the contrasting roles of planck's constant in …the experimental blackbody data when using...

5
The contrasting roles of Planck's constant in classical and quantum theories Timothy H. Boyer Citation: American Journal of Physics 86, 280 (2018); doi: 10.1119/1.5021355 View online: https://doi.org/10.1119/1.5021355 View Table of Contents: http://aapt.scitation.org/toc/ajp/86/4 Published by the American Association of Physics Teachers Articles you may be interested in Fresnel's original interpretation of complex numbers in 19th century optics American Journal of Physics 86, 245 (2018); 10.1119/1.5011366 The Toy model: Understanding the early universe American Journal of Physics 86, 290 (2018); 10.1119/1.5026694 On the optical path length in refracting media American Journal of Physics 86, 268 (2018); 10.1119/1.5013008 Deriving mass-energy equivalence and mass-velocity relation without light American Journal of Physics 86, 284 (2018); 10.1119/1.5021794 User's guide to Monte Carlo methods for evaluating path integrals American Journal of Physics 86, 293 (2018); 10.1119/1.5024926 Sinking bubbles in stout beers American Journal of Physics 86, 250 (2018); 10.1119/1.5021361

Upload: others

Post on 14-Mar-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The contrasting roles of Planck's constant in …the experimental blackbody data when using theoretical ideas associated with Wien’s proposed form for the radiation spec-trum. Planck’s

The contrasting roles of Planck's constant in classical and quantum theoriesTimothy H. Boyer

Citation: American Journal of Physics 86, 280 (2018); doi: 10.1119/1.5021355View online: https://doi.org/10.1119/1.5021355View Table of Contents: http://aapt.scitation.org/toc/ajp/86/4Published by the American Association of Physics Teachers

Articles you may be interested inFresnel's original interpretation of complex numbers in 19th century opticsAmerican Journal of Physics 86, 245 (2018); 10.1119/1.5011366

The Toy model: Understanding the early universeAmerican Journal of Physics 86, 290 (2018); 10.1119/1.5026694

On the optical path length in refracting mediaAmerican Journal of Physics 86, 268 (2018); 10.1119/1.5013008

Deriving mass-energy equivalence and mass-velocity relation without lightAmerican Journal of Physics 86, 284 (2018); 10.1119/1.5021794

User's guide to Monte Carlo methods for evaluating path integralsAmerican Journal of Physics 86, 293 (2018); 10.1119/1.5024926

Sinking bubbles in stout beersAmerican Journal of Physics 86, 250 (2018); 10.1119/1.5021361

Page 2: The contrasting roles of Planck's constant in …the experimental blackbody data when using theoretical ideas associated with Wien’s proposed form for the radiation spec-trum. Planck’s

The contrasting roles of Planck’s constant in classical and quantumtheories

Timothy H. BoyerDepartment of Physics, City College of the City University of New York, New York, New York 10031

(Received 1 October 2017; accepted 31 December 2017)

We trace the historical appearance of Planck’s constant in physics, and we note that initially the

constant did not appear in connection with quanta. Furthermore, we emphasize that Planck’s

constant can appear in both classical and quantum theories. In both theories, Planck’s constant sets

the scale of atomic phenomena. However, the roles played in the foundations of the theories are

sharply different. In quantum theory, Planck’s constant is crucial to the structure of the theory. On

the other hand, in classical electrodynamics, Planck’s constant is optional, since it appears only as

the scale factor for the (homogeneous) source-free contribution to the general solution of

Maxwell’s equations. Since classical electrodynamics can be solved while taking the homogenous

source-free contribution in the solution as zero or non-zero, there are naturally two different

theories of classical electrodynamics, one in which Planck’s constant is taken as zero and one

where it is taken as non-zero. The textbooks of classical electromagnetism present only the version

in which Planck’s constant is taken to vanish. VC 2018 American Association of Physics Teachers.

