Dihydrogen cation

Dihydrogen cation

The hydrogen molecular ion, dihydrogen cation, or H2+, is the simplest molecular ion. It is composed of two positively-charged protons and one negatively-charged electron, and can be formed from ionization of a neutral hydrogen molecule. It is of great historical and theoretical interest because, having only one electron, the Schrödinger equation for the system can be solved in a relatively straightforward way due to the lack of electron–electron repulsion (electron correlation). The analytical solutions for the energy eigenvalues are a generalization of the Lambert W function (see Lambert W function and references therein for more details on this function).[1] Thus, the case of clamped nuclei can be completely done analytically using a Computer algebra system. Consequently, it is included as an example in most quantum chemistry textbooks.

The first successful quantum mechanical treatment of H2+ was published by the Danish physicist Øyvind Burrau in 1927,[2] just one year after the publication of wave mechanics by Erwin Schrödinger. Earlier attempts using the old quantum theory had been published in 1922 by Karel Niessen[3] and Wolfgang Pauli,[4] and in 1925 by Harold Urey.[5] In 1928, Linus Pauling published a review putting together the work of Burrau with the work of Walter Heitler and Fritz London on the hydrogen molecule.[6]

Bonding in H2+ can be described as a covalent one-electron bond, which has a formal bond order of one half.[7]

The ion is commonly formed in molecular clouds in space, and is important in the chemistry of the interstellar medium.

Contents

Quantum mechanical treatment, symmetries, and asymptotics

Hydrogen Molecular Ion H2+ with clamped nuclei A and B, internuclear distance R and plane of symmetry M.

The simplest electronic Schrödinger wave equation for the hydrogen molecular ion  H_2^{+} is modeled with two fixed nuclear centers, labeled A and B, and one electron. It can be written as


\left( -\frac{\hbar^2}{2m} \nabla^2 + V \right) \psi = E \psi ~,

where V is the electron-nuclear Coulomb potential energy function:


V =  - \frac{e^{2}}{4 \pi \varepsilon_0 } \left( \frac{1}{r_a} + \frac{1}{r_b}   \right)

and E is the (electronic) energy of a given quantum mechanical state (eigenstate), with the electronic state function  \psi=\psi(\mathbf{r}) depending on the spatial coordinates of the electron. An additive term 1 / R, which is constant for fixed inter-nuclear distance R, has been omitted from the potential V, since it merely shifts the eigenvalue. The distances between the electron and the nuclei are denoted r_a^{} and r_b^{}. In atomic units (\hbar=m=e=4 \pi\varepsilon_0 =1) the wave equation is

\left( {} - \frac{1}{2} \nabla^2 + V \right) \psi = E \psi    \qquad \mbox{with} \qquad  V = {} - \frac{1}{r_a^{}} - \frac{1}{r_b^{}} \; .

We can choose the midpoint between the nuclei as the origin of coordinates. It follows from general symmetry principles that the wave functions can be characterized by their symmetry behavior with respect to space inversion (r  \to -r). There are wave functions :\psi_{+}(\mathbf{r}), which are symmetric with respect to space inversion, and there are wave functions :\psi_{-}(\mathbf{r}), which are anti-symmetric under this symmetry operation:  \psi_{\pm}(-{\mathbf{r}}) = {} \pm \psi_{\pm}({\mathbf r}) \; . We note that the permutation (exchange) of the nuclei has a similar effect on the electronic wave function. We only mention that for a many-electron system proper behavior of ψ with respect to the permutational symmetry of the electrons (Pauli exclusion principle) must be guaranteed, in addition to those symmetries just discussed above. Now the Schrödinger equations for these symmetry-adapted wave functions are

 \begin{align}
\left( -\frac{1}{2} \nabla^2 + V \right) \psi_{+} = E_{+} \psi_{+} \\  
\left( -\frac{1}{2} \nabla^2 + V \right) \psi_{-} = E_{-} \psi_{-}
\end{align}

The ground state (the lowest discrete state) of  H_{2}^{+} is the  {\rm X}_{}^{2}\Sigma_{\rm g}^{+} state [8] with the corresponding wave function ψ + denoted as 1s \sigma_{\rm g}^{}. There is also the first excited  {\rm A}_{}^{2}\Sigma_{\rm u}^{+} state, with its ψ labeled as  {\rm 2p}\sigma_{\rm u}^{}. (The suffixes g and u are from the German gerade and ungerade) occurring here denote just the symmetry behavior under space inversion. Their use is standard practice for the designation of electronic states of diatomic molecules, whereas for atomic states the terms even and odd are used.

Energies (E) of the lowest discrete states of the Hydrogen Molecular Ion H_2^{+} as a function of inter-nuclear distance (R) in atomic units. See text for details.

