Home Search Collections Journals About Contact us My IOPscience Electron and photon dissociation cross sections of the H2 singlet ungerade continua This article has been downloaded from IOPscience. Please scroll down to see the full text article. 2012 J. Phys. B: At. Mol. Opt. Phys. 45 015201 (http://iopscience.iop.org/0953-4075/45/1/015201) View the table of contents for this issue, or go to the journal homepage for more Download details: IP Address: 128.138.131.243 The article was downloaded on 02/02/2012 at 20:50 Please note that terms and conditions apply. IOP PUBLISHING JOURNAL OF PHYSICS B: ATOMIC, MOLECULAR AND OPTICAL PHYSICS doi:10.1088/0953-4075/45/1/015201 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 (15pp) Electron and photon dissociation cross sections of the H2 singlet ungerade continua Xianming Liu1 , Donald E Shemansky1 , Paul V Johnson2 , Charles P Malone2,3 , Murtadha A Khakoo3 and Isik Kanik2 1 Planetary and Space Science Division, Space Environment Technologies, 1676 Palisades Drive, Pacific Palisades, CA 90272, USA 2 Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena, CA 91109, USA 3 Department of Physics, California State University, Fullerton, CA 92834, USA E-mail: xliu@spacenvironment.net Received 17 October 2011, in final form 10 November 2011 Published 9 December 2011 Online at stacks.iop.org/JPhysB/45/015201 Abstract Photodissociation cross sections and oscillator strengths for H2 from the X 1 g+ (vi , Ji ) levels to the continuum levels of the B 1 u+ , C 1 u , B B̄ 1 u+ , D 1 u and 5pσ 1 u+ states have been calculated. The (vi , Ji ) state-specific electron impact dissociation cross sections to the continuum levels of these states have been obtained for the first time over a wide energy range using calculated continuum oscillator strengths along with previously published excitation functions of the Lyman and Werner bands. Estimated cross sections to the higher (n 5) npσ 1 u+ and npπ 1 u continua are also provided. Both photon and electron impact excitation cross sections show strong dependences on the initial (vi , Ji ) quantum numbers. Thermally averaged electron impact cross sections of all singlet ungerade states increase monotonically with temperature. While excitation to the B 1 u+ continuum is the dominant dissociation channel at room temperature, the C 1 u and B 1 u+ continua become more important at high temperature (>5000 K). This work, along with the previous calculation of the B 1 u+ and D 1 u states by Liu et al (2009a J. Phys. B: At. Mol. Opt. Phys. 42 185203), provides the complete electron impact dissociation cross section of H2 through the singlet ungerade continua. Electron dissociation cross sections of the singlet ungerade continua are provided for the purpose of modelling atmospheric heating, analysis of occultation measurements and the hot atomic hydrogen plume observed at Saturn. (Some figures in this article are in colour only in the electronic version) 1. Introduction B 1 u+ and C 1 u continua are expected to increase rapidly with increasing vibrational quantum number of the X 1 g+ state. These calculated cross sections, along with the estimated values for higher Rydberg states, give a reliable estimate of the total dissociation cross sections of the singlet ungerade continua. In H2 -dominated atmospheres, photon and electron impact dissociations of H2 are the primary mechanisms for the production of kinetically hot atomic hydrogen fragments. These processes efficiently convert electron and photon energy into heat. The dissociation cross section of H2 is vital for This paper presents (vi , Ji ) state-specific and thermally averaged electron and photon dissociation cross sections of the B 1 u+ , C 1 u , B B̄ 1 u+ , D 1 u and 5pσ 1 u+ continua. Photon and electron impact dissociation cross sections of H2 X 1 g+ (vi , Ji ) to the continuum levels of B 1 u+ and D 1 u states and the predissociative levels of the D 1 +u state were presented in a previous paper (Liu et al 2009a). While dissociation via the B 1 u+ continuum dominates other singlet ungerade continua at room temperature, the importance of the 0953-4075/12/015201+15$33.00 1 © 2012 IOP Publishing Ltd Printed in the UK & the USA J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al a quantitative understanding of heating of the interstellar medium (ISM) and the atmospheres of comets and outer planets. The fact that H2 is homonuclear results in significant deviation from local thermodynamic equilibrium (LTE) (Liu et al 2007, Shemansky et al 2009, Habart et al 2005). Modelling planetary and astrophysical observations generally requires state-specific excitation, emission and dissociation cross sections. Photoexcitations from the X 1 g+ state to the