https://doi.org/10.1119/1.5021355

I. INTRODUCTION

“Planck’s constant h is a ‘quantum constant”’ is what I amtold by my students. “Planck’s constant is not allowed in aclassical theory” is the view of many physicists. I believethat these views misunderstand the role of physical constantsand, in particular, the role of Planck’s constant h within clas-sical and quantum theories. It is noteworthy that despite thecurrent orthodox view which restricts the constant h to quan-tum theory, Planck searched for many years for a way to fithis constant h into classical electrodynamics, and felt hecould find no place. Late in life, Planck1 acknowledged thathis colleagues felt that his long futile search “bordered on atragedy.” However, Planck’s constant h indeed has a naturalplace in classical electrodynamics. Yet even today, mostphysicists are unaware of this natural place, and some evenwish to suppress information regarding this role. In this arti-cle, we review the historical2 appearance of Planck’s con-stant, and then emphasize its contrasting roles in classicaland quantum theories.

II. PHYSICAL CONSTANTS AS SCALES FOR

PHYSICAL PHENOMENA

Physical constants appear in connection with measure-ments of the scale of natural phenomena. Thus, for example,Cavendish’s constant G appeared first in connection withgravitational forces using the Newtonian theory of gravity.At the end of the 18th century, Cavendish’s experiment usinga torsion balance in combination with Newton’s theory pro-vided the information needed for an accurate evaluation ofthe constant. However, because physical constants refer tonatural phenomena and are not the exclusive domain of anyone theory, Cavendish’s constant G also reappears inEinstein’s 20th-century general-relativistic description ofgravitational phenomena.

In a similar fashion, Planck’s constant h sets the scale ofelectromagnetic phenomena at the atomic level. It was firstevaluated in 1899 in connection with a fit of experimentaldata to Wien’s theoretical suggestion for the spectrum of

blackbody radiation. Subsequently, the constant has reap-peared both in quantum theory and in classical electrody-namics. In its role as a scale, the constant h can appear inany theory which attempts to explain aspects of atomicphysics.

III. PHYSICAL CONSTANTS OF THE 19TH

CENTURY

The 19th century saw developments in theories involving anumber of physical constants which are still in use today. Thekinetic theory of gases involves the gas constant R andAvogadro’s number NA. The unification of electricity andmagnetism in Maxwell’s equations involves an implicit orexplicit (depending on the units) appearance of the speed oflight in vacuum c. Measurements of blackbody radiation ledto the introduction of new physical constants. Thus theappearance of the Stefan-Boltzmann law UT ¼ aST4V for thethermal energy UT of radiation in a cavity of volume V at tem-perature T introduced a new constant aS; Stefan’s constant, in1879. Today this constant is reexpressed in terms of later con-stants as aS ¼ p2k4

B=ð15�h3c3Þ: In the 1890s, careful experi-mental work on the spectrum of blackbody radiation led toWien’s theoretical suggestion for the blackbody radiationspectrum with its two constants, (labeled here as a and bÞ;

qWð�; TÞ ¼ a�3 exp �b�=T½ �: (1)

Also at the end of the century, the measurement of the ratioof charge to mass for cathode rays led to new constantsinvolving the charge and mass of the electron.

IV. APPEARANCE OF PLANCK’S CONSTANT

Planck’s great interest in thermodynamics led him to con-sider the equilibrium of electric dipole oscillators whenlocated in random classical radiation. Planck found that therandom radiation spectrum could be connected to the aver-age energy Uð�; TÞ of an electric dipole oscillator of naturalfrequency � as qð�;TÞ ¼ ð8p�2=c3ÞUð�; TÞ. Introducing

280 Am. J. Phys. 86 (4), April 2018 http://aapt.org/ajp VC 2018 American Association of Physics Teachers 280

Page 3: The contrasting roles of Planck's constant in …the experimental blackbody data when using theoretical ideas associated with Wien’s proposed form for the radiation spec-trum. Planck’s

Wien’s suggestion qW in Eq. (1) for the spectral form,Planck found for the average energy of an oscillator3

UWð�;TÞ ¼ h� exp �b�=T½ �: (2)

Here is where Planck’s constant h first appeared in physics. Atthe meeting of the Prussian Academy of Sciences on May 18,1899, Planck reported4 the value b as b ¼ 0:4818� 10�10 s �K and the value of the constant h as h ¼ 6:885� 10�27 erg-s.Thus initially, Planck’s constant appeared as a numerical fit tothe experimental blackbody data when using theoretical ideasassociated with Wien’s proposed form for the radiation spec-trum. Planck’s constant h had nothing to do with quanta in itsfirst appearance in physics.

