Planck constant

Last updated
Planck constant
Common symbols
SI unit joule per hertz (joule second)
Other units
electronvolt per hertz (electronvolt second)
In SI base units kgm 2s −1
Dimension
Value 6.62607015×10−34 J⋅Hz−1
4.135667696...×10−15 eV⋅Hz−1
Reduced Planck constant
Common symbols
SI unit joule-second
Other units
electronvolt-second
In SI base units kgm 2s −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 closely related reduced Planck constant, equal to and denoted is commonly used in quantum physics equations.

Contents

The constant was postulated by Max Planck in 1900 as a proportionality constant needed to explain experimental 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 metrology, the Planck constant is used, together with other constants, to define the kilogram, the 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: "In this edifice taught Max Planck, the discoverer of the elementary quantum of action h, from 1889 to 1928." Max Planck Wirkungsquantums 20050815.jpg
Plaque at the Humboldt University of Berlin: "In this edifice taught Max Planck, the discoverer of the elementary quantum of action h, 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. Wiens law.svg
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 Kirchhoff some 40 years earlier. Every physical body spontaneously and continuously emits electromagnetic radiation. There was no expression or explanation for the overall shape of the observed emission spectrum. At the time, Wien's law fit the data for short wavelengths and high temperatures, but failed for long wavelengths. [7] :141 Also around this time, but unknown to Planck, 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), [8] and subsequently became known as the Planck constant. The expression formulated by Planck showed that the spectral radiance per unit frequency of a body for frequency ν at 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. Substituting in the relation above we get

,

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 unit of is W·sr −1·m −2·Hz −1, while that of is 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, [14] but one that would revolutionize physics. Applying this new approach to Wien's displacement law showed that the "energy element" must be proportional to the frequency of the oscillator, the first version of what is now sometimes termed the "Planck–Einstein relation":

Planck was able to calculate the value of from experimental data on black-body radiation: his result, 6.55×10−34 Js, 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. Black body.svg
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 (together) and Albert Einstein 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. Bohr atom model.svg
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 standard deviation of the measured value from its expected value. There are several other such pairs of physically measurable 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, Louis de Broglie generalized the Planck–Einstein relation by postulating that the Planck constant represents the proportionality between the momentum and the quantum wavelength of not just the photon, but the quantum wavelength of any particle. This was confirmed by experiments soon afterward. This holds throughout the quantum theory, including electrodynamics. The 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 statistical mechanics requires the existence of h (but does not define its value). [30] Eventually, following upon Planck's discovery, it was speculated that physical action could not take on an arbitrary value, but instead was restricted to integer multiples of a very small quantity, the "[elementary] quantum of action", now called the Planck constant. [31] This was a significant conceptual part of the so-called "old quantum theory" developed by physicists including Bohr, Sommerfeld, and Ishiwara, in which particle trajectories exist but are 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 revision of the SI.

Since 2019, the numerical value of the Planck constant has been fixed, with a finite decimal representation. This fixed value is used to define the SI unit of mass, the kilogram: "the kilogram [...] is defined by taking the fixed numerical value of h to be 6.62607015×10−34 when expressed in the unit J⋅s, which is equal to kg⋅m2⋅s−1, where the metre and the second are defined in terms of speed of light c and duration of hyperfine transition of the ground state of an unperturbed 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

Many equations in quantum physics are customarily written using the reduced Planck constant, [38] : 104 equal to and denoted (pronounced h-bar [39] : 336 ). [40]

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 . [41] :8–9

Names and symbols

The reduced Planck constant is known by many other names: reduced Planck's constant [42] : 5 [43] : 788 ), the rationalized Planck constant [44] : 726 [45] : 10 [46] : - (or rationalized Planck's constant [47] : 334 [48] : ix , [49] : 112 the Dirac constant [50] : 275 [44] : 726 [51] : xv (or Dirac's constant [52] : 148 [53] : 604 [54] : 313 ), the Dirac [55] [56] : xviii (or Dirac's [57] : 17 ), the Dirac [58] : 187 (or Dirac's [59] : 273 [60] : 14 ), and h-bar. [61] : 558 [62] : 561 It is also common to refer to this as "Planck's constant" [63] : 55 [lower-alpha 1] while retaining the relationship .

