

In September and October 2010, after Webb's released research, physicists Chad Orzel and Sean M. Carroll suggested various approaches of how Webb's observations may be wrong. Orzel argues[40] that the study may contain wrong data due to subtle differences in the two telescopes, in which one of the telescopes the data set was slightly high and on the other slightly low, so that they cancel each other out when they overlapped. He finds it suspicious that the sources showing the greatest changes are all observed by one telescope, with the region observed by both telescopes aligning so well with the sources where no effect is observed. Carroll suggested[41] a totally different approach; he looks at the fine-structure constant as a scalar field and claims that if the telescopes are correct and the fine-structure constant varies smoothly over the universe, then the scalar field must have a very small mass. However, previous research has shown that the mass is not likely to be extremely small. Both of these scientists' early criticisms point to the fact that different techniques are needed to confirm or contradict the results, as Webb, et al., also concluded in their study.

In October 2011, Webb et al. reported[42] a variation in α dependent on both redshift and spatial direction. They report "the combined data set fits a spatial dipole" with an increase in α with redshift in one direction and a decrease in the other. "[I]ndependent VLT and Keck samples give consistent dipole directions and amplitudes...."

Anthropic explanation

The anthropic principle is a controversial argument of why the fine-structure constant has the value it does: stable matter, and therefore life and intelligent beings, could not exist if its value were much different. For instance, were α to change by 4%, stellar fusion would not produce carbon, so that carbon-based life would be impossible. If α were > 0.1, stellar fusion would be impossible and no place in the universe would be warm enough for life as we know it.[43]

Numerological explanations

As a dimensionless constant which does not seem to be directly related to any mathematical constant, the fine-structure constant has long fascinated physicists.

Arthur Eddington argued that the value could be "obtained by pure deduction" and he related it to the Eddington number, his estimate of the number of protons in the Universe.[44] This led him in 1929 to conjecture that its reciprocal was precisely the integer 137. Other physicists neither adopted this conjecture nor accepted his arguments but by the 1940s experimental values for 1/α deviated sufficiently from 137 to refute Eddington's argument.[45]

The fine-structure constant so intrigued physicist Wolfgang Pauli that he collaborated with psychoanalyst Carl Jung in a quest to understand its significance.[46] Similarly, Max Born believed if the value of alpha were any different, the universe would be degenerate, and thus that 1/137 was a law of nature.[47]

Richard Feynman, one of the originators and early developers of the theory of quantum electrodynamics (QED), referred to the fine-structure constant in these terms:

There is a most profound and beautiful question associated with the observed coupling constant, e – the amplitude for a real electron to emit or absorb a real photon. It is a simple number that has been experimentally determined to be close to 0.08542455. (My physicist friends won't recognize this number, because they like to remember it as the inverse of its square: about 137.03597 with about an uncertainty of about 2 in the last decimal place. It has been a mystery ever since it was discovered more than fifty years ago, and all good theoretical physicists put this number up on their wall and worry about it.) Immediately you would like to know where this number for a coupling comes from: is it related to pi or perhaps to the base of natural logarithms? Nobody knows. It's one of the greatest damn mysteries of physics: a magic number that comes to us with no understanding by man. You might say the "hand of God" wrote that number, and "we don't know how He pushed his pencil." We know what kind of a dance to do experimentally to measure this number very accurately, but we don't know what kind of dance to do on the computer to make this number come out, without putting it in secretly!
Richard Feynman, Richard P. Feynman (1985). QED: The Strange Theory of Light and Matter. Princeton University Press. p. 129. ISBN 0-691-08388-6. 

Conversely, statistician I. J. Good argued that a numerological explanation would only be acceptable if it came from a more fundamental theory that also provided a Platonic explanation of the value.[48]

Attempts to find a mathematical basis for this dimensionless constant have continued up to the present time. However, no numerological explanation has ever been accepted by the community.

Quotes

The mystery about α is actually a double mystery. The first mystery – the origin of its numerical value α ≈ 1/137 has been recognized and discussed for decades. The second mystery – the range of its domain – is generally unrecognized.
Malcolm H. Mac Gregor, M.H. MacGregor (2007). The Power of Alpha. World Scientific. p. 69. ISBN 978-981-256-961-5. 

