Planck constant

Source: Wikipedia, the free encyclopedia.
Planck constant
Common symbols
m2 s−1
Dimension
Value
6.62607015×10−34 J⋅Hz−1
4.135667696...×10−15 eV⋅Hz−1
Reduced Planck constant
Common symbols
m2 s
−1
Derivations from
other quantities
Dimension
Value
1.054571817...×10−34 J⋅s
6.582119569...×10−16 eV⋅s

The Planck constant, or Planck's constant, denoted by ,[1] is a fundamental physical constant[1] of foundational importance in quantum mechanics: a photon's energy is equal to its frequency multiplied by the Planck constant, and the wavelength of a matter wave equals the Planck constant divided by the associated particle momentum.

The constant was postulated by

black-body radiation.[2] Planck later referred to the constant as the "quantum of action".[3] In 1905, Albert Einstein associated the "quantum" or minimal element of the energy to the electromagnetic wave itself. Max Planck received the 1918 Nobel Prize in Physics
"in recognition of the services he rendered to the advancement of Physics by his discovery of energy quanta".

In

SI unit of mass.[4]
The SI units are defined in such a way that, when the Planck constant is expressed in SI units, it has the exact value = 6.62607015×10−34 J⋅Hz−1.[5][6]

History

Origin of the constant

Plaque at the Humboldt University of Berlin: "Max Planck, who discovered the elementary quantum of action h, taught here from 1889 to 1928."
Intensity of light emitted from a black body. Each curve represents behavior at different body temperatures. The Planck constant h is used to explain the shape of these curves.

Planck's constant was formulated as part of Max Planck's successful effort to produce a mathematical expression that accurately predicted the observed spectral distribution of thermal radiation from a closed furnace (black-body radiation).[7] This mathematical expression is now known as Planck's law.

In the last years of the 19th century, Max Planck was investigating the problem of black-body radiation first posed by

Lord Rayleigh had derived theoretically a formula, now known as the Rayleigh–Jeans law
, that could reasonably predict long wavelengths but failed dramatically at short wavelengths.

Approaching this problem, Planck hypothesized that the equations of motion for light describe a set of harmonic oscillators, one for each possible frequency. He examined how the entropy of the oscillators varied with the temperature of the body, trying to match Wien's law, and was able to derive an approximate mathematical function for the black-body spectrum,[2] which gave a simple empirical formula for long wavelengths.

Planck tried to find a mathematical expression that could reproduce Wien's law (for short wavelengths) and the empirical formula (for long wavelengths). This expression included a constant, , which is thought to be for Hilfsgrösse (auxiliary variable),

absolute temperature
T is given by

,

where is the Boltzmann constant, is the Planck constant, and is the speed of light in the medium, whether material or vacuum.[9][10][11]

The spectral radiance of a body, , describes the amount of energy it emits at different radiation frequencies. It is the power emitted per unit area of the body, per unit solid angle of emission, per unit frequency. The spectral radiance can also be expressed per unit wavelength instead of per unit frequency. In this case, it is given by

,

showing how radiated energy emitted at shorter wavelengths increases more rapidly with temperature than energy emitted at longer wavelengths.[12]

Planck's law may also be expressed in other terms, such as the number of photons emitted at a certain wavelength, or the energy density in a volume of radiation. The SI units of are W·sr−1·m−2·Hz−1, while those of are W·sr−1·m−3.

Planck soon realized that his solution was not unique. There were several different solutions, each of which gave a different value for the entropy of the oscillators.[2] To save his theory, Planck resorted to using the then-controversial theory of statistical mechanics,[2] which he described as "an act of desperation".[13] One of his new boundary conditions was

to interpret UN [the vibrational energy of N oscillators] not as a continuous, infinitely divisible quantity, but as a discrete quantity composed of an integral number of finite equal parts. Let us call each such part the energy element ε;

— Planck, On the Law of Distribution of Energy in the Normal Spectrum[2]

With this new condition, Planck had imposed the quantization of the energy of the oscillators, "a purely formal assumption ... actually I did not think much about it ..." in his own words,

Planck–Einstein relation
":

Planck was able to calculate the value of from experimental data on black-body radiation: his result, 6.55×10−34 J⋅s, is within 1.2% of the currently defined value.[2] He also made the first determination of the Boltzmann constant from the same data and theory.[15]

The observed Planck curves at different temperatures, and the divergence of the theoretical Rayleigh–Jeans (black) curve from the observed Planck curve at 5000K.

