Hydrogen atom

Hydrogen-1

General
Name, symbol protium, 1H
Neutrons 0
Protons 1
Nuclide Data
Natural abundance 99.985%
Half-life Stable
Isotope mass 1.007825 u
Spin ½+
Excess energy 7,288.969±0.001 keV
Binding energy 0±keV

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. The most abundant isotope, hydrogen-1, protium, or light hydrogen, contains no neutrons; other isotopes of hydrogen, such as deuterium, contain one or more neutrons. The formulas below are valid for all three isotopes of hydrogen, however slightly different values of the Rydberg constant (correction formula given below) must be used for each hydrogen isotope.

The hydrogen atom has special significance in quantum mechanics and quantum field theory as a simple two-body problem physical system which has yielded many simple analytical solutions in closed-form.

In 1914, Niels Bohr obtained the spectral frequencies of the hydrogen atom after making a number of simplifying assumptions. These assumptions, the cornerstones of the Bohr model, were not fully correct but did yield the correct energy answers. Bohr's results for the frequencies and underlying energy values were confirmed by the full quantum-mechanical analysis which uses the Schrödinger equation, as was shown in 1925–1926. The solution to the Schrödinger equation for hydrogen is analytical. From this, the hydrogen energy levels and thus the frequencies of the hydrogen spectral lines can be calculated. The solution of the Schrödinger equation goes much further than the Bohr model however, because it also yields the shape of the electron's wave function ("orbital") for the various possible quantum-mechanical states, thus explaining the anisotropic character of atomic bonds.

The Schrödinger equation also applies to more complicated atoms and molecules. However, in most such cases the solution is not analytical and either computer calculations are necessary or simplifying assumptions must be made.

Contents

Solution of Schrödinger equation: Overview of results

The solution of the Schrödinger equation (wave equations) for the hydrogen atom uses the fact that the Coulomb potential produced by the nucleus is isotropic (it is radially symmetric in space and only depends on the distance to the nucleus). Although the resulting energy eigenfunctions (the orbitals) are not necessarily isotropic themselves, their dependence on the angular coordinates follows completely generally from this isotropy of the underlying potential: The eigenstates of the Hamiltonian (that is, the energy eigenstates) can be chosen as simultaneous eigenstates of the angular momentum operator. This corresponds to the fact that angular momentum is conserved in the orbital motion of the electron around the nucleus. Therefore, the energy eigenstates may be classified by two angular momentum quantum numbers, and m (both are integers). The angular momentum quantum number = 0, 1, 2, ... determines the magnitude of the angular momentum. The magnetic quantum number m = −, ..., + determines the projection of the angular momentum on the (arbitrarily chosen) z-axis.

In addition to mathematical expressions for total angular momentum and angular momentum projection of wavefunctions, an expression for the radial dependence of the wave functions must be found. It is only here that the details of the 1/r Coulomb potential enter (leading to Laguerre polynomials in r). This leads to a third quantum number, the principal quantum number n = 1, 2, 3, .... The principal quantum number in hydrogen is related to atom's total energy.

Note that the maximum value of the angular momentum quantum number is limited by the principal quantum number: it can run only up to n − 1, i.e. = 0, 1, ..., n − 1.

Due to angular momentum conservation, states of the same but different m have the same energy (this holds for all problems with rotational symmetry). In addition, for the hydrogen atom, states of the same n but different are also degenerate (i.e. they have the same energy). However, this is a specific property of hydrogen and is no longer true for more complicated atoms which have a (effective) potential differing from the form 1/r (due to the presence of the inner electrons shielding the nucleus potential).

Taking into account the spin of the electron adds a last quantum number, the projection of the electron's spin angular momentum along the z-axis, which can take on two values. Therefore, any eigenstate of the electron in the hydrogen atom is described fully by four quantum numbers. According to the usual rules of quantum mechanics, the actual state of the electron may be any superposition of these states. This explains also why the choice of z-axis for the directional quantization of the angular momentum vector is immaterial: an orbital of given and m′ obtained for another preferred axis z′ can always be represented as a suitable superposition of the various states of different m (but same l) that have been obtained for z.

Alternatives to the Schrödinger Theory

In the language of Heisenberg's Matrix Mechanics, the hydrogen atom was first solved by Wolfgang Pauli[1] using a rotational symmetry in four dimension [O(4)-symmetry] generated by the angular momentum and the Laplace–Runge–Lenz vector. By extending the symmetry group O(4) to the dynamical group O(4,2), the entire spectrum and all transitions were embedded in a single irreducible group representation.[2]

In 1979 the (non relativistic) hydrogen atom was solved for the first time within Feynman's path integral formulation of quantum mechanics.[3][4] This work greatly extended the range of applicability of Feynman's method.