In the middle of the year 1900, experimentalists Rubensand Kurlbaum5 found that their measurements of the black-body radiation spectrum departed from the form suggestedby Wien in Eq. (1). It was found that at low frequencies(long wavelengths) the spectrum qð�; TÞ seem to be propor-tional to �2T: Planck learned of the new experimental results,and, using ideas of energy and entropy for a dipole oscillator,introduced a simple interpolation between the newly sug-gested low-frequency form and the well-established Wienhigh-frequency form. His interpolation gave him the Planckradiation form corresponding to a dipole oscillator energy6

UPð�; TÞ ¼h�

exp b�=T½ � � 1: (3)

Planck reported7 this suggested blackbody radiation spec-trum to the German Physical Society on October 19, 1900.The Planck spectrum (3) involved exactly the same constantsas appeared in his work when starting from Wien’s spectrum.The experimentalists confirmed that Planck’s new suggestedspectrum was an excellent fit to the data. Once againPlanck’s constant h appeared as a parameter in a fit to experi-mental data when working with an assumed theoretical form.There was still no suggestion of quanta in connection withPlanck’s constant h.

Planck still needed a theoretical justification for his newspectral form. Although Planck had hoped initially that hisdipole oscillators would act as black particles and wouldbring random radiation into the equilibrium blackbody spec-trum, he had come to realize that his small linear oscillatorswould not change the frequency spectrum of the randomradiation; the incident and the scattered radiation were at thesame frequency.8 In late 1900 in an “act of desperation,”Planck turned to Boltzmann’s statistical work which he hadearlier “vehemently rejected.”9 It was in connection with theuse of statistical ideas that Planck found that the constant bin the radiation spectra could be rewritten as b ¼ h=kB wherekB ¼ R=NA; where R and NA were the constants which hadalready appeared in the kinetic theory of gases.10 Indeed, itwas Planck who introduced Boltzmann’s constant kB intophysics. Boltzmann had always stated that entropy S wasproportional to the logarithm of probability W without evergiving the constant of proportionality. Planck introduced theequality S ¼ kBlnW: Also, in his calculations of the probabil-ity, Planck departed from Boltzmann’s procedures in retain-ing the energy connection E ¼ h� without taking theexpected limit E ! 0: Only by avoiding the limit, couldPlanck recover his radiation spectrum. It was here, in retain-ing E ¼ h� rather than taking Boltzmann’s limit, that theassociation of Planck’s constant h with quanta first appeared.

V. PLANCK’S CONSTANT IN CURRENT QUANTUM

THEORY

Although Planck’s constant did not originally appear inconnection with quantum theory, the constant became a use-ful scale factor when dealing with the photoelectric effect,specific heats of solids, and the Bohr atom. The theoreticalcontext for all these phenomena was quantum theory.Quantum theory developed steadily during the first third ofthe 20th century. Today quantum theory is regarded as thetheory which gives valid results for phenomena at the atomiclevel.

Our current textbooks emphasize that quantum theoryincorporates Planck’s constant h into the essential aspects ofthe theory involving non-commuting operators. Thus posi-tion and momentum are associated with operators satisfyingthe commutation relation ½x̂; p̂x� ¼ ih=ð2pÞ ¼ i�h: The non-commutativity is associated with a non-zero value ofPlanck’s constant h and disappears when h is taken to zero.Associated with the non-commuting operators is the zero-point energy U ¼ ð1=2Þh�0 of a harmonic oscillator of natu-ral frequency �0: The oscillator zero-point energy vanishesalong with a vanishing value for h.