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 " [89] : 43 [90] (or "Dirac's " [91] : 21 ).

History

The combination appeared in Niels Bohr's 1913 paper, [92] : 15 where it was denoted by . [26] :169 [lower-alpha 2] For the next 15 years, the combination continued to appear in the literature, but normally without a separate symbol. [93] : 180 [lower-alpha 3] Then, in 1926, in their seminal papers, Schrödinger and Dirac again introduced special symbols for it: in the case of Schrödinger, [105] and in the case of Dirac. [106] Dirac continued to use in this way until 1930, [107] : 291 when he introduced the symbol in his book The Principles of Quantum Mechanics. [107] : 291 [108]

See also

Notes

  1. Notable examples of such usage include Landau and Lifshitz [64] : 20 and Griffiths, [65] : 3 but there are many others, e.g. [66] [67] :449 [68] : 284 [69] : 3 [70] : 365 [71] : 14 [72] : 18 [73] : 4 [74] : 138 [75] : 251 [76] : 1 [77] : 622 [78] : xx [79] : 20 [80] : 4 [81] : 36 [82] : 41 [83] : 199 [84] :846 [85] [86] [87] :25 [88] :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 [93] and in which appeared without a separate symbol: [94] :428 [95] : 549 [96] : 508 [97] : 230 [98] :458 [99] [100] : 276 [101] [102] [103] . [104]

Related Research Articles

<span class="mw-page-title-main">Bohr model</span> Atomic model introduced by Niels Bohr in 1913

In atomic physics, the Bohr model or Rutherford–Bohr model was the first successful model of the atom. Developed from 1911 to 1918 by Niels Bohr and building on Ernest Rutherford's nuclear model, it supplanted the plum pudding model of J J Thomson only to be replaced by the quantum atomic model in the 1920s. It consists of a small, dense nucleus surrounded by orbiting electrons. It is analogous to the structure of the Solar System, but with attraction provided by electrostatic force rather than gravity, and with the electron energies quantized.

A photon is an elementary particle that is a quantum of the electromagnetic field, including electromagnetic radiation such as light and radio waves, and the force carrier for the electromagnetic force. Photons are massless particles that always move at the speed of light measured in vacuum. The photon belongs to the class of boson particles.

<span class="mw-page-title-main">Quantum mechanics</span> Description of physical properties at the atomic and subatomic scale

Quantum mechanics is a fundamental theory that describes the behavior of nature at and below the scale of atoms. It is the foundation of all quantum physics, which includes quantum chemistry, quantum field theory, quantum technology, and quantum information science.

<span class="mw-page-title-main">Schrödinger equation</span> Description of a quantum-mechanical system

The Schrödinger equation is a partial differential equation that governs the wave function of a quantum-mechanical system. Its discovery was a significant landmark in the development of quantum mechanics. It is named after Erwin Schrödinger, who postulated the equation in 1925 and published it in 1926, forming the basis for the work that resulted in his Nobel Prize in Physics in 1933.

<span class="mw-page-title-main">Wave function</span> Mathematical description of quantum state

In quantum physics, a wave function is a mathematical description of the quantum state of an isolated quantum system. The most common symbols for a wave function are the Greek letters ψ and Ψ. Wave functions are complex-valued. For example, a wave function might assign a complex number to each point in a region of space. The Born rule provides the means to turn these complex probability amplitudes into actual probabilities. In one common form, it says that the squared modulus of a wave function that depends upon position is the probability density of measuring a particle as being at a given place. The integral of a wavefunction's squared modulus over all the system's degrees of freedom must be equal to 1, a condition called normalization. Since the wave function is complex-valued, only its relative phase and relative magnitude can be measured; its value does not, in isolation, tell anything about the magnitudes or directions of measurable observables. One has to apply quantum operators, whose eigenvalues correspond to sets of possible results of measurements, to the wave function ψ and calculate the statistical distributions for measurable quantities.

<span class="mw-page-title-main">Wavenumber</span> Spatial frequency of a wave

In the physical sciences, the wavenumber, also known as repetency, is the spatial frequency of a wave, measured in cycles per unit distance or radians per unit distance. It is analogous to temporal frequency, which is defined as the number of wave cycles per unit time or radians per unit time.