Asymptotically, the (total) eigenenergies E_{\pm} for these two lowest lying states have the same asymptotic expansion in inverse powers of the inter-nuclear distance R [9]:

 
E_{\pm} = {} - \frac{1}{2} - \frac{9}{4 R^4} + O(R^{-6})  + \cdots

The actual difference between these two energies is called the exchange energy splitting and is given by [10]:

 
\Delta E = E_{-} - E_{+} = \frac{4}{e} \, R \, e^{-R}      \left[ \, 1 + \frac{1}{2R} + O(R^{-2}) \, \right]

which exponentially vanishes as the inter-nuclear distance R gets greater. The lead term  {\textstyle \frac{4}{e}} R e^{-R} was first obtained by the Holstein–Herring method. Similarly, asympotic expansions in powers of 1/R have been obtained to high order by Cizek et al. for the lowest ten discrete states of the hydrogen molecular ion (clamped nuclei case). For general diatomic and polyatomic molecular systems, the exchange energy is thus very elusive to calculate at large inter-nuclear distances but is nonetheless needed for long-range interactions including studies related to magnetism and charge exchange effects. These are of particular importance in stellar and atmospheric physics.

The energies for the lowest discrete states are shown in the graph above. These can be obtained to within arbitrary accuracy using computer algebra from the generalized Lambert W function (see eq. (3) in that site and the reference of Scott, Aubert-Frécon, and Grotendorst) but were obtained initially by numerical means to within double precision by the most precise program available, namely ODKIL.[11] The red full lines are  {\rm {}}_{}^{2}\Sigma_{\rm g}^{+} states. The green dashed lines are  {\rm {}}_{}^{2}\Sigma_{\rm u}^{+} states. The blue dashed line is a  {\rm {}}_{}^{2}\Pi_{\rm u} state and the pink dotted line is a  {\rm {}}_{}^{2}\Pi_{\rm g} state. Note that although the generalized Lambert W function eigenvalue solutions supersede these asymptotic expansions, in practice, they are most useful near the bond length. These solutions are possible because the partial differential equation of the wave equation here separates into two coupled ordinary differential equations using prolate spheroidal coordinates.

Formation

The dihydrogen ion is formed in nature by the interaction of cosmic rays and the hydrogen molecule. An electron is knocked off leaving the cation behind.[12]

H2 + cosmic ray → H2+ + e- + cosmic ray.

Cosmic ray particles have enough energy to ionize many molecules before coming to a stop.

In nature the ion is destroyed by reacting with other hydrogen molecules:

H2+ + H2 → H3+ + H.

The ionization energy of the hydrogen molecule is 15.603 eV. The dissociation energy of the ion is 1.8 eV. High speed electrons also cause ionization of hydrogen molecules. The peak cross section for ionization for high speed protons is 70000 eV with a cross section of 2.5x10−16 cm2. A cosmic ray proton at lower energy can also strip an electron off a neutral hydrogen molecule to form a neutral hydrogen atom, with a peak cross section at around 8000 eV of 8x10−16 cm2.[13]

An artificial plasma discharge cell can also produce the ion.

