Atomic units

Last updated

The atomic units are a system of natural units of measurement that is especially convenient for calculations in atomic physics and related scientific fields, such as computational chemistry and atomic spectroscopy. They were originally suggested and named by the physicist Douglas Hartree. [1] Atomic units are often abbreviated "a.u." or "au", not to be confused with similar abbreviations used for astronomical units, arbitrary units, and absorbance units in other contexts.

Contents

Motivation

In the context of atomic physics, using the atomic units system can be a convenient shortcut, eliminating unnecessary symbols and numbers with very small orders of magnitude. For example, the Hamiltonian operator in the Schrödinger equation for the helium atom with standard quantities, such as when using SI units, is [2]

but adopting the convention associated with atomic units that transforms quantities into dimensionless equivalents, it becomes

In this convention, the constants , , , and all correspond to the value (see § Definition below). The distances relevant to the physics expressed in SI units are naturally on the order of , while expressed in atomic units distances are on the order of (one Bohr radius, the atomic unit of length). An additional benefit of expressing quantities using atomic units is that their values calculated and reported in atomic units do not change when values of fundamental constants are revised. The fundamental constants are built into the conversion factors between atomic units and SI.

History

Hartree defined units based on three physical constants: [1] :91

Both in order to eliminate various universal constants from the equations and also to avoid high powers of 10 in numerical work, it is convenient to express quantities in terms of units, which may be called 'atomic units', defined as follows:

Unit of length, , on the orbital mechanics the radius of the 1-quantum circular orbit of the H-atom with fixed nucleus.
Unit of charge, , the magnitude of the charge on the electron.
Unit of mass, , the mass of the electron.

Consistent with these are:

Unit of action, .
Unit of energy, [...]
Unit of time, .
 
D.R. Hartree, The Wave Mechanics of an Atom with a Non-Coulomb Central Field. Part I. Theory and Methods

Here, the modern equivalent of is the Rydberg constant , of is the electron mass , of is the Bohr radius , and of is the reduced Planck constant . Hartree's expressions that contain differ from the modern form due to a change in the definition of , as explained below.

In 1957, Bethe and Salpeter's book Quantum mechanics of one-and two-electron atoms [3] built on Hartree's units, which they called atomic units abbreviated "a.u.". They chose to use , their unit of action and angular momentum in place of Hartree's length as the base units. They noted that the unit of length in this system is the radius of the first Bohr orbit and their velocity is the electron velocity in Bohr's model of the first orbit.

In 1959, Shull and Hall [4] advocated atomic units based on Hartree's model but again chose to use as the defining unit. They explicitly named the distance unit a "Bohr radius"; in addition, they wrote the unit of energy as and called it a Hartree. These terms came to be used widely in quantum chemistry. [5] :349

In 1973 McWeeny extended the system of Shull and Hall by adding permittivity in the form of as a defining or base unit. [6] [7] Simultaneously he adopted the SI definition of so that his expression for energy in atomic units is , matching the expression in the 8th SI brochure. [8]

Definition

A set of base units in the atomic system as in one proposal are the electron rest mass, the magnitude of the electronic charge, the Planck constant, and the permittivity. [6] [9] In the atomic units system, each of these takes the value 1; the corresponding values in the International System of Units [10] :132 are given in the table.

Table notes
This choice of base units, which is essentially arbitrary, is McWeeny's proposal.
W represents the dimensions of energy, ML2T−2. [6]
In the 'atomic units' column, the convention that uses dimensionless equivalents has been applied.
Base atomic units
Symbol and NameQuantity (dimensions)Atomic unitsSI units
, reduced Planck constant action (ML2T−1)11.054571817...×10−34 J⋅s  [11]
, elementary charge charge (Q)11.602176634×10−19 C  [12]
, electron rest mass mass (M)19.1093837139(28)×10−31 kg  [13]
, permittivity permittivity (Q2W−1L−1)11.11265005620(17)×10−10 F⋅m−1  [14]

Units

Three of the defining constants (reduced Planck constant, elementary charge, and electron rest mass) are atomic units themselves – of action, [15] electric charge, [16] and mass, [17] respectively. Two named units are those of length (Bohr radius ) and energy (hartree ).