Mathematical summary of eigenstates of hydrogen atom

Energy levels

The energy levels of hydrogen, including fine structure, are given by

E_{jn} = \frac{-13.6 \ \mathrm{eV}}{n^2} \left[1 %2B \frac{\alpha^2}{n^2}\left(\frac{n}{j%2B\frac{1}{2}} - \frac{3}{4} \right) \right] ,

where α is the fine-structure constant and j is a number which is the total angular momentum eigenvalue; that is, ± 1/2 depending on the direction of the electron spin. The quantity in square brackets arises from relativistic (spin-orbit) coupling interactions (as further described below in the section entitled "Features going beyond the Schrödinger solution").

The value of 13.6 eV is called the Rydberg constant and can be found from the Bohr model, and is given by

-13.6 \ \mathrm{eV} = -\frac{m_e q_e^4}{8 h^2 \varepsilon_{0}^2},

where me is the mass of the electron, qe is the charge of the electron, h is the Planck constant, and ε0 is the vacuum permittivity.

The Rydberg constant is connected to the fine-structure constant by the relation

-13.6 \ \mathrm{eV} = -\frac{m_e c^2\alpha^2}{2}  = -\frac{0.51\mathrm{MeV}}{2 \cdot 137^2} .

This constant is often used in atomic physics in the form of the Rydberg unit of energy:

1 \ \mathrm{Ry} \equiv h c R_\infty = 13.605\;692\;53(30) \ \mathrm{eV}.[5]

The exact value of the Rydberg constant above assumes that the nucleus is infinitely massive with respect to the electron. For hydrogen-1, hydrogen-2 (deuterium), and hydrogen-3 (tritium) the constant must be slight modified to use the reduced mass of the system, rather than simply the mass of the electron. However, since the nucleus is much heavier than the electron, the values are nearly the same. The Rydberg constant RM for a hydrogen atom (one electron), R is given by :R_M = \frac{R_\infty}{1%2Bm_{\mathrm{e}}/M}, where m_{\mathrm{e}} is the rest mass of the electron, and M is the mass of the atomic nucleus. For hydrogen-1, the quantity m_{\mathrm{e}/M}, is about 1/1836, reflecting the ratio electron to proton mass. For deuterium and tritium, the numbers are about 1/3670 and 1/5497 respectively. These figures, when added to 1 in the denominator, represent very small corrections in the value of R, and thus only small corrections to all energy levels in corresponding hydrogen isotopes.

Wavefunction

The normalized position wavefunctions, given in spherical coordinates are:[6]

 \psi_{n\ell m}(r,\vartheta,\varphi) = \sqrt {{\left (  \frac{2}{n a_0} \right )}^3\frac{(n-\ell-1)!}{2n[(n%2B\ell)!]^3} } e^{- \rho / 2} \rho^{\ell} L_{n-\ell-1}^{2\ell%2B1}(\rho) \cdot Y_{\ell}^{m}(\vartheta, \varphi )

where:

 \rho = {2r \over {na_0}} ,
 a_0 is the Bohr radius,
 L_{n-\ell-1}^{2\ell%2B1}(\rho) are the generalized Laguerre polynomials of degree n − 1, and
 Y_{\ell}^{m}(\vartheta, \varphi ) \, is a spherical harmonic function of degree and order m. Note that the generalized Laguerre polynomials are defined differently by different authors. The usage here is consistent with the definitions used by Messiah[7] and Griffiths.[8] But not Mathematica where it differs by a factor of (n%2Bl)! [9] In other places,[10] the Generalized Laguerre polynomial appearing in the Hydrogen wave function may be  L_{n%2B\ell}^{2\ell%2B1}(\rho).

The quantum numbers can take the following values:

 n=1,2,3,\ldots
\ell=0,1,2,\ldots,n-1
m=-\ell,\ldots,\ell.

Additionally, these wavefunctions are normalized (i.e., the integral of their modulus square equals 1) and orthogonal:

\int_0^{%2B\infty} r^2 dr\int_0^{\pi} \sin \vartheta d\vartheta \int_0^{2 \pi} d\varphi\; \psi^*_{n\ell m}(r,\vartheta,\varphi)\psi_{n'\ell'm'}(r,\vartheta,\varphi)=\langle n,\ell, m | n', \ell', m' \rangle = \delta_{nn'} \delta_{\ell\ell'} \delta_{mm'},

where | n, \ell, m \rangle is the representation of the wavefunction  \psi_{n\ell m} in Dirac notation, and  \delta is the Kronecker delta function.[11]

Angular momentum

The eigenvalues for Angular momentum operator:

 L^2\, | n, \ell, m\rangle = {\hbar}^2 \ell(\ell%2B1)\, | n, \ell, m \rang
 L_z\, | n, \ell, m \rang = \hbar m \,| n, \ell, m \rang.

Visualizing the hydrogen electron orbitals

The image to the right shows the first few hydrogen atom orbitals (energy eigenfunctions). These are cross-sections of the probability density that are color-coded (black represents zero density and white represents the highest density). The angular momentum (orbital) quantum number is denoted in each column, using the usual spectroscopic letter code (s means  = 0, p means  = 1, d means  = 2). The main (principal) quantum number n (= 1, 2, 3, ...) is marked to the right of each row. For all pictures the magnetic quantum number m has been set to 0, and the cross-sectional plane is the xz-plane (z is the vertical axis). The probability density in three-dimensional space is obtained by rotating the one shown here around the z-axis.