VI. PLANCK’S CONSTANT IN CLASSICAL

ELECTRODYNAMICS

The natural place for Planck’s constant within classicalelectrodynamics is as the scale factor of the source-free contri-bution to the general solution of Maxwell’s equations.Maxwell’s equations for the electromagnetic fields Eðr; tÞ andBðr; tÞ can be rewritten11 in terms of the potentials Uðr; tÞ andAðr; tÞ where E ¼ �rU� ð1=cÞ@A=@t and B ¼ r� A: Inthe Lorenz gauge, Maxwell’s equations for the potentialsbecome wave equations with sources in the charge densityqðr; tÞ and current density Jðr; tÞ

r2 � 1

c2

@2

@t2

� �Uðr; tÞ ¼ �4pqðr; tÞ; (4)

r2 � 1

c2

@2

@t2

� �Aðr; tÞ ¼ �4p

Jðr; tÞc

: (5)

The general solutions of these differential equations in allspacetime with outgoing wave boundary conditions are

Uðr; tÞ ¼ Uinðr; tÞ

þð

d3r0ð

dt0d t� t0 � jr� r0j=cð Þ

jr� r0j q r0; t0ð Þ; (6)

Aðr; tÞ ¼ Ainðr; tÞ

þð

d3r0ð

dt0d t� t0 � jr� r0j=cð Þ

jr� r0jJ r0; t0ð Þ

c; (7)

where Uinðr; tÞ and Ainðr; tÞ are the (homogeneous) source-free contributions to the general solutions of Eqs. (4) and (5).It is universally accepted that these are the accurate solu-tions. However, all the textbooks12 of electromagnetism omitthe source-free contributions Uinðr; tÞ and Ainðr; tÞ; and pre-sent only the contributions due to the charge and currentssources qðr; tÞ and Jðr; tÞ:

Within a laboratory situation, the experimenter’s sourcescorrespond to qðr; tÞ and Jðr; tÞ while the source-free terms

281 Am. J. Phys., Vol. 86, No. 4, April 2018 Timothy H. Boyer 281

Page 4: The contrasting roles of Planck's constant in …the experimental blackbody data when using theoretical ideas associated with Wien’s proposed form for the radiation spec-trum. Planck’s

Uinðr; tÞ and Ainðr; tÞ correspond to contributions the experi-menter does not control which are present when the experi-menter enters his laboratory. The terms Uinðr; tÞ and Ainðr; tÞmight correspond to radio waves coming from a nearbybroadcasting station, or might correspond to thermal radia-tion from the walls of the laboratory. In a shielded laboratoryheld at zero temperature, there is still random classical radia-tion present which can be measured using the Casimirforce13 between two parallel conducting plates. Interpretedwithin classical electromagnetic theory, this Casimir forceresponds to all the classical radiation surrounding the con-ducting parallel plates, and it is found experimentally14 thatthis force does not go to zero as the temperature goes tozero. The Casimir force, interpreted within classical electro-magnetic theory, indicates that there is classical electromag-netic zero-point radiation present throughout spacetime. Thezero-point spectrum is Lorentz invariant with a scale set byPlanck’s constant h. Thus in a shielded laboratory at zero-temperature, the scalar potential Uin can be taken to vanishand the vector potential Aðr; tÞ should be written as

Aðr; tÞ ¼X2

k¼1

ðd3k�̂ k; kð Þ h

4p3x

� �1=2

� sin k � r� xtþ h k; kð Þ½ �

þð

d3r0ð

dt0d t� t0 � jr� r0j=cð Þ

jr� r0jJ r0; t0ð Þ

c;

including a source-free contribution which is a Lorentz-invariant spectrum of plane waves with random phaseshðk; kÞ and with Planck’s constant h setting the scale.15