See also

References

  1. Peskin, M.; Schroeder, D. (1995). An Introduction to Quantum Field Theory. Westview Press. p. 125. ISBN 0-201-50397-2.
  2. 1 2 3 Mohr, P. J.; Taylor, B. N.; Newell, D. B. (2015). "Fine structure constant". CODATA Internationally recommended 2014 values of the fundamental physical constants. National Institute of Standards and Technology.
  3. Aoyama, T.; Hayakawa, M.; Kinoshita, T.; Nio, M. (2012). "Tenth-order QED contribution to the electron g - 2 and an improved value of the fine structure constant". Physical Review Letters. 109 (11): 111807. arXiv:1205.5368Freely accessible. Bibcode:2012PhRvL.109k1807A. doi:10.1103/PhysRevLett.109.111807.
  4. Bouchendira, Rym; Cladé, Pierre; Guellati-Khélifa, Saïda; Nez, François; Biraben, François (2010). "New determination of the fine-structure constant and test of the quantum electrodynamics". Physical Review Letters. 106 (8): 080801. arXiv:1012.3627Freely accessible. Bibcode:2011PhRvL.106h0801B. doi:10.1103/PhysRevLett.106.080801. PMID 21405559.
  5. NIST Reference on Constants, Units, and Uncertainty Current advances: The fine-structure constant and quantum Hall effect
  6. Nair, R. R.; Blake, P.; Grigorenko, A. N.; Novoselov, K. S.; Booth, T. J.; Stauber, T.; Peres, N. M. R.; Geim, A. K. (2008). "Fine Structure Constant Defines Visual Transparency of Graphene". Science. 320 (5881): 1308–1308. Bibcode:2008Sci...320.1308N. doi:10.1126/science.1156965.
  7. http://www.nobelprize.org/nobel_prizes/physics/laureates/1983/chandrasekhar-lecture.pdf
  8. Bedford, D.; Krumm, P. (2004). "Heisenberg indeterminacy and the fine structure constant". American Journal of Physics. 72 (7): 969. doi:10.1119/1.1646135.
  9. Poelz, G. (5 Oct 2016). "An Electron Model with Synchrotron Radiation". arXiv:1206.0620v24Freely accessible [physics.class-ph].
  10. "Introduction to the Constants for Nonexperts – Current Advances: The Fine-Structure Constant and Quantum Hall Effect". The NIST Reference on Constants, Units, and Uncertainty. NIST. Retrieved 11 April 2009.
  11. Milne, E. A. (1935). Relativity, Gravitation and World Structure. Clarendon Press.
  12. Dirac, P.A.M. (1937). "The Cosmological Constants". Nature. 139 (3512): 323. Bibcode:1937Natur.139..323D. doi:10.1038/139323a0.
  13. Gamow, G. (1967). "Electricity, Gravity, and Cosmology". Physical Review Letters. 19 (13): 759. Bibcode:1967PhRvL..19..759G. doi:10.1103/PhysRevLett.19.759.
  14. Gamow, G. (1967). "Variability of Elementary Charge and Quasistellar Objects". Physical Review Letters. 19 (16): 913. Bibcode:1967PhRvL..19..913G. doi:10.1103/PhysRevLett.19.913.
  15. Uzan, J.-P. (2003). "The Fundamental Constants and Their Variation: Observational Status and Theoretical Motivations". Reviews of Modern Physics. 75 (2): 403–455. arXiv:hep-ph/0205340Freely accessible. Bibcode:2003RvMP...75..403U. doi:10.1103/RevModPhys.75.403.
  16. Uzan, J.-P. (2004). "Variation of the Constants in the Late and Early Universe". arXiv:astro-ph/0409424Freely accessible.
  17. Olive, K.; Qian, Y.-Z. (2003). "Were Fundamental Constants Different in the Past?". Physics Today. 57 (10): 40–45. Bibcode:2004PhT....57j..40O. doi:10.1063/1.1825267.
  18. Barrow, J.D. (2002). The Constants of Nature: From Alpha to Omega—the Numbers That Encode the Deepest Secrets of the Universe. Vintage. ISBN 0-09-928647-5.
  19. Uzan, J.-P.; Leclercq, B. (2008). The Natural Laws of the Universe: Understanding Fundamental Constants. Springer Praxis. ISBN 978-0-387-73454-5.
  20. Yasunori, F. (2004). "Oklo Constraint on the Time-Variability of the Fine-Structure Constant". Astrophysics, Clocks and Fundamental Constants. Lecture Notes in Physics. Springer Berlin. pp. 167–185. ISBN 978-3-540-21967-5.