The Bohr radius is a physical constant, approximately equal to the most probable distance between the nucleus and the electron in a hydrogen atom in its ground state. It is named after Niels Bohr, due to its role in the Bohr model of an atom. Its value is 5.29177210544(82)×10−11 m.

<span class="mw-page-title-main">Planck's law</span> Spectral density of light emitted by a black body

In physics, Planck's law describes the spectral density of electromagnetic radiation emitted by a black body in thermal equilibrium at a given temperature T, when there is no net flow of matter or energy between the body and its environment.

In spectroscopy, the Rydberg constant, symbol for heavy atoms or for hydrogen, named after the Swedish physicist Johannes Rydberg, is a physical constant relating to the electromagnetic spectra of an atom. The constant first arose as an empirical fitting parameter in the Rydberg formula for the hydrogen spectral series, but Niels Bohr later showed that its value could be calculated from more fundamental constants according to his model of the atom.

<span class="mw-page-title-main">Ultraviolet catastrophe</span> Classical physics prediction that black body radiation grows unbounded with frequency

The ultraviolet catastrophe, also called the Rayleigh–Jeans catastrophe, was the prediction of late 19th century to early 20th century classical physics that an ideal black body at thermal equilibrium would emit an unbounded quantity of energy as wavelength decreased into the ultraviolet range. The term "ultraviolet catastrophe" was first used in 1911 by Paul Ehrenfest, but the concept originated with the 1900 statistical derivation of the Rayleigh–Jeans law.

Matter waves are a central part of the theory of quantum mechanics, being half of wave–particle duality. At all scales where measurements have been practical, matter exhibits wave-like behavior. For example, a beam of electrons can be diffracted just like a beam of light or a water wave.

<span class="mw-page-title-main">Rydberg formula</span> Formula for spectral line wavelengths in alkali metals

In atomic physics, the Rydberg formula calculates the wavelengths of a spectral line in many chemical elements. The formula was primarily presented as a generalization of the Balmer series for all atomic electron transitions of hydrogen. It was first empirically stated in 1888 by the Swedish physicist Johannes Rydberg, then theoretically by Niels Bohr in 1913, who used a primitive form of quantum mechanics. The formula directly generalizes the equations used to calculate the wavelengths of the hydrogen spectral series.

The old quantum theory is a collection of results from the years 1900–1925 which predate modern quantum mechanics. The theory was never complete or self-consistent, but was instead a set of heuristic corrections to classical mechanics. The theory has come to be understood as the semi-classical approximation to modern quantum mechanics. The main and final accomplishments of the old quantum theory were the determination of the modern form of the periodic table by Edmund Stoner and the Pauli exclusion principle which were both premised on the Arnold Sommerfeld enhancements to the Bohr model of the atom.

The Compton wavelength is a quantum mechanical property of a particle, defined as the wavelength of a photon whose energy is the same as the rest energy of that particle. It was introduced by Arthur Compton in 1923 in his explanation of the scattering of photons by electrons.

In atomic physics, the electron magnetic moment, or more specifically the electron magnetic dipole moment, is the magnetic moment of an electron resulting from its intrinsic properties of spin and electric charge. The value of the electron magnetic moment is −9.2847646917(29)×10−24 J⋅T−1. In units of the Bohr magneton (μB), it is −1.00115965218059(13) μB, a value that was measured with a relative accuracy of 1.3×10−13.

<span class="mw-page-title-main">Einstein coefficients</span> Quantities describing probability of absorption or emission of light

In atomic, molecular, and optical physics, the Einstein coefficients are quantities describing the probability of absorption or emission of a photon by an atom or molecule. The Einstein A coefficients are related to the rate of spontaneous emission of light, and the Einstein B coefficients are related to the absorption and stimulated emission of light. Throughout this article, "light" refers to any electromagnetic radiation, not necessarily in the visible spectrum.

The Rydberg–Ritz combination principle is an empirical rule proposed by Walther Ritz in 1908 to describe the relationship of the spectral lines for all atoms, as a generalization of an earlier rule by Johannes Rydberg for the hydrogen atom and the alkali metals. The principle states that the spectral lines of any element include frequencies that are either the sum or the difference of the frequencies of two other lines. Lines of the spectra of elements could be predicted from existing lines. Since the frequency of light is proportional to the wavenumber or reciprocal wavelength, the principle can also be expressed in terms of wavenumbers which are the sum or difference of wavenumbers of two other lines.