Defined atomic units
Atomic unit ofExpressionValue in SI unitsOther equivalents
electric charge density 1.08120238677(51)×1012 C⋅m−3  [18]
electric current 6.6236182375082(72)×10−3 A  [19]
electric charge 1.602176634×10−19 C  [20]
electric dipole moment 8.4783536198(13)×10−30 C⋅m  [21] 2.541746473  D
electric quadrupole moment4.4865515185(14)×10−40 C⋅m2  [22]
electric potential 27.211386245981(30) V  [23]
electric field 5.14220675112(80)×1011 V⋅m−1  [24]
electric field gradient 9.7173624424(30)×1021 V⋅m−2  [25]
permittivity 1.11265005620(17)×10−10 F⋅m−1  [14]
electric polarizability 1.64877727212(51)×10−41 C2⋅m2⋅J−1  [26]
1st hyperpolarizability 3.2063612996(15)×10−53 C3⋅m3⋅J−2  [27]
2nd hyperpolarizability6.2353799735(39)×10−65 C4⋅m4⋅J−3  [28]
magnetic dipole moment 1.85480201315(58)×10−23 J⋅T−1  [29]
magnetic flux density 2.35051757077(73)×105 T  [30] 2.3505×109  G
magnetizability 7.8910365794(49)×10−29 J⋅T−2  [31]
action 1.054571817...×10−34 J⋅s  [32]
energy 4.3597447222060(48)×10−18 J  [33] , , 27.211386245988(53)  eV   [34]
force 8.2387235038(13)×10−8 N  [35] 82.387 nN, 51.421 eV·Å−1
length 5.29177210544(82)×10−11 m  [36] , 0.529177  Å
mass 9.1093837139(28)×10−31 kg  [37]
momentum 1.99285191545(31)×10−24 kg⋅m⋅s−1  [38]
time 2.4188843265864(26)×10−17 s  [39]
velocity 2.18769126216(34)×106 m⋅s−1  [40]

:  speed of light, :  vacuum permittivity, :  Rydberg constant, : Planck constant, :  fine-structure constant, :  Bohr magneton, : correspondence

Conventions

Different conventions are adopted in the use of atomic units, which vary in presentation, formality and convenience.

Explicit units

A convention that eliminates units

In atomic physics, it is common to simplify mathematical expressions by a transformation of all quantities:

Physical constants

Dimensionless physical constants retain their values in any system of units. Of note is the fine-structure constant , which appears in expressions as a consequence of the choice of units. For example, the numeric value of the speed of light, expressed in atomic units, is [44] :597

Some physical constants expressed in atomic units
NameSymbol/DefinitionValue in atomic units
speed of light
classical electron radius
reduced Compton wavelength
of the electron
ƛe
proton mass

Bohr model in atomic units

Atomic units are chosen to reflect the properties of electrons in atoms, which is particularly clear in the classical Bohr model of the hydrogen atom for the bound electron in its ground state:

Related Research Articles

<span class="mw-page-title-main">Hydrogen atom</span> Atom of the element hydrogen

A hydrogen atom is an atom of the chemical element hydrogen. The electrically neutral atom contains a single positively charged proton and a single negatively charged electron bound to the nucleus by the Coulomb force. Atomic hydrogen constitutes about 75% of the baryonic mass of the universe.

<span class="mw-page-title-main">Fine-structure constant</span> Dimensionless number that quantifies the strength of the electromagnetic interaction

In physics, the fine-structure constant, also known as the Sommerfeld constant, commonly denoted by α, is a fundamental physical constant which quantifies the strength of the electromagnetic interaction between elementary charged particles.

Physisorption, also called physical adsorption, is a process in which the electronic structure of the atom or molecule is barely perturbed upon adsorption.

<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.

The hartree, also known as the Hartree energy, is the unit of energy in the atomic units system, named after the British physicist Douglas Hartree. Its CODATA recommended value is Eh = 4.3597447222060(48)×10−18 J = 27.211386245981(30) eV.

In atomic physics, the Bohr magneton is a physical constant and the natural unit for expressing the magnetic moment of an electron caused by its orbital or spin angular momentum. In SI units, the Bohr magneton is defined as and in the Gaussian CGS units as where

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">Fine structure</span> Details in the emission spectrum of an atom

In atomic physics, the fine structure describes the splitting of the spectral lines of atoms due to electron spin and relativistic corrections to the non-relativistic Schrödinger equation. It was first measured precisely for the hydrogen atom by Albert A. Michelson and Edward W. Morley in 1887, laying the basis for the theoretical treatment by Arnold Sommerfeld, introducing the fine-structure constant.