Development and application

The black-body problem was revisited in 1905, when Lord Rayleigh and James Jeans (on the one hand) and Albert Einstein (on the other hand) independently proved that classical electromagnetism could never account for the observed spectrum. These proofs are commonly known as the "ultraviolet catastrophe", a name coined by Paul Ehrenfest in 1911. They contributed greatly (along with Einstein's work on the photoelectric effect) in convincing physicists that Planck's postulate of quantized energy levels was more than a mere mathematical formalism. The first Solvay Conference in 1911 was devoted to "the theory of radiation and quanta".[16]

Photoelectric effect

The photoelectric effect is the emission of electrons (called "photoelectrons") from a surface when light is shone on it. It was first observed by

Alexandre Edmond Becquerel in 1839, although credit is usually reserved for Heinrich Hertz,[17] who published the first thorough investigation in 1887. Another particularly thorough investigation was published by Philipp Lenard (Lénárd Fülöp) in 1902.[18] Einstein's 1905 paper[19] discussing the effect in terms of light quanta would earn him the Nobel Prize in 1921,[17] after his predictions had been confirmed by the experimental work of Robert Andrews Millikan.[20] The Nobel committee awarded the prize for his work on the photo-electric effect, rather than relativity, both because of a bias against purely theoretical physics not grounded in discovery or experiment, and dissent amongst its members as to the actual proof that relativity was real.[21][22]

Before Einstein's paper, electromagnetic radiation such as visible light was considered to behave as a wave: hence the use of the terms "frequency" and "wavelength" to characterize different types of radiation. The energy transferred by a wave in a given time is called its intensity. The light from a theatre spotlight is more intense than the light from a domestic lightbulb; that is to say that the spotlight gives out more energy per unit time and per unit space (and hence consumes more electricity) than the ordinary bulb, even though the color of the light might be very similar. Other waves, such as sound or the waves crashing against a seafront, also have their intensity. However, the energy account of the photoelectric effect did not seem to agree with the wave description of light.

The "photoelectrons" emitted as a result of the photoelectric effect have a certain kinetic energy, which can be measured. This kinetic energy (for each photoelectron) is independent of the intensity of the light,[18] but depends linearly on the frequency;[20] and if the frequency is too low (corresponding to a photon energy that is less than the work function of the material), no photoelectrons are emitted at all, unless a plurality of photons, whose energetic sum is greater than the energy of the photoelectrons, acts virtually simultaneously (multiphoton effect).[23] Assuming the frequency is high enough to cause the photoelectric effect, a rise in intensity of the light source causes more photoelectrons to be emitted with the same kinetic energy, rather than the same number of photoelectrons to be emitted with higher kinetic energy.[18]

Einstein's explanation for these observations was that light itself is quantized; that the energy of light is not transferred continuously as in a classical wave, but only in small "packets" or quanta. The size of these "packets" of energy, which would later be named photons, was to be the same as Planck's "energy element", giving the modern version of the Planck–Einstein relation:

Einstein's postulate was later proven experimentally: the constant of proportionality between the frequency of incident light and the kinetic energy of photoelectrons was shown to be equal to the Planck constant .[20]

Atomic structure

A schematization of the Bohr model of the hydrogen atom. The transition shown from the n = 3 level to the n = 2 level gives rise to visible light of wavelength 656 nm (red), as the model predicts.

In 1912 John William Nicholson developed[24] an atomic model and found the angular momentum of the electrons in the model were related by h/2π.[25][26] Nicholson's nuclear quantum atomic model influenced the development of Niels Bohr 's atomic model[27][28][26] and Bohr quoted him in his 1913 paper of the Bohr model of the atom.[29] Bohr's model went beyond Planck's abstract harmonic oscillator concept: an electron in a Bohr atom could only have certain defined energies

where is the speed of light in vacuum, is an experimentally determined constant (the Rydberg constant) and . This approach also allowed Bohr to account for the Rydberg formula, an empirical description of the atomic spectrum of hydrogen, and to account for the value of the Rydberg constant in terms of other fundamental constants. In discussing angular momentum of the electrons in his model Bohr introduced the quantity , now known as the reduced Planck constant as the quantum of angular momentum.[29]

Uncertainty principle

The Planck constant also occurs in statements of Werner Heisenberg's uncertainty principle. Given numerous particles prepared in the same state, the uncertainty in their position, , and the uncertainty in their momentum, , obey

where the uncertainty is given as the

conjugate variables
which obey a similar rule. One example is time vs. energy. The inverse relationship between the uncertainty of the two conjugate variables forces a tradeoff in quantum experiments, as measuring one quantity more precisely results in the other quantity becoming imprecise.