See also

References

  1. ^ Scott, T. C.; Aubert-Frécon, M.; Grotendorst, J. (2006). "New Approach for the Electronic Energies of the Hydrogen Molecular Ion". Chem. Phys. 324 (2–3): 323–338. arXiv:physics/0607081. doi:10.1016/j.chemphys.2005.10.031. 
  2. ^ Burrau Ø (1927). "Berechnung des Energiewertes des Wasserstoffmolekel-Ions (H2+) im Normalzustand." (in German). Danske Vidensk. Selskab. Math.-fys. Meddel. M 7:14: 1–18. http://www.royalacademy.dk/CatalogEntry.asp?id=862. 
    Burrau Ø (1927). "The calculation of the Energy value of Hydrogen molecule ions (H2+) in their normal position" (in German) (PDF). Naturwissenschaften 15 (1): 16–7. doi:10.1007/BF01504875. http://www.springerlink.com/content/h60148l4717uv805/fulltext.pdf. 
  3. ^ Karel F. Niessen Zur Quantentheorie des Wasserstoffmolekülions, doctoral dissertation, University of Utrecht, Utrecht: I. Van Druten (1922) as cited in Mehra, Volume 5, Part 2, 2001, p. 932.
  4. ^ Pauli W (1922). "Über das Modell des Wasserstoffmolekülions". Ann. D. Phys. 373 (11): 177–240. doi:10.1002/andp.19223731101.  Extended doctoral dissertation; received 4 March 1922, published in issue No. 11 of 3 August 1922.
  5. ^ Urey HC (October 1925). "The Structure of the Hydrogen Molecule Ion". Proc. Natl. Acad. Sci. U.S.A. 11 (10): 618–21. doi:10.1073/pnas.11.10.618. PMC 1086173. PMID 16587051. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1086173. 
  6. ^ Pauling, L. (1928). "The Application of the Quantum Mechanics to the Structure of the Hydrogen Molecule and Hydrogen Molecule-Ion and to Related Problems". Chemical Reviews 5 (2): 173–213. doi:10.1021/cr60018a003. 
  7. ^ Clark R. Landis; Frank Weinhold (2005). Valency and bonding: a natural bond orbital donor-acceptor perspective. Cambridge, UK: Cambridge University Press. pp. 96–100. ISBN 0-521-83128-8. 
  8. ^ Huber, K.-P.; Herzberg, G. (1979). Molecular Spectra and Molecular Structure. IV. Constants of Diatomic Molecules. New York: Van Nostrand Reinhold. 
  9. ^ Čížek, J.; Damburg, R. J.; Graffi, S.; Grecchi, V.; Harrel II, E. M.; Harris, J. G.; Nakai, S.; Paldus, J. et al. (1986). "1/R expansion for H2+: Calculation of exponentially small terms and asymptotics". Phys. Rev. A 33: 12–54. doi:10.1103/PhysRevA.33.12. 
  10. ^ Scott, T. C.; Dalgarno, A.; Morgan III, J. D. (1991). "Exchange Energy of H2+ Calculated from Polarization Perturbation Theory and the Holstein-Herring Method". Phys. Rev. Lett. 67 (11): 1419–1422. doi:10.1103/PhysRevLett.67.1419. 
  11. ^ Hadinger, G.; Aubert-Frécon, M.; Hadinger, G. (1989). "The Killingbeck method for the one-electron two-centre problem". J. Phys. B 22 (5): 697–712. doi:10.1088/0953-4075/22/5/003. 
  12. ^ Herbst, E. (2000). "The Astrochemistry of H3+". Phil. Trans. R. Soc. Lond. A. 358 (1774): 2523–2534. doi:10.1098/rsta.2000.0665. 
  13. ^ Padovani, Marco; Galli, Daniele; Glassgold, Alfred E. (2009). "Cosmic-ray ionization of molecular clouds". Preprint. arXiv:0904.4149. 

Wikimedia Foundation. 2010.

Игры ⚽ Нужен реферат?

Look at other dictionaries:

  • Dihydrogen complex — Formation and equilibrium structures of metal dihydrogen and dihydride complexes (L = ligand). Dihydrogen complexes are coordination complexes containing intact H2 as a ligand.[1] The prototypical complex is W(CO)3(PCy3)2(H2). This class of… …   Wikipedia

  • Dihydrogen bond — In chemistry, a dihydrogen bond is a kind of hydrogen bond, an interaction between a metal hydride bond and an OH or NH group or another proton donor. The first example of this phenomenon is credited to Brown and Heseltine.[1] They observed… …   Wikipedia

  • Chemical bond — A chemical bond is an attraction between atoms that allows the formation of chemical substances that contain two or more atoms. The bond is caused by the electromagnetic force attraction between opposite charges, either between electrons and… …   Wikipedia

  • Water (molecule) — Chembox new Name = Water (H2O) ImageFileL1 = Water 2D labelled.png ImageSizeL1 = 120px ImageNameL1 = The water molecule has this basic geometric structure ImageFileR1 = Water molecule 3D.svg ImageSizeR1 = 100px ImageNameR1 = Space filling model… …   Wikipedia

  • Hydrogen — This article is about the chemistry of hydrogen. For the physics of atomic hydrogen, see Hydrogen atom. For other meanings, see Hydrogen (disambiguation). ← hydrogen → helium …   Wikipedia

  • Molecular orbital diagram — A molecular orbital diagram, or MO diagram for short, is a qualitative descriptive tool explaining chemical bonding in molecules in terms of molecular orbital theory in general and the Linear combination of atomic orbitals molecular orbital… …   Wikipedia

  • IUPAC nomenclature of inorganic chemistry — For the current Red Book version, see IUPAC nomenclature of inorganic chemistry 2005. The IUPAC nomenclature of inorganic chemistry is a systematic method of naming inorganic chemical compounds, as recommended by the International Union of Pure… …   Wikipedia

  • Intercalation (chemistry) — Intercalation induces structural distortions. Left: unchanged DNA strand. Right: DNA strand intercalated at three locations (red areas). In chemistry, intercalation is the reversible inclusion of a molecule (or group) between two other molecules… …   Wikipedia

  • oxyacid — /ok see as id/, n. Chem. an inorganic acid containing oxygen. Also called oxygen acid. [1830 40; OXY 2 + ACID] * * * ▪ chemical compound Introduction       any oxygen containing acid. Most covalent nonmetallic oxides (oxide) react with water to… …   Universalium

  • Properties of water — H2O and HOH redirect here. For other uses, see H2O (disambiguation) and HOH (disambiguation). This article is about the physical and chemical properties of pure water. For general discussion and its distribution and importance in life, see Water …   Wikipedia

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”