The history of quantum mechanics is a fundamental part of the history of modern physics. The major chapters of this history begin with the emergence of quantum ideas to explain individual phenomena—blackbody radiation, the photoelectric effect, solar emission spectra—an era called the Old or Older quantum theories. Building on the technology developed in classical mechanics, the invention of wave mechanics by Erwin Schrödinger and expansion by many others triggers the "modern" era beginning around 1925. Paul Dirac's relativistic quantum theory work lead him to explore quantum theories of radiation, culminating in quantum electrodynamics, the first quantum field theory. The history of quantum mechanics continues in the history of quantum field theory. The history of quantum chemistry, theoretical basis of chemical structure, reactivity, and bonding, interlaces with the events discussed in this article.

In particle physics and physical cosmology, Planck units are a system of units of measurement defined exclusively in terms of four universal physical constants: c, G, ħ, and kB. Expressing one of these physical constants in terms of Planck units yields a numerical value of 1. They are a system of natural units, defined using fundamental properties of nature rather than properties of a chosen prototype object. Originally proposed in 1899 by German physicist Max Planck, they are relevant in research on unified theories such as quantum gravity.

The Planck relation is a fundamental equation in quantum mechanics which states that the energy E of a photon, known as photon energy, is proportional to its frequency ν: The constant of proportionality, h, is known as the Planck constant. Several equivalent forms of the relation exist, including in terms of angular frequency ω: where . Written using the symbol f for frequency, the relation is