As indicated here, there is a natural place for Planck’sconstant h within classical electromagnetic theory. Once thiszero-point radiation is present in the theory, it causes zero-point energy for dipole oscillators, and it can be used toexplain Casimir forces, van der Waals forces, oscillator spe-cific heats, diamagnetism, the blackbody radiation spectrum,and the absence of “atomic collapse.”16

VII. CONTRASTING ROLES FOR PLANCK’S

CONSTANT

Planck’s constant h sets the scale for atomic phenomena.However, the constant plays contrasting roles in classicaland quantum theories. Within quantum theory, Planck’s con-stant h is embedded in the essential aspects of the theory. Ifone sets Planck’s constant to zero, then the quantum charac-ter of the theory disappears; the non-commuting operatorsbecome simply commuting c-numbers, and the zero-pointenergy of a harmonic oscillator drops to zero.

On the other hand, within classical electrodynamics,Planck’s constant does not appear in Maxwell’s fundamentaldifferential equations. Rather, Planck’s constant appearsonly in the (homogeneous) source-free contribution to thegeneral solution of Maxwell’s equations. Thus classical elec-trodynamics can exist in two natural forms. In one formPlanck’s constant is taken as non-zero, and one can explain anumber of natural phenomena at the atomic level. In theother form, Planck’s constant is taken to vanish. It is onlythis second form which appears in the textbooks of electro-magnetism and modern physics. However, even this form isquite sufficient to account for our macroscopic electromag-netic technology.

VIII. COMMENTS ON PLANCK’S CONSTANT IN

CLASSICAL THEORY

Planck was a “reluctant revolutionary” who tried to con-serve as much of 19th century physics as possible. Late inlife Planck wrote1 in his scientific autobiography, “My futileattempts to fit the elementary quantum of action somehowinto the classical theory continued for a number of years, andthey cost me a great deal of effort. Many of my colleaguessaw in this something bordering on a tragedy.” Within hislifetime, there seems to have been no recognition of a naturalplace for Planck’s constant h within classical electrodynam-ics. It was only in the 1960s, beginning with careful work byMarshall,17 that it became clear that Planck’s constant hcould be incorporated in an natural way within classicalelectrodynamics.

Despite the realization that Planck’s constant is associatedwith atomic phenomena and that at least some of atomic phe-nomena can be described within either classical or quantumtheory, there seems great reluctance on the part of somephysicists to acknowledge the possibility of Planck’s con-stant appearing within classical theory. One referee wrotethe following justification18 in rejecting the idea of Planck’sconstant appearing within classical electromagnetism: “Butas a pedagogical matter, doesn’t it muddy the distinctionbetween classical and quantum physics? The traditionaldividing line may be in some aspects arbitrary, but at least itis clear (�h) quantum).” This referee clearly misunderstandsthe nature of physical constants and seems willing to sacri-fice scientific accuracy to pedagogical simplicity.

I believe that many physicists would prefer truth to conve-nient pedagogy. Certainly, the introduction of Planck’s con-stant as the scale factor for source-free classical zero-pointradiation expands the range of phenomena described by clas-sical electromagnetic theory.16 The recognition that Planck’sconstant has a natural place within classical theory may pro-vide a broadening perspective beyond a confining quantumorthodoxy. In any case, one suspects that Planck would havebeen pleased to find that there is indeed a natural role for hisconstant h within classical electromagnetic theory.

1M. Planck, Scientific Autobiography and Other Papers, translated by F.

Gaynor (Philosophical Library, New York, 1949), p. 35.2The historical information is taken from the following sources: T. S.

Kuhn, Black-Body Theory and the Quantum Discontinuity 1894-1912(Oxford U.P., New York, 1978); A. Hermann, The Genesis of QuantumTheory (1899-1913), translated by C. W. Nash (MIT Press, Cambridge,

MA, 1971); M. J. Klein, “Planck, entropy, and quanta 1901-1906,” TheNatural Philosopher I (Blaisdell Publishing Co., New York, 1963), pp.