<span class="mw-page-title-main">Klein–Nishina formula</span> Electron-photon scattering cross section

In particle physics, the Klein–Nishina formula gives the differential cross section of photons scattered from a single free electron, calculated in the lowest order of quantum electrodynamics. It was first derived in 1928 by Oskar Klein and Yoshio Nishina, constituting one of the first successful applications of the Dirac equation. The formula describes both the Thomson scattering of low energy photons and the Compton scattering of high energy photons, showing that the total cross section and expected deflection angle decrease with increasing photon energy.

<span class="mw-page-title-main">Lamb shift</span> Difference in energy of hydrogenic atom electron states not predicted by the Dirac equation

In physics, the Lamb shift, named after Willis Lamb, is an anomalous difference in energy between two electron orbitals in a hydrogen atom. The difference was not predicted by theory and it cannot be derived from the Dirac equation, which predicts identical energies. Hence the Lamb shift is a deviation from theory seen in the differing energies contained by the 2S1/2 and 2P1/2 orbitals of the hydrogen atom.

<span class="mw-page-title-main">Dirac large numbers hypothesis</span> Hypothesis relating age of the universe to physical constants

The Dirac large numbers hypothesis (LNH) is an observation made by Paul Dirac in 1937 relating ratios of size scales in the Universe to that of force scales. The ratios constitute very large, dimensionless numbers: some 40 orders of magnitude in the present cosmological epoch. According to Dirac's hypothesis, the apparent similarity of these ratios might not be a mere coincidence but instead could imply a cosmology with these unusual features:

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.

Vacuum permittivity, commonly denoted ε0, is the value of the absolute dielectric permittivity of classical vacuum. It may also be referred to as the permittivity of free space, the electric constant, or the distributed capacitance of the vacuum. It is an ideal (baseline) physical constant. Its CODATA value is:

In 1927, a year after the publication of the Schrödinger equation, Hartree formulated what are now known as the Hartree equations for atoms, using the concept of self-consistency that Lindsay had introduced in his study of many electron systems in the context of Bohr theory. Hartree assumed that the nucleus together with the electrons formed a spherically symmetric field. The charge distribution of each electron was the solution of the Schrödinger equation for an electron in a potential , derived from the field. Self-consistency required that the final field, computed from the solutions, was self-consistent with the initial field, and he thus called his method the self-consistent field method.

The isotopic shift is the shift in various forms of spectroscopy that occurs when one nuclear isotope is replaced by another.

In condensed matter physics, Lindhard theory is a method of calculating the effects of electric field screening by electrons in a solid. It is based on quantum mechanics and the random phase approximation. It is named after Danish physicist Jens Lindhard, who first developed the theory in 1954.

An LC circuit can be quantized using the same methods as for the quantum harmonic oscillator. An LC circuit is a variety of resonant circuit, and consists of an inductor, represented by the letter L, and a capacitor, represented by the letter C. When connected together, an electric current can alternate between them at the circuit's resonant frequency:

The Mott–Bethe formula is an approximation used to calculate atomic electron scattering form factors, , from atomic X-ray scattering form factors, . The formula was derived independently by Hans Bethe and Neville Mott both in 1930, and simply follows from applying the first Born approximation for the scattering of electrons via the Coulomb interaction together with the Poisson equation for the charge density of an atom in the Fourier domain. Following the first Born approximation,

In nuclear physics, the astrophysical S-factorS(E) is a rescaling of a nuclear reaction's total cross section σ(E) to account for the Coulomb repulsion between the charged reactants. It determines the rates of nuclear fusion reactions that occur in the cores of stars.