References

Citations

  1. 1 2 "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. 1 2 3 4 5 6 Planck, Max (1901), "Ueber das Gesetz der Energieverteilung im Normalspectrum" (PDF), Ann. Phys. , 309 (3): 553–63, Bibcode:1901AnP...309..553P, 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. The International System of Units (PDF) (9th ed.), International Bureau of Weights and Measures, Dec 2022, p. 131, ISBN   978-92-822-2272-0
  5. 1 2 "2022 CODATA Value: Planck constant". The NIST Reference on Constants, Units, and Uncertainty. NIST. May 2024. Retrieved 2024-05-18.
  6. "Resolutions of the 26th CGPM" (PDF). BIPM. 2018-11-16. Archived from the original (PDF) on 2018-11-19. Retrieved 2018-11-20.
  7. 1 2 Bitter, Francis; Medicus, Heinrich A. (1973). Fields and particles. New York: Elsevier. pp. 137–144.
  8. Boya, Luis J. (2004). "The Thermal Radiation Formula of Planck (1900)". arXiv: physics/0402064v1 .
  9. Planck, M. (1914). The Theory of Heat Radiation. Masius, M. (transl.) (2nd ed.). P. Blakiston's Son. pp. 6, 168. OL   7154661M.
  10. Chandrasekhar, S. (1960) [1950]. Radiative Transfer (Revised reprint ed.). Dover. p. 8. ISBN   978-0-486-60590-6.
  11. Rybicki, G. B.; Lightman, A. P. (1979). Radiative Processes in Astrophysics. Wiley. p. 22. ISBN   978-0-471-82759-7. Archived from the original on 2020-07-27. Retrieved 2020-05-20.
  12. Shao, Gaofeng; et al. (2019). "Improved oxidation resistance of high emissivity coatings on fibrous ceramic for reusable space systems". Corrosion Science. 146: 233–246. arXiv: 1902.03943 . Bibcode:2019Corro.146..233S. doi:10.1016/j.corsci.2018.11.006. S2CID   118927116.
  13. Kragh, Helge (1 December 2000), Max Planck: the reluctant revolutionary, PhysicsWorld.com, archived from the original on 2009-01-08
  14. Kragh, Helge (1999), Quantum Generations: A History of Physics in the Twentieth Century, Princeton University Press, p. 62, ISBN   978-0-691-09552-3, archived from the original on 2021-12-06, retrieved 2021-10-31
  15. 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
  16. Previous Solvay Conferences on Physics, International Solvay Institutes, archived from the original on 16 December 2008, retrieved 12 December 2008
  17. 1 2 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
  18. 1 2 3 Lenard, P. (1902), "Ueber die lichtelektrische Wirkung", Annalen der Physik , 313 (5): 149–98, Bibcode:1902AnP...313..149L, doi:10.1002/andp.19023130510, archived from the original on 2019-08-18, retrieved 2019-07-03
  19. Einstein, Albert (1905), "Über einen die Erzeugung und Verwandlung des Lichtes betreffenden heuristischen Gesichtspunkt" (PDF), Annalen der Physik , 17 (6): 132–48, Bibcode:1905AnP...322..132E, doi: 10.1002/andp.19053220607 , archived (PDF) from the original on 2011-07-09, retrieved 2009-12-03
  20. 1 2 3 Millikan, R. A. (1916), "A Direct Photoelectric Determination of Planck's h", Physical Review , 7 (3): 355–88, Bibcode:1916PhRv....7..355M, doi: 10.1103/PhysRev.7.355
  21. Isaacson, Walter (2007-04-10), Einstein: His Life and Universe, Simon and Schuster, ISBN   978-1-4165-3932-2, archived from the original on 2020-01-09, retrieved 2021-10-31, pp. 309–314.
  22. "The Nobel Prize in Physics 1921". Nobelprize.org. Archived from the original on 2018-07-03. Retrieved 2014-04-23.
  23. Heilbron, John L. (2013). "The path to the quantum atom". Nature . 498 (7452): 27–30. doi:10.1038/498027a. PMID   23739408. S2CID   4355108.
  24. 1 2 3 McCormmach, Russell (1966). "The Atomic Theory of John William Nicholson". Archive for History of Exact Sciences . 3 (2): 160–184. doi:10.1007/BF00357268. JSTOR   41133258. S2CID   120797894.
  25. Hirosige, Tetu; Nisio, Sigeko (1964). "Formation of Bohr's theory of atomic constitution". Japanese Studies in History of Science. 3: 6–28.
  26. 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).
  27. 1 2 Bohr, N. (1913). "On the constitution of atoms and molecules". The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science . 6th series. 26 (151): 1–25. Bibcode:1913PMag...26..476B. doi:10.1080/14786441308634955. Archived from the original on 2023-03-07. Retrieved 2023-07-23.