In addition to some assumptions underlying the interpretation of certain values in the quantum mechanical formulation, one of the fundamental cornerstones to the entire theory lies in the commutator relationship between the position operator and the momentum operator :

where is the Kronecker delta.

Photon energy

The Planck relation connects the particular photon energy E with its associated wave frequency f:

This energy is extremely small in terms of ordinarily perceived everyday objects.

Since the frequency f, wavelength λ, and speed of light c are related by , the relation can also be expressed as

de Broglie wavelength

In 1923,

de Broglie wavelength
λ of the particle is given by

where p denotes the linear momentum of a particle, such as a photon, or any other elementary particle.

The energy of a photon with angular frequency ω = 2πf is given by

while its linear momentum relates to

where k is an angular wavenumber.

These two relations are the temporal and spatial parts of the special relativistic expression using

4-vectors
.

Statistical mechanics

Classical

hidden, but quantum laws constrain them based on their action. This view has been replaced by fully modern quantum theory, in which definite trajectories of motion do not even exist; rather, the particle is represented by a wavefunction spread out in space and in time.[32]
: 373  Related to this is the concept of energy quantization which existed in old quantum theory and also exists in altered form in modern quantum physics. Classical physics cannot explain quantization of energy.

Dimension and value

The Planck constant has the same

dimensions as action and as angular momentum. In SI units, the Planck constant is expressed with the unit joule per hertz
(J⋅Hz−1) or joule-second (J⋅s).

= 6.62607015×10−34 J⋅Hz−1[5]
= 1.054571817...×10−34 J⋅s[33] = 6.582119569...×10−16 eV⋅s.[34]

The above values have been adopted as fixed in the 2019 redefinition of the SI base units.

Since 2019, the numerical value of the Planck constant has been fixed, with a

caesium-133 atom ΔνCs."[35] Technologies of mass metrology such as the Kibble balance
measure refine the value of kilogram applying fixed value of the Planck constant.

Significance of the value

The Planck constant is one of the smallest constants used in physics. This reflects the fact that on a scale adapted to humans, where energies are typical of the order of kilojoules and times are typical of the order of seconds or minutes, the Planck constant is very small. When the product of energy and time for a physical event approaches the Planck constant, quantum effects dominate.[36]

Equivalently, the order of the Planck constant reflects the fact that everyday objects and systems are made of a large number of microscopic particles. For example, in green light (with a wavelength of 555 nanometres or a frequency of 540 THz) each photon has an energy E = hf = 3.58×10−19 J. That is a very small amount of energy in terms of everyday experience, but everyday experience is not concerned with individual photons any more than with individual atoms or molecules. An amount of light more typical in everyday experience (though much larger than the smallest amount perceivable by the human eye) is the energy of one mole of photons; its energy can be computed by multiplying the photon energy by the Avogadro constant, NA = 6.02214076×1023 mol−1[37], with the result of 216 kJ, about the food energy in three apples.[citation needed]

Reduced Planck constant ℏ

In many applications, the Planck constant naturally appears in combination with as , which can be traced to the fact that in these applications it is natural to use the angular frequency (in radians per second) rather than plain frequency (in cycles per second or hertz). For this reason, it is often useful to absorb that factor of 2π into the Planck constant by introducing the reduced Planck constant[38][39]: 482 (or reduced Planck's constant[40]: 5 [41]: 788), equal to the Planck constant divided by [38] and denoted by (pronounced h-bar[42]: 336).