The "ground state", i.e. the state of lowest energy, in which the electron is usually found, is the first one, the 1s state (principal quantum level n = 1, = 0).

An image with more orbitals is also available (up to higher numbers n and ).

Black lines occur in each but the first orbital: these are the nodes of the wavefunction, i.e. where the probability density is zero. (More precisely, the nodes are Spherical harmonics that appear as a result of solving Schrödinger's equation in polar coordinates.)

The quantum numbers determine the layout of these nodes.[12] There are:

Features going beyond the Schrödinger solution

There are several important effects that are neglected by the Schrödinger equation and which are responsible for certain small but measurable deviations of the real spectral lines from the predicted ones:

Both of these features (and more) are incorporated in the relativistic Dirac equation, with predictions that come still closer to experiment. Again the Dirac equation may be solved analytically in the special case of a two-body system, such as the hydrogen atom. The resulting solution quantum states now must be classified by the total angular momentum number j (arising through the coupling between electron spin and orbital angular momentum). States of the same j and the same n are still degenerate.

For these developments, it was essential that the solution of the Dirac equation for the hydrogen atom could be worked out exactly, such that any experimentally observed deviation had to be taken seriously as a signal of failure of the theory.

Due to the high precision of the theory also very high precision for the experiments is needed, which utilize a frequency comb.

Hydrogen ion

Hydrogen is not found without its electron in ordinary chemistry (room temperatures and pressures), as ionized hydrogen is highly chemically reactive. When ionized hydrogen is written as "H+" as in the solvation of classical acids such hydrochloric acid, the hydronium ion, H3O+, is meant, not a literal ionized single hydrogen atom. In that case, the acid transfers the proton to H2O to form H3O+.

Ionized hydrogen without its electron, or free protons, are common in the interstellar medium, and solar wind.

See also

References

  1. ^ Pauli, W (1926). "Über das Wasserstoffspektrum vom Standpunkt der neuen Quantenmechanik". Zeitschrift für Physik 36: 336–363. Bibcode 1926ZPhy...36..336P. doi:10.1007/BF01450175. 
  2. ^ Kleinert H. (1968). "Group Dynamics of the Hydrogen Atom". Lectures in Theoretical Physics, edited by W.E. Brittin and A.O. Barut, Gordon and Breach, N.Y. 1968: 427–482. http://www.physik.fu-berlin.de/~kleinert/kleiner_re4/4.pdf. 
  3. ^ Duru I.H., Kleinert H. (1979). "Solution of the path integral for the H-atom". Physics Letters B 84 (2): 185–188. Bibcode 1979PhLB...84..185D. doi:10.1016/0370-2693(79)90280-6. http://www.physik.fu-berlin.de/~kleinert/kleiner_re65/65.pdf. 
  4. ^ Duru I.H., Kleinert H. (1982). "Quantum Mechanics of H-Atom from Path Integrals". Fortschr. Phys 30 (2): 401–435. doi:10.1002/prop.19820300802. http://www.physik.fu-berlin.de/~kleinert/kleiner_re83/83.pdf. 
  5. ^ P.J. Mohr, B.N. Taylor, and D.B. Newell (2011), "The 2010 CODATA Recommended Values of the Fundamental Physical Constants" (Web Version 6.0). This database was developed by J. Baker, M. Douma, and S. Kotochigova. Available: http://physics.nist.gov/constants. National Institute of Standards and Technology, Gaithersburg, MD 20899. Link to R, Link to hcR
  6. ^ David Griffiths (2008). Introduction to elementary particles. Wiley-VCH. pp. 162–. ISBN 9783527406012. http://books.google.com/books?id=w9Dz56myXm8C&pg=PA162. Retrieved 27 June 2011. 
  7. ^ Messiah, Albert (1999). Quantum Mechanics. New York: Dover. pp. 1136. ISBN 0-486-40924-4. 
  8. ^ Griffiths, David (1995). Introduction to Quantum Mechanics. New Jersey: Pearson Education, Inc.. pp. 152. ISBN 0131118927. 
  9. ^ LaguerreL. Wolfram Mathematica page
  10. ^ Condon and Shortley (1963). The Theory of Atomic Spectra. London: Cambridge. pp. 441. 
  11. ^ Introduction to Quantum Mechanics, Griffiths 4.89
  12. ^ Summary of atomic quantum numbers. Lecture notes. 28 July 2006

Books

External links

Lighter:
(none, lightest possible)
Hydrogen atom is an
isotope of hydrogen
Heavier:
hydrogen-2
Decay product of:
neutronium-1
helium-2
Decay chain
of Hydrogen atom
Decays to:
Stable