  28. Giuseppe Morandi; F. Napoli; E. Ercolessi (2001), Statistical mechanics: an intermediate course, World Scientific, p. 84, ISBN   978-981-02-4477-4, archived from the original on 2021-12-06, retrieved 2021-10-31
  29. ter Haar, D. (1967). The Old Quantum Theory . Pergamon Press. p.  133. ISBN   978-0-08-012101-7.
  30. Einstein, Albert (2003), "Physics and Reality" (PDF), Daedalus, 132 (4): 24, doi:10.1162/001152603771338742, 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.
  31. "2022 CODATA Value: reduced Planck constant". The NIST Reference on Constants, Units, and Uncertainty. NIST. May 2024. Retrieved 2024-05-18.
  32. "CODATA Value: reduced Planck constant in eV s". physics.nist.gov.
  33. The International System of Units (PDF) (9th ed.), International Bureau of Weights and Measures, Dec 2022, ISBN   978-92-822-2272-0
  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. Schwarz, Patricia M.; Schwarz, John H. (25 March 2004). Special Relativity: From Einstein to Strings. Cambridge University Press. ISBN   978-1-139-44950-2.
  37. Chabay, Ruth W.; Sherwood, Bruce A. (20 November 2017). Matter and Interactions. John Wiley & Sons. ISBN   978-1-119-45575-2.
  38. "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.
  39. Lévy-Leblond, Jean-Marc (2002). "The meanings of Planck's constant" (PDF). In Beltrametti, E.; Rimini, A.; Robotti, Nadia (eds.). One Hundred Years of H: Pavia, 14-16 September 2000. Italian Physical Society. ISBN   978-88-7438-003-9. Archived from the original (PDF) on 2023-10-14.
  40. Huang, Kerson (26 April 2010). Quantum Field Theory: From Operators to Path Integrals. John Wiley & Sons. ISBN   978-3-527-40846-7.
  41. Schmitz, Kenneth S. (11 November 2016). Physical Chemistry: Concepts and Theory. Elsevier. ISBN   978-0-12-800600-9.
  42. 1 2 Rennie, Richard; Law, Jonathan, eds. (2017). "Planck constant". A Dictionary of Physics. Oxford Quick Reference (7th ed.). Oxford, UK: OUP Oxford. ISBN   978-0198821472.
  43. The International Encyclopedia of Physical Chemistry and Chemical Physics. Pergamon Press. 1960.
  44. Vértes, Attila; Nagy, Sándor; Klencsár, Zoltán; Lovas, Rezso György; Rösch, Frank (10 December 2010). Handbook of Nuclear Chemistry. Springer Science & Business Media. ISBN   978-1-4419-0719-6.
  45. 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.
  46. Lang, Kenneth (11 November 2013). Astrophysical Formulae: A Compendium for the Physicist and Astrophysicist. Springer Science & Business Media. ISBN   978-3-662-11188-8.
  47. 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. doi:10.1142/9789812776273_0011. ISBN   978-981-238-235-1.
  48. Fox, Mark (14 June 2018). A Student's Guide to Atomic Physics. Cambridge University Press. ISBN   978-1-316-99309-5.
  49. Kleiss, Ronald (10 June 2021). Quantum Field Theory: A Diagrammatic Approach. Cambridge University Press. ISBN   978-1-108-78750-5.
  50. Zohuri, Bahman (5 January 2021). Thermal Effects of High Power Laser Energy on Materials. Springer Nature. ISBN   978-3-030-63064-5.
  51. Balian, Roger (26 June 2007). From Microphysics to Macrophysics: Methods and Applications of Statistical Physics. Volume II. Springer Science & Business Media. ISBN   978-3-540-45480-9.
  52. Chen, C. Julian (15 August 2011). Physics of Solar Energy. John Wiley & Sons. ISBN   978-1-118-04459-9.
  53. "Dirac h". Britannica. Archived from the original on 2023-02-17. Retrieved 2023-09-27.
  54. Shoenberg, D. (3 September 2009). Magnetic Oscillations in Metals. Cambridge University Press. ISBN   978-1-316-58317-3.
  55. Powell, John L.; Crasemann, Bernd (5 May 2015). Quantum Mechanics. Courier Dover Publications. ISBN   978-0-486-80478-1.
  56. Dresden, Max (6 December 2012). H.A. Kramers Between Tradition and Revolution. Springer Science & Business Media. ISBN   978-1-4612-4622-0.
  57. Johnson, R. E. (6 December 2012). Introduction to Atomic and Molecular Collisions. Springer Science & Business Media. ISBN   978-1-4684-8448-9.