References

[45]

  1. 1 2 3 Hartree, D. R. (1928), "The Wave Mechanics of an Atom with a Non-Coulomb Central Field. Part I. Theory and Methods", Mathematical Proceedings of the Cambridge Philosophical Society, vol. 24, no. 1, Cambridge University Press, pp. 89–110, Bibcode:1928PCPS...24...89H, doi:10.1017/S0305004100011919, S2CID   122077124
  2. McQuarrie, Donald A. (2008). Quantum Chemistry (2nd ed.). New York, NY: University Science Books.
  3. Bethe, Hans A.; Salpeter, Edwin E. (1957). Introduction. Units. Berlin, Heidelberg: Springer Berlin Heidelberg. pp. 2–4. doi:10.1007/978-3-662-12869-5_1. ISBN   978-3-662-12871-8.
  4. 1 2 3 Shull, H.; Hall, G. G. (1959). "Atomic Units". Nature . 184 (4698): 1559. Bibcode:1959Natur.184.1559S. doi:10.1038/1841559a0. S2CID   23692353.
  5. Levine, Ira N. (1991). Quantum chemistry. Pearson advanced chemistry series (4 ed.). Englewood Cliffs, NJ: Prentice-Hall International. ISBN   978-0-205-12770-2.
  6. 1 2 3 McWeeny, R. (May 1973). "Natural Units in Atomic and Molecular Physics". Nature. 243 (5404): 196–198. Bibcode:1973Natur.243..196M. doi:10.1038/243196a0. ISSN   0028-0836. S2CID   4164851.
  7. 1 2 Jerrard, H. G.; McNeill, D. B. (1992). Systems of units. Dordrecht: Springer Netherlands. pp. 3–8. doi:10.1007/978-94-011-2294-8_2. ISBN   978-0-412-46720-2.
  8. International Bureau of Weights and Measures (2006), The International System of Units (SI) (PDF) (8th ed.), p. 125, ISBN   92-822-2213-6, archived (PDF) from the original on 2021-06-04, retrieved 2021-12-16. Note that this information is omitted in the 9th edition.
  9. Paul Quincey; Peter J Mohr; William D Phillips (2019), "Angles are inherently neither length ratios nor dimensionless", Metrologia, 56, arXiv: 1909.08389 , doi:10.1088/1681-7575/ab27d7, In [the Hartree system of atomic] units, me, e, ħ and 1/4πε0 are all set equal to unity. – a reference giving an equivalent set of defining constants.
  10. "9th edition of the SI Brochure". BIPM. 2019. Retrieved 2019-05-20.
  11. "reduced Planck constant". CODATA .
  12. "elementary charge". CODATA .
  13. "electron mass". CODATA .
  14. 1 2 "atomic unit of permittivity". CODATA .
  15. "atomic unit of action". CODATA .
  16. "atomic unit of charge". CODATA .
  17. "atomic unit of mass". CODATA .
  18. "atomic unit of charge density". CODATA .
  19. "atomic unit of current". CODATA .
  20. "atomic unit of charge". CODATA .
  21. "atomic unit of electric dipole moment". CODATA .
  22. "atomic unit of electric quadrupole moment". CODATA .
  23. "atomic unit of electric potential". CODATA .
  24. "atomic unit of electric field". CODATA .
  25. "atomic unit of electric field gradient". CODATA .
  26. "atomic unit of electric polarizability". CODATA .
  27. "atomic unit of 1st hyperpolarizability". CODATA .
  28. "atomic unit of 2nd hyperpolarizability". CODATA .
  29. "atomic unit of magnetic dipole moment". CODATA .
  30. "atomic unit of magnetic flux density". CODATA .
  31. "atomic unit of magnetizability". CODATA .
  32. "atomic unit of action". CODATA .
  33. "atomic unit of energy". CODATA .
  34. "Hartree energy in eV". CODATA .
  35. "atomic unit of force". CODATA .
  36. "atomic unit of length". CODATA .
  37. "atomic unit of mass". CODATA .
  38. "atomic unit of momentum". CODATA .
  39. "atomic unit of time". CODATA .
  40. "atomic unit of velocity". CODATA .
  41. 1 2 Pilar, Frank L. (2001). Elementary Quantum Chemistry. Dover Publications. p. 155. ISBN   978-0-486-41464-5.
  42. Bishop, David M. (1993). Group Theory and Chemistry. Dover Publications. p. 217. ISBN   978-0-486-67355-4.
  43. Drake, Gordon W. F. (2006). Springer Handbook of Atomic, Molecular, and Optical Physics (2nd ed.). Springer. p. 5. ISBN   978-0-387-20802-2.
  44. 1 2 Karplus, Martin; Porter, Richard Needham (1970), Atoms and Molecules: An Introduction for Students of Physical Chemistry, Netherlands: W. A. Benjamin
  45. "CODATA Internationally recommended 2022 values of the Fundamental Physical Constants". NIST Reference on Constants, Units, and Uncertainty. NIST.