Many of the most important equations, relations, definitions, and results of quantum mechanics are customarily written using the reduced Planck constant rather than the Planck constant , including the

Heisenberg's uncertainty principle, and Planck units.[43]: 104

The fundamental equations look simpler when written using as opposed to , and it is usually rather than that gives the most reliable results when used in order-of-magnitude estimates. For example, using dimensional analysis to estimate the ionization energy of a hydrogen atom, the relevant parameters that determine the ionization energy are the mass of the electron , the electron charge , and either the Planck constant or the reduced Planck constant :

Since both constants have the same dimensions, they will enter the dimensional analysis in the same way, but with the estimate is within a factor of two, while with the error is closer to .[44]: 8–9 

Names

The reduced Planck constant is known by many other names: the rationalized Planck constant[45]: 726 [46]: 10 [47]: - (or rationalized Planck's constant[48]: 334 [49]: ix ,[50]: 112 the Dirac constant[51]: 275 [45]: 726 [52]: xv (or Dirac's constant[53]: 148 [54]: 604 [55]: 313), the Dirac [56][57]: xviii (or Dirac's [58]: 17 ), the Dirac [59]: 187 (or Dirac's [60]: 273 [61]: 14 ), and h-bar.[62]: 558[63]: 561 It is also common to refer to this as "Planck's constant"[64]: 55 [a] while retaining the relationship .

Symbols

By far the most common symbol for the reduced Planck constant is . However, there are some sources that denote it by instead, in which case they usually refer to it as the "Dirac "[90]: 43 [91]: 151 (or "Dirac's "[92]: 21).

History

The combination appeared in Niels Bohr's 1913 paper,[93]: 15 where it was denoted by .[26]: 169 [b] For the next 15 years, the combination continued to appear in the literature, but normally without a separate symbol.[94]: 180[c] Then, in 1926, in their seminal papers, Schrödinger and Dirac again introduced special symbols for it: in the case of Schrödinger,[106] and in the case of Dirac.[107] Dirac continued to use in this way until 1930,[108]: 291 when he introduced the symbol in his book The Principles of Quantum Mechanics.[108]: 291 [109]

See also

Notes

  1. ^ Notable examples of such usage include Landau and Lifshitz[65]: 20 and Giffiths,[66]: 3 but there are many others, e.g.[67][68]: 449  [69]: 284 [70]: 3 [71]: 365 [72]: 14 [73]: 18 [74]: 4 [75]: 138 [76]: 251 [77]: 1 [78]: 622 [79]: xx [80]: 20 [81]: 4 [82]: 36 [83]: 41 [84]: 199 [85]: 846  [86][87][88]: 25  [89]: 653 
  2. ^ Bohr denoted by the angular momentum of the electron around the nucleus, and wrote the quantization condition as , where is a positive integer. (See the Bohr model.)
  3. ^ Here are some papers that are mentioned in[94] and in which appeared without a separate symbol: [95]: 428  [96]: 549 [97]: 508 [98]: 230 [99]: 458  [100][101]: 276 [102][103][104].[105]