  58. Garcia, Alejandro; Henley, Ernest M. (13 July 2007). Subatomic Physics (3rd ed.). World Scientific Publishing Company. ISBN   978-981-310-167-8.
  59. Holbrow, Charles H.; Lloyd, James N.; Amato, Joseph C.; Galvez, Enrique; Parks, M. Elizabeth (14 September 2010). Modern Introductory Physics. New York: Springer Science & Business Media. ISBN   978-0-387-79080-0.
  60. Polyanin, Andrei D.; Chernoutsan, Alexei (18 October 2010). A Concise Handbook of Mathematics, Physics, and Engineering Sciences. CRC Press. ISBN   978-1-4398-0640-1.
  61. Dowling, Jonathan P. (24 August 2020). Schrödinger's Web: Race to Build the Quantum Internet. CRC Press. ISBN   978-1-000-08017-9.
  62. Landau, L. D.; Lifshitz, E. M. (22 October 2013). Quantum Mechanics: Non-Relativistic Theory. Elsevier. ISBN   978-1-4831-4912-7.
  63. Griffiths, David J.; Schroeter, Darrell F. (20 November 2019). Introduction to Quantum Mechanics. Cambridge University Press. ISBN   978-1-108-10314-5.
  64. "Planck's constant". The Great Soviet Encyclopedia (1970–1979, 3rd ed.). The Gale Group.
  65. Itzykson, Claude; Zuber, Jean-Bernard (20 September 2012). Quantum Field Theory. Courier Corporation. ISBN   978-0-486-13469-7.
  66. Kaku, Michio (1993). Quantum Field Theory: A Modern Introduction. Oxford University Press. ISBN   978-0-19-507652-3.
  67. Bogoli︠u︡bov, Nikolaĭ Nikolaevich; Shirkov, Dmitriĭ Vasilʹevich (1982). Quantum Fields. Benjamin/Cummings Publishing Company, Advanced Book Program/World Science Division. ISBN   978-0-8053-0983-6.
  68. Aitchison, Ian J. R.; Hey, Anthony J. G. (17 December 2012). Gauge Theories in Particle Physics: A Practical Introduction: From Relativistic Quantum Mechanics to QED, Fourth Edition. CRC Press. ISBN   978-1-4665-1299-3.
  69. de Wit, B.; Smith, J. (2 December 2012). Field Theory in Particle Physics, Volume 1. Elsevier. ISBN   978-0-444-59622-2.
  70. Brown, Lowell S. (1992). Quantum Field Theory. Cambridge University Press. ISBN   978-0-521-46946-3.
  71. Buchbinder, Iosif L.; Shapiro, Ilya (March 2021). Introduction to Quantum Field Theory with Applications to Quantum Gravity. Oxford University Press. ISBN   978-0-19-883831-9.
  72. Jaffe, Arthur (25 March 2004). "9. Where does quantum field theory fit into the big picture?". In Cao, Tian Yu (ed.). Conceptual Foundations of Quantum Field Theory. Cambridge University Press. ISBN   978-0-521-60272-3.
  73. Cabibbo, Nicola; Maiani, Luciano; Benhar, Omar (28 July 2017). An Introduction to Gauge Theories. CRC Press. ISBN   978-1-4987-3452-3.
  74. Casalbuoni, Roberto (6 April 2017). Introduction To Quantum Field Theory (Second ed.). World Scientific Publishing Company. ISBN   978-981-314-668-6.
  75. Das, Ashok (24 July 2020). Lectures On Quantum Field Theory (2nd ed.). World Scientific. ISBN   978-981-12-2088-3.
  76. Desai, Bipin R. (2010). Quantum Mechanics with Basic Field Theory. Cambridge University Press. ISBN   978-0-521-87760-2.
  77. Donoghue, John; Sorbo, Lorenzo (8 March 2022). A Prelude to Quantum Field Theory. Princeton University Press. ISBN   978-0-691-22348-3.
  78. Folland, Gerald B. (3 February 2021). Quantum Field Theory: A Tourist Guide for Mathematicians. American Mathematical Soc. ISBN   978-1-4704-6483-7.
  79. Fradkin, Eduardo (23 March 2021). Quantum Field Theory: An Integrated Approach. Princeton University Press. ISBN   978-0-691-14908-0.
  80. Gelis, François (11 July 2019). Quantum Field Theory. Cambridge University Press. ISBN   978-1-108-48090-1.
  81. Greiner, Walter; Reinhardt, Joachim (9 March 2013). Quantum Electrodynamics. Springer Science & Business Media. ISBN   978-3-662-05246-4.
  82. Liboff, Richard L. (2003). Introductory Quantum Mechanics (4th ed.). San Francisco: Pearson Education. ISBN   978-81-317-0441-7.
  83. Barut, A. O. (1 August 1978). "The Creation of a Photon: A Heuristic Calculation of Planck's Constant ħ or the Fine Structure Constant α". Zeitschrift für Naturforschung A. 33 (8): 993–994. Bibcode:1978ZNatA..33..993B. doi: 10.1515/zna-1978-0819 . S2CID   45829793.