81–108; M. J. Klein, “Thermodynamics and quanta in Planck’s work,” in

History of Physics, edited by S. R. Weart and M. Philips (AIP, New York,

1985), pp. 294–302.3See, for example, Ref. 2, M. Klein, History, p. 297.4M. Planck, S.-B. Preuss, Akad. Wiss. 440 (1899).5See, for example, Ref. 2, A. Hermann, p. 13.6See, for example, Ref. 2. A. Hermann, p. 14.7M. Planck, Verh. d. Deutsch. Phys. Ges. 2, 202 (1900).8See, for example, Ref. 2, T. Kuhn, pp. 34–36.9See, for example, Ref. 2, A. Hermann, p. 17.

10See, for example, Ref. 2, M. Klein, Philosopher, pp. 91–92.11We are using Gaussian units.12See, for example, D. J. Griffiths, Introduction to Electrodynamics, 4th ed.

(Pearson, New York, 2013), Sec. 10.2.1, Eq. (10.26); or L. Eyges, TheClassical Electrodynamic Field (Cambridge U.P., New York, 1972), p. 186,

Eqs. (11.45) and (11.46); or J. D. Jackson, Classical Electrodynamics, 3rd ed.

(John Wiley & Sons, New York, 1999), p. 246, Eq. (6.48); or A. Zangwill,

Modern Electrodynamics (Cambridge U.P., 2013), pp. 724–725, Eqs. (20.57)

282 Am. J. Phys., Vol. 86, No. 4, April 2018 Timothy H. Boyer 282

Page 5: The contrasting roles of Planck's constant in …the experimental blackbody data when using theoretical ideas associated with Wien’s proposed form for the radiation spec-trum. Planck’s

and (20.58); or A. Garg, Classical Electromagnetism in a Nutshell (Princeton

U.P., Princeton, NJ, 2012), p. 204, Eqs. (54.16) and (54.17).13H. B. G. Casimir, “On the attraction between two perfectly conducting

plates,” Proc. Ned. Akad. Wetenschap. 51, 793–795 (1948).14M. J. Sparnaay, “Measurement of the attractive forces between flat plates,”

Physica (Amsterdam) 24, 751–764 (1958); S. K. Lamoreaux,

“Demonstration of the Casimir force in the 0.6 to 6 lm range,” Phys. Rev.

Lett. 78, 5–8 (1997); 81, 5475–5476 (1998); U. Mohideen, “Precision

measurement of the Casimir force from 0.1 to 0.9 lm,” ibid. 81,

4549–4552 (1998); H. B. Chan, V. A. Aksyuk, R. N. Kleinman, D. J.

Bishop, and F. Capasso, “Quantum mechanical actuation of microelectro-

mechanical systems by the Casimir force,” Science 291, 1941–1944

(2001); G. Bressi, G. Caarugno, R. Onofrio, and G. Ruoso, “Measurement

of the Casimir force between parallel metallic surfaces,” Phys. Rev. Lett.

88, 041804 (2002).

15T. H. Boyer, “Random electrodynamics: The theory of classical electrody-

namics with classical electromagnetic zero-point radiation,” Phys. Rev. D

11, 790–808 (1975).16T. H. Boyer, “Any classical description of nature requires classical electro-

magnetic zero-point radiation,” Am. J. Phys. 79, 1163–1167 (2011); See

also D. C. Cole and Y. Zou, “Quantum mechanical ground state of hydro-

gen obtained from classical electrodynamics,” Phys. Lett. A 317, 14–20

(2003); A review of the work on classical electromagnetic zero-point radi-

ation up to 1996 is provided by L. de la Pena and A. M. Cetto, TheQuantum Dice—An Introduction to Stochastic Electrodynamics (Kluwer

Academic, Dordrecht 1996).17T. W. Marshall, “Random electrodynamics,” Proc. R. Soc. A 276,

475–491 (1963); “Statistical electrodynamics,” Proc. Camb. Phil. Soc. 61,

537–546 (1965).18Author’s personal correspondence with the American Journal of Physics.

283 Am. J. Phys., Vol. 86, No. 4, April 2018 Timothy H. Boyer 283