References

Citations

  1. ^ a b "Planck constant". The NIST Reference on Constants, Units, and Uncertainty. NIST. 20 May 2019. Archived from the original on 2022-05-27. Retrieved 2023-09-03.
  2. ^
    doi:10.1002/andp.19013090310, archived (PDF) from the original on 2012-06-10, retrieved 2008-12-15. English translation: "On the Law of Distribution of Energy in the Normal Spectrum". Archived from the original on 2008-04-18.". "On the Law of Distribution of Energy in the Normal Spectrum" (PDF). Archived from the original
    (PDF) on 2011-10-06. Retrieved 2011-10-13.
  3. ^ "Max Planck Nobel Lecture". Archived from the original on 2023-07-14. Retrieved 2023-07-14.
  4. ^ a b "2022 CODATA Value: Planck constant". The NIST Reference on Constants, Units, and Uncertainty. NIST. May 2024. Retrieved 2024-05-18.
  5. ^ "Resolutions of the 26th CGPM" (PDF). BIPM. 2018-11-16. Archived from the original (PDF) on 2018-11-19. Retrieved 2018-11-20.
  6. ^ a b Bitter, Francis; Medicus, Heinrich A. (1973). Fields and particles. New York: Elsevier. pp. 137–144.
  7. .
  8. .
  9. .
  10. from the original on 2020-07-27. Retrieved 2020-05-20.
  11. .
  12. ^ Kragh, Helge (1 December 2000), Max Planck: the reluctant revolutionary, PhysicsWorld.com, archived from the original on 2009-01-08
  13. from the original on 2021-12-06, retrieved 2021-10-31
  14. ^ Planck, Max (2 June 1920), The Genesis and Present State of Development of the Quantum Theory (Nobel Lecture), archived from the original on 15 July 2011, retrieved 13 December 2008
  15. ^ Previous Solvay Conferences on Physics, International Solvay Institutes, archived from the original on 16 December 2008, retrieved 12 December 2008
  16. ^ a b See, e.g., Arrhenius, Svante (10 December 1922), Presentation speech of the 1921 Nobel Prize for Physics, archived from the original on 4 September 2011, retrieved 13 December 2008
  17. ^ from the original on 2019-08-18, retrieved 2019-07-03
  18. (PDF) from the original on 2011-07-09, retrieved 2009-12-03
  19. ^
  20. from the original on 2020-01-09, retrieved 2021-10-31, pp. 309–314.
  21. ^ "The Nobel Prize in Physics 1921". Nobelprize.org. Archived from the original on 2018-07-03. Retrieved 2014-04-23.
  22. .
  23. ^
  24. ^ Heilbron, John L. (2013). "The path to the quantum atom".
    S2CID 4355108
    .
  25. ^ .
  26. ^ Hirosige, Tetu; Nisio, Sigeko (1964). "Formation of Bohr's theory of atomic constitution". Japanese Studies in History of Science. 3: 6–28.
  27. ^ J. L. Heilbron, A History of Atomic Models from the Discovery of the Electron to the Beginnings of Quantum Mechanics, diss. (University of California, Berkeley, 1964).
  28. ^ from the original on 2023-03-07. Retrieved 2023-07-23.
  29. from the original on 2021-12-06, retrieved 2021-10-31
  30. .
  31. S2CID 57559543, archived from the original
    (PDF) on 2012-04-15, The question is first: How can one assign a discrete succession of energy values Hσ to a system specified in the sense of classical mechanics (the energy function is a given function of the coordinates qr and the corresponding momenta pr)? The Planck constant h relates the frequency Hσ/h to the energy values Hσ. It is therefore sufficient to give to the system a succession of discrete frequency values.
  32. ^ "2022 CODATA Value: reduced Planck constant". The NIST Reference on Constants, Units, and Uncertainty. NIST. May 2024. Retrieved 2024-05-18.
  33. ^ "CODATA Value: reduced Planck constant in eV s". physics.nist.gov.
  34. ^ "The Feynman Lectures on Physics Vol. II Ch. 19: The Principle of Least Action". www.feynmanlectures.caltech.edu. Retrieved 2023-11-03.
  35. ^ "2022 CODATA Value: Avogadro constant". The NIST Reference on Constants, Units, and Uncertainty. NIST. May 2024. Retrieved 2024-05-18.
  36. ^ a b "reduced Planck constant". The NIST Reference on Constants, Units, and Uncertainty. NIST. 20 May 2019. Archived from the original on 2023-04-08. Retrieved 2023-09-03.
  37. .
  38. .
  39. .
  40. .
  41. .
  42. (PDF) on 2023-10-14.
