Hostname: page-component-76fb5796d-dfsvx Total loading time: 0 Render date: 2024-04-26T18:20:38.763Z Has data issue: false hasContentIssue false

H3+: the initiator of interstellar chemistry

Published online by Cambridge University Press:  07 August 2008

Mats Larsson
Affiliation:
Department of Physics, AlbaNova University Center, Stockholm University, SE-10691 Stockholm, Sweden e-mail: mats.larsson@physto.se

Abstract

Second only to H2, protonated molecular hydrogen, H3+, is the most abundantly produced interstellar molecule. Owing to its high reactivity and acidity, it plays the pivotal role in initiating interstellar chemical reactions, something which also reduces its steady-state concentration. Interstellar H3+ is not only destroyed in chemical reactions but also in dissociative recombination with electrons. The rate constant and mechanism of recombination have long been controversial, but great advances have been made during recent years, with the important consequence that the cosmic ray ionization rate in diffuse clouds is now believed to be higher by an order of magnitude than previously assumed.

Type
Research Article
Copyright
Copyright © 2008 Cambridge University Press

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Adams, N.G. & Smith, D. (1987). Recent advances in the studies of reaction rates relevant to interstellar chemistry. In Astrochemistry, IAU Symp. 120, eds Vardya, M.S. & Tarafdar, S.P., pp. 118. Reidel, Dordrecht.Google Scholar
Adams, N.G., Smith, D. & Alge, E. (1984). Measurements of dissociative recombination rate coefficients of H3+, HCO+, N2H+, and CH5+ at 95 K and 300 K using the FALP apparatus. J. Chem. Phys. 81, 17781784.CrossRefGoogle Scholar
Amano, T. (1988). Is the dissociative recombination of H3+ really slow? A new spectroscopic measurement of the rate constant. Astrophys. J. Lett. 329, L121L124.Google Scholar
Amano, T. (1990). The dissociative recombination rate coefficient of H3+, HN2+, and HCO+. J. Chem. Phys. 92, 64926501.Google Scholar
Dalgarno, A. (2006). The galactic cosmic ray ionization rate. Proc. Natl. Acad. Sci. USA 103, 1226912273.CrossRefGoogle ScholarPubMed
Fonseca dos Santos, S., Kokoouline, V. & Greene, C.H. (2007). Dissociative recombination of H3+ in the ground and excited vibrational states. J. Chem. Phys. 127, 124309.Google Scholar
Geballe, T.R. & Oka, T. (1996). Detection of H3+ in interstellar space. Nature 384, 334335.CrossRefGoogle ScholarPubMed
Glosik, J., Plašil, R., Poterya, V., Kudrna, P. & Tichý, M. (2000). The recombination of H3+ ions with electrons: dependence of partial pressure of H2. Chem. Phys. Lett. 331, 209214.CrossRefGoogle Scholar
Glosik, J. et al. (2008). Recombination of H3+ ions in the afterglow of a He-Ar-H2 plasma. Phys. Rev. Lett. (unpublished).Google Scholar
Greene, C.H., Kokoouline, V. & Esry, B.D. (2003). Importance of Jahn–Teller coupling in the dissociative recombination of H3+ by low energy electrons. In Dissociative Recombination of Molecular Ions with Electrons, ed. Guberman, S.L., pp. 221233. Kluwer Academic/Plenum, New York.Google Scholar
Guberman, S.L. (ed.) (2003). Dissociative Recombination of Molecular Ions with Electrons. Kluwer Academic/Plenum, New York.Google Scholar
Herbst, E. & Klemperer, W. (1973). The formation and depletion of molecules in dense interstellar clouds. Astrophys. J. 185, 505533.Google Scholar
Herbst, E., Miller, S., Oka, T. & Watson, J.K.G.(eds.) (2000). Astronomy, physics and chemistry of H3+. Phil. Trans. R. Soc. Lond. A 358, 23592559.Google Scholar
Indriolo, N., Geballe, T.R., Oka, T. & McCall, B.J. (2007). H3+ in diffuse interstellar clouds: a tracer for the cosmic-ray ionization rate. Astrophys. J. 671, 17361747.Google Scholar