  84. Kocia, Lucas; Love, Peter (12 July 2018). "Measurement contextuality and Planck's constant". New Journal of Physics. 20 (7): 073020. arXiv: 1711.08066 . Bibcode:2018NJPh...20g3020K. doi:10.1088/1367-2630/aacef2. S2CID   73623448.
  85. Humpherys, David (28 November 2022). "The Implicit Structure of Planck's Constant". European Journal of Applied Physics. 4 (6): 22–25. doi: 10.24018/ejphysics.2022.4.6.227 . S2CID   254359279.
  86. Bais, F. Alexander; Farmer, J. Doyne (2008). "The Physics of Information". In Adriaans, Pieter; van Benthem, Johan (eds.). Philosophy of Information. Handbook of the Philosophy of Science. Vol. 8. Amsterdam: North-Holland. arXiv: 0708.2837 . ISBN   978-0-444-51726-5.
  87. Hirota, E.; Sakakima, H.; Inomata, K. (9 March 2013). Giant Magneto-Resistance Devices. Springer Science & Business Media. ISBN   978-3-662-04777-4.
  88. Gardner, John H. (1988). "An Invariance Theory". Encyclia. 65: 139.
  89. Levine, Raphael D. (4 June 2009). Molecular Reaction Dynamics. Cambridge University Press. ISBN   978-1-139-44287-9.
  90. Bohr, N. (July 1913). "I. On the constitution of atoms and molecules". The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science. 26 (151): 1–25. Bibcode:1913PMag...26....1B. doi:10.1080/14786441308634955.
  91. 1 2 Mehra, Jagdish; Rechenberg, Helmut (3 August 1982). The Historical Development of Quantum Theory. Vol. 1. Springer New York. ISBN   978-0-387-90642-3.
  92. Sommerfeld, A. (1915). "Zur Theorie der Balmerschen Serie" (PDF). Sitzungsberichte der mathematisch-physikalischen Klasse der K. B. Akademie der Wissenschaften zu München. 33 (198): 425–458. doi:10.1140/epjh/e2013-40053-8.
  93. Schwarzschild, K. (1916). "Zur Quantenhypothese". Sitzungsberichte der Königlich Preussischen Akademie der Wissenschaften zu Berlin: 548–568.
  94. Ehrenfest, P. (June 1917). "XLVIII. Adiabatic invariants and the theory of quanta". The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science. 33 (198): 500–513. doi:10.1080/14786440608635664.
  95. Landé, A. (June 1919). "Das Serienspektrum des Heliums". Physikalische Zeitschrift. 20: 228–234.
  96. Bohr, N. (October 1920). "Über die Serienspektra der Elemente". Zeitschrift für Physik. 2 (5): 423–469. Bibcode:1920ZPhy....2..423B. doi:10.1007/BF01329978.
  97. Stern, Otto (December 1921). "Ein Weg zur experimentellen Prüfung der Richtungsquantelung im Magnetfeld". Zeitschrift für Physik. 7 (1): 249–253. Bibcode:1921ZPhy....7..249S. doi:10.1007/BF01332793.
  98. Heisenberg, Werner (December 1922). "Zur Quantentheorie der Linienstruktur und der anomalen Zeemaneflekte". Zeitschrift für Physik. 8 (1): 273–297. Bibcode:1922ZPhy....8..273H. doi:10.1007/BF01329602.
  99. Kramers, H. A.; Pauli, W. (December 1923). "Zur Theorie der Bandenspektren". Zeitschrift für Physik. 13 (1): 351–367. Bibcode:1923ZPhy...13..351K. doi:10.1007/BF01328226.
  100. Born, M.; Jordan, P. (December 1925). "Zur Quantenmechanik". Zeitschrift für Physik. 34 (1): 858–888. Bibcode:1925ZPhy...34..858B. doi:10.1007/BF01328531.
  101. Dirac, P. A. M. (December 1925). "The fundamental equations of quantum mechanics". Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character. 109 (752): 642–653. Bibcode:1925RSPSA.109..642D. doi: 10.1098/rspa.1925.0150 .
  102. Born, M.; Heisenberg, W.; Jordan, P. (August 1926). "Zur Quantenmechanik. II". Zeitschrift für Physik. 35 (8–9): 557–615. Bibcode:1926ZPhy...35..557B. doi:10.1007/BF01379806.
  103. Schrödinger, E. (1926). "Quantisierung als Eigenwertproblem". Annalen der Physik. 384 (4): 361–376. Bibcode:1926AnP...384..361S. doi: 10.1002/andp.19263840404 .
  104. Dirac, P. A. M. (October 1926). "On the theory of quantum mechanics". Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character. 112 (762): 661–677. Bibcode:1926RSPSA.112..661D. doi: 10.1098/rspa.1926.0133 .
  105. 1 2 Mehra, Jagdish; Rechenberg, Helmut (2000). The Historical Development of Quantum Theory. Vol. 6. New York: Springer.
  106. Dirac, P. A. M. (1930). The Principles of Quantum Mechanics (1st ed.). Oxford, U.K.: Clarendon.

Sources