  43. ^ a b Rennie, Richard; Law, Jonathan, eds. (2017). "Planck constant". A Dictionary of Physics. Oxford Quick Reference (7th ed.). Oxford, UK: OUP Oxford. .
  44. ^ The International Encyclopedia of Physical Chemistry and Chemical Physics. Pergamon Press. 1960.
  45. ^ Vértes, Attila; Nagy, Sándor; Klencsár, Zoltán; Lovas, Rezso György; Rösch, Frank (10 December 2010). Handbook of Nuclear Chemistry. Springer Science & Business Media. .
  46. ^ Bethe, Hans A.; Salpeter, Edwin E. (1957). "Quantum Mechanics of One- and Two-Electron Atoms". In Flügge, Siegfried (ed.). Handbuch der Physik: Atome I-II. Springer.
  47. ^ Lang, Kenneth (11 November 2013). Astrophysical Formulae: A Compendium for the Physicist and Astrophysicist. Springer Science & Business Media. .
  48. ^ Galgani, L.; Carati, A.; Pozzi, B. (December 2002). "The Problem of the Rate of Thermalization, and the Relations between Classical and Quantum Mechanics". In Fabrizio, Mauro; Morro, Angelo (eds.). Mathematical Models and Methods for Smart Materials, Cortona, Italy, 25 – 29 June 2001. pp. 111–122. .
  49. ^ Fox, Mark (14 June 2018). A Student's Guide to Atomic Physics. Cambridge University Press. .
  50. ^ Kleiss, Ronald (10 June 2021). Quantum Field Theory: A Diagrammatic Approach. Cambridge University Press. .
  51. ^ Zohuri, Bahman (5 January 2021). Thermal Effects of High Power Laser Energy on Materials. Springer Nature. .
  52. ^ .
  53. ^ Chen, C. Julian (15 August 2011). Physics of Solar Energy. John Wiley & Sons. .
  54. ^ "Dirac h". Britannica. Archived from the original on 2023-02-17. Retrieved 2023-09-27.
  55. ^ .
  56. ^ Powell, John L.; Crasemann, Bernd (5 May 2015). Quantum Mechanics. Courier Dover Publications. .
  57. ^ .
  58. ^ Johnson, R. E. (6 December 2012). Introduction to Atomic and Molecular Collisions. Springer Science & Business Media. .
  59. ^ Garcia, Alejandro; .
  60. ^ .
  61. ^ .
  62. ^ .
  63. ^ .
  64. ^ .
  65. ^ "Planck's constant". The Great Soviet Encyclopedia (1970–1979, 3rd ed.). The Gale Group.
  66. ^ .
  67. ^ .
  68. ^ .
  69. ^ .
  70. ^ .
  71. ^ .
  72. ^ Buchbinder, Iosif L.; Shapiro, Ilya (March 2021). Introduction to Quantum Field Theory with Applications to Quantum Gravity. Oxford University Press. .
  73. ^ .
  74. ^ .
  75. ^ Casalbuoni, Roberto (6 April 2017). Introduction To Quantum Field Theory (Second ed.). World Scientific Publishing Company. .
  76. ^ .
  77. ^ Desai, Bipin R. (2010). Quantum Mechanics with Basic Field Theory. Cambridge University Press. .
  78. ^ Donoghue, John; Sorbo, Lorenzo (8 March 2022). A Prelude to Quantum Field Theory. Princeton University Press. .
  79. ^ .
  80. ^ Fradkin, Eduardo (23 March 2021). Quantum Field Theory: An Integrated Approach. Princeton University Press. .
  81. ^ Gelis, François (11 July 2019). Quantum Field Theory. Cambridge University Press. .
  82. ^ .
  83. ^ .
  84. ^
    S2CID 45829793
    .
  85. ^ Kocia, Lucas; Love, Peter (12 July 2018). "Measurement contextuality and Planck's constant". New Journal of Physics. 20 (7): 073020.
    S2CID 73623448
    .
  86. ^ Humpherys, David (28 November 2022). "The Implicit Structure of Planck's Constant". European Journal of Applied Physics. 4 (6): 22–25.
    S2CID 254359279
    .
  87. ^ Bais, F. Alexander; .
  88. ^ Hirota, E.; Sakakima, H.; Inomata, K. (9 March 2013). Giant Magneto-Resistance Devices. Springer Science & Business Media. .
  89. ^ Gardner, John H. (1988). "An Invariance Theory". Encyclia. 65: 139.
  90. ^ .
  91. .
  92. ^ .
  93. .
  94. ^ Schwarzschild, K. (1916). "Zur Quantenhypothese". Sitzungsberichte der Königlich Preussischen Akademie der Wissenschaften zu Berlin: 548–568.
  95. .
  96. ^ Landé, A. (June 1919). "Das Serienspektrum des Heliums". Physikalische Zeitschrift. 20: 228–234.
  97. .
  98. .
  99. .
  100. .
  101. .
  102. .
  103. .
  104. .
  105. .
  106. ^ a b Mehra, Jagdish; Rechenberg, Helmut (2000). The Historical Development of Quantum Theory. Vol. 6. New York: Springer.
  107. ^ Dirac, P. A. M. (1930). The Principles of Quantum Mechanics (1st ed.). Oxford, U.K.: Clarendon.

Sources

External links