Jensen, M.J., Pedersen, H.B., Safvan, C.P., Seiersen, K., Urbain, X. & Andersen, L.H. (2001). Dissociative recombination and excitation of H3+. Phys. Rev. A 63, 052701-15.Google Scholar
Kokoouline, V. & Greene, C.H. (2003a). Theory of dissociative recombination of D 3h triatomic ions applied to H3+. Phys. Rev. Lett. 90, 13201-14.Google Scholar
Kokoouline, V. & Greene, C.H. (2003b). Unified theoretical treatment of dissociative recombination of D 3h triatomic ions: Applications to H3+ and D3+. Phys. Rev. A 68, 012703-123.CrossRefGoogle Scholar
Kokoouline, V., Greene, C.H. & Esry, B.D. (2001). Mechanism for the destruction of H3+ ions by electron impact. Nature 412, 891894.CrossRefGoogle ScholarPubMed
Kreckel, H., Krohn, S., Lammich, L. et al. (2002). Vibrational and rotational cooling of H3+. Phys. Rev. A 66, 052509-111.Google Scholar
Kreckel, H. et al. (2005). High resolution dissociative recombination of cold H3+ and first evidence for nuclear spin effects. Phys. Rev. Lett. 95, 263201-14.Google Scholar
Larsson, M. (2000). Experimental studies of the dissociative recombination of H3+, Phil. Trans. R. Soc. Lond. A 358, 24332444.Google Scholar
Larsson, M. & Orel, A.E. (2008). Dissociative Recombination of Molecular Ions. Cambridge University Press, Cambridge.Google Scholar
Larsson, M. et al. (1993). Direct high-energy neutral-channel dissociative recombination of cold H3+ in an ion storage ring. Phys. Rev. Lett. 70, 430433.Google Scholar
Larsson, M. et al. (2003). Studies of dissociative recombination in CRYRING. In Dissociative Recombination of Molecular Ions with Electrons, ed. Guberman, S.L., pp. 8794. Kluwer Academic/Plenum, New YorkGoogle Scholar
Leu, M. T., Biondi, M. A. & Johnsen, R. (1973). Measurements of recombination of electrons with H3+ and H5+, Phys. Rev. A 8, 413419.Google Scholar
McCall, B.J. (2001). Spectroscopy of H3+ in laboratory and astrophysical plasmas. PhD Thesis, The University of Chicago.Google Scholar
McCall, B.J., Geballe, T.R., Hinkle, K.H. & Oka, T. (1998). Detection of H3+ in the diffuse interstellar medium toward Cygnus OB2 no. 12. Science 279, 19101913.Google Scholar
McCall, B.J. et al. (2003). An enhanced cosmic-ray flux towards ζ Persei inferred from a laboratory study of the H3+−e recombination rate. Nature 422, 500502.Google Scholar
McCall, B.J. et al. (2004). Dissociative recombination of rotationally cold H3+. Phys. Rev. A 70, 052716-113.CrossRefGoogle Scholar
Oka, T. (1980). Observation of the infrared spectrum of H3+. Phys. Rev. Lett. 45, 531534.Google Scholar
Oka, T. (2003). Help!!! Theory for H3+ recombination badly needed. In Dissociative Recombination of Mole-cular Ions with Electrons, ed. Guberman, S.L., pp. 209220. Kluwer Academic/Plenum, New York.Google Scholar
Oka, T. (2006). Interstellar H3+. Proc. Natl. Acad. Sci. USA 103, 1223512242.CrossRefGoogle ScholarPubMed
Schneider, I.F., Suzor-Weiner, A. & Orel, A.E. (2000). Channel mixing effects in the dissociative recombination of H3+ with slow electrons. Phys. Rev. Lett. 85, 37853788.Google Scholar
Sundström, G. et al. (1994). Destruction rate of H3+ by low energy electrons measured in a storage-ring experiment. Science 263, 785787.Google Scholar
Thomson, J.J. (1911). Rays of positive electricity. Phil. Mag. 21, 225249.CrossRefGoogle Scholar
Van Dishoeck, E.F. & Black, J.H. (1986). Comprehensive models of diffuse interstellar clouds: Physical conditions and molecular abundances. Astrophys. J. Suppl. Ser. 62, 109145.Google Scholar
Watson, W.D. (1973). Rate of formation of interstellar molecules by ion-molecule reactions. Astrophys. J. 183, L17L20.Google Scholar