  21. Webb, J. K.; et al. (1999). "Search for Time Variation of the Fine Structure Constant". Physical Review Letters. 82 (5): 884–887. arXiv:astro-ph/9803165Freely accessible. Bibcode:1999PhRvL..82..884W. doi:10.1103/PhysRevLett.82.884.
  22. Murphy, M. T.; et al. (2001). "Possible evidence for a variable fine-structure constant from QSO absorption lines: motivations, analysis and results". Monthly Notices of the Royal Astronomical Society. 327 (4): 1208. arXiv:astro-ph/0012419Freely accessible. Bibcode:2001MNRAS.327.1208M. doi:10.1046/j.1365-8711.2001.04840.x.
  23. Murphy, M. T.; et al. (2001). "Further Evidence for Cosmological Evolution of the Fine Structure Constant". Physical Review Letters. 87 (9): 091301. arXiv:astro-ph/0012539Freely accessible. Bibcode:2001PhRvL..87i1301W. doi:10.1103/PhysRevLett.87.091301. PMID 11531558.
  24. Murphy, M. T.; Webb, J .K.; Flambaum, V. V. (2003). "Further Evidence for a Variable Fine-Structure Constant from Keck/HIRES QSO Absorption Spectra". Monthly Notices of the Royal Astronomical Society. 345 (2): 609. arXiv:astro-ph/0306483Freely accessible. Bibcode:2003MNRAS.345..609M. doi:10.1046/j.1365-8711.2003.06970.x.
  25. Chand, H.; et al. (2004). "Probing the Cosmological Variation of the Fine-Structure Constant: Results Based on VLT-UVES Sample". Astronomy & Astrophysics. 417 (3): 853. arXiv:astro-ph/0401094Freely accessible. Bibcode:2004A&A...417..853C. doi:10.1051/0004-6361:20035701.
  26. R. Srianand; et al. (2004). "Limits on the Time Variation of the Electromagnetic Fine-Structure Constant in the Low Energy Limit from Absorption Lines in the Spectra of Distant Quasars". Physical Review Letters. 92 (12): 121302. arXiv:astro-ph/0402177Freely accessible. Bibcode:2004PhRvL..92l1302S. doi:10.1103/PhysRevLett.92.121302. PMID 15089663.
  27. M.T. Murphy; J.K. Webb; V.V. Flambaum (2007). "Comment on "Limits on the Time Variation of the Electromagnetic Fine-Structure Constant in the Low Energy Limit from Absorption Lines in the Spectra of Distant Quasars"". Physical Review Letters. 99 (23): 239001. arXiv:0708.3677Freely accessible. Bibcode:2007PhRvL..99w9001M. doi:10.1103/PhysRevLett.99.239001.
  28. M.T. Murphy; J.K. Webb; V.V. Flambaum (2008). "Revision of VLT/UVES Constraints on a Varying Fine-Structure Constant". Monthly Notices of the Royal Astronomical Society. 384 (3): 1053. arXiv:astro-ph/0612407Freely accessible. Bibcode:2008MNRAS.384.1053M. doi:10.1111/j.1365-2966.2007.12695.x.
  29. J. King; D. Mortlock; J. Webb; M. Murphy (2009). "Markov Chain Monte Carlo methods applied to measuring the fine-structure constant from quasar spectroscopy". arXiv:0910.2699Freely accessible [astro-ph].
  30. R. Kurzweil (2005). The Singularity Is Near. Viking Penguin. pp. 139–140. ISBN 0-670-03384-7.
  31. S.K. Lamoreaux; J.R. Torgerson (2004). "Neutron Moderation in the Oklo Natural Reactor and the Time Variation of Alpha". Physical Review D. 69 (12): 121701. arXiv:nucl-th/0309048Freely accessible. Bibcode:2004PhRvD..69l1701L. doi:10.1103/PhysRevD.69.121701.
  32. E.S. Reich (30 June 2004). "Speed of Light May Have Changed Recently". New Scientist. Retrieved 30 January 2009.
  33. "Scientists Discover One Of The Constants Of The Universe Might Not Be Constant". ScienceDaily. 12 May 2005. Retrieved 30 January 2009.
  34. 1 2 R. Khatri; B.D. Wandelt (2007). "21-cm Radiation: A New Probe of Variation in the Fine-Structure Constant". Physical Review Letters. 98 (11): 111301. arXiv:astro-ph/0701752Freely accessible. Bibcode:2007PhRvL..98k1301K. doi:10.1103/PhysRevLett.98.111301. PMID 17501040.
  35. T. Rosenband; et al. (2008). "Frequency Ratio of Al+ and Hg+ Single-Ion Optical Clocks; Metrology at the 17th Decimal Place". Science. 319 (5871): 1808–12. Bibcode:2008Sci...319.1808R. doi:10.1126/science.1154622. PMID 18323415.
  36. J.D. Barrow; H.B. Sandvik; J. Magueijo (2001). "The Behaviour of Varying-Alpha Cosmologies". Phys. Rev. D. 65 (6): 063504. arXiv:astro-ph/0109414Freely accessible. Bibcode:2002PhRvD..65f3504B. doi:10.1103/PhysRevD.65.063504.
  37. H. Johnston (2 September 2010). "Changes spotted in fundamental constant". Physics World. Retrieved 11 September 2010.
  38. J.K.Webb; King, J. A.; Murphy, M. T.; Flambaum, V. V.; Carswell, R. F.; Bainbridge, M. B. (23 August 2010). "Evidence for spatial variation of the fine-structure constant". Physical Review Letters. 107 (19): 191101. arXiv:1008.3907Freely accessible. Bibcode:2011PhRvL.107s1101W. doi:10.1103/PhysRevLett.107.191101. PMID 22181590.
  39. J. A. King (2010). Searching for variations in the fine-structure constant and the proton-to-electron mass ratio using quasar absorption lines (PhD thesis). University of New South Wales.
  40. Orzel, Chad (14 October 2010). "Why I'm Skeptical About the Changing Fine-Structure Constant". ScienceBlogs.
  41. Carroll, Sean M. (18 October 2010). "The Fine Structure Constant is Probably Constant".
  42. J. K. Webb; J. A. King; M. T. Murphy; V. V. Flambaum; R. F. Carswell; M. B. Bainbridge (4 November 2011). "Indications of a Spatial Variation of the Fine Structure Constant" (PDF). Physical Review Letters. 107 (19): 191101. arXiv:1008.3907Freely accessible. Bibcode:2011PhRvL.107s1101W. doi:10.1103/PhysRevLett.107.191101.
  43. J.D. Barrow (2001). "Cosmology, Life, and the Anthropic Principle". Annals of the New York Academy of Sciences. 950 (1): 139–153. Bibcode:2001NYASA.950..139B. doi:10.1111/j.1749-6632.2001.tb02133.x.
  44. A.S Eddington (1956). "The Constants of Nature". In J.R. Newman. The World of Mathematics. 2. Simon & Schuster. pp. 1074–1093.
  45. H. Kragh (2003). "Magic Number: A Partial History of the Fine-Structure Constant". Archive for History of Exact Sciences. 57 (5): 395. doi:10.1007/s00407-002-0065-7 (inactive 2016-10-06).
  46. P. Varlaki; L. Nadai; J. Bokor (2008). "Number Archetypes and Background Control Theory Concerning the Fine Structure Constant" (PDF). Acta Polytechnica Hungarica. 5 (2): 71.
  47. A.I. Miller (2009). Deciphering the Cosmic Number: The Strange Friendship of Wolfgang Pauli and Carl Jung. W.W. Norton & Co. p. 253. ISBN 978-0-393-06532-9. Max Born: If alpha were bigger than it really is, we should not be able to distinguish matter from ether [the vacuum, nothingness], and our task to disentangle the natural laws would be hopelessly difficult. The fact however that alpha has just its value 1/137 is certainly no chance but itself a law of nature. It is clear that the explanation of this number must be the central problem of natural philosophy.
  48. I. J. Good (1990). "A Quantal Hypothesis for Hadrons and the Judging of Physical Numerology." in G. R. Grimmett (Editor), D. J. A. Welsh (Editor). Disorder in Physical Systems. Oxford University Press. p. 141. ISBN 978-0-19-853215-6. I. J. Good: There have been a few examples of numerology that have led to theories that transformed society: see the mention of Kirchhoff and Balmer in Good (1962, p. 316) … and one can well include Kepler on account of his third law. It would be fair enough to say that numerology was the origin of the theories of electromagnetism, quantum mechanics, gravitation.... So I intend no disparagement when I describe a formula as numerological. When a numerological formula is proposed, then we may ask whether it is correct. … I think an appropriate definition of correctness is that the formula has a good explanation, in a Platonic sense, that is, the explanation could be based on a good theory that is not yet known but ‘exists’ in the universe of possible reasonable ideas.

External links

This article is issued from Wikipedia - version of the 12/4/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.