continuum and predissociative levels of singlet ungerade states and to the X 2 g+ ground ionic state are the two dominant mechanisms for photo-destruction of H2 . While dissociation and ionization via excited ionic states and doubly excited states also take place, the cross sections of these transitions are very small, as they all require a change of two electron configurations and take place via electron correlation. Experimental measurements by Chung et al (1993) and He et al (1995) have shown that the cumulative photoionization cross sections of the excited ionic states and doubly excited states are very small. Electron impact destruction of H2 differs from photon destruction in several ways. In addition to excitation to the dissociative and predissociative levels of singlet ungerade states and to the X 2 g+ ionic state, electron impact dissociation also takes place by excitation to the continuum and predissociative levels of the triplet states and singlet gerade states. Electron impact dissociation occurs via a larger number of doubly excited states and excited ionic states than photon excitation and thus is generally more important than its photon dissociation counterpart. The contribution of the additional electron exchange channels is either small or important only in the low energy region. In fact, when electron impact energy is higher than 25 eV, dissociation via the X 2 g+ ionic state and the dissociative and predissociative levels of the singlet ungerade states is more probable than that via all other states combined (i.e. triplet, singlet gerade, dissociative ionic and doubly excited states). The dominance of the former becomes progressively more prominent with the increase of electron impact energy. The fundamental reason for such dominance is that excitation to both the X 2 g+ state and singlet ungerade states is dipole and spin allowed, whereas other channels are either dipole or spin forbidden. Corrigan (1965) carried out the only existing published electron impact total dissociation measurement by observing the rate of H2 pressure decrease in a closed system where the dissociation fragments were trapped on a molybdenum trioxide (MoO3 ) surface. MoO3 absorbs hydrogen atoms with 100% efficiency. The H+ and H+2 produced by ionization recombine with electrons rapidly on the surface to form hydrogen atoms, which are then absorbed by the surface. Corrigan also derived the so-called total triplet cross section by subtracting the ionization cross section of Tate and Smith (1932) from the measured total dissociation cross section. His approach, however, encountered a number of difficulties. First, the Tate and Smith (1932) ionization cross section becomes larger than his dissociation cross section when E is above ∼90 eV. More importantly, dissociation via singlet ungerade and gerade states was completely neglected in the Corrigan treatment. The dissociation cross section of singlet Figure 1. Singlet ungerade adiabatic potential energy curves and corresponding dissociation limits. Energies are relative to the (vi = 0, Ji = 0) level of the X 1 g+ state, which is not shown. Note that the dissociation limit of the B B̄ 1 u+ state is H(1s)+H(3). See table 1 for more accurate values of the dissociation energies. All potential energy curves are based on the calculations of Staszewska and Wolniewicz (2002), and Wolniewicz and Staszewska (2003a, 2003b). ungerade continua obtained in this work is an important step towards the examination of the accuracy of the Corrigan (1965) measurement and, ultimately, extraction of the total triplet excitation cross section from the experimental measurement. The vacuum ultraviolet spectrum of H2 arises from transitions between the X 1 g+ state and the npσ 1 u+ and npπ 1 u Rydberg series. The emission spectrum is dominated by the first two members of the series: the B 1 u+ and C 1 u states. Below 1000 Å, the contributions from the higher states, such as B 1 u+ and D 1 u , become significant. Figure 1 shows the adiabatic potential energy curves and dissociation limits for several of the lowest singlet ungerade states. The B 1 u+ , C 1 u and B 1 u+ states are not predissociated by other singlet ungerade states. However, excitation to the H(1s)+H(2) continuum is significant. At room temperature, dissociative excitation to the continua of the B 1 u+ and C 1 u states from the ground X 1 g+ (0) level is weak but not negligible (Allison and Dalgarno 1969, Glass-Maujean et al 1985a), while the dissociation via the B 1 u+ continuum is very significant (Glass-Maujean et al 1985a, Glass-Maujean 1986). Glass-Maujean (1986) has also shown that dissociation via the C 1 u continuum from the vibrationally excited X 1 g+ state is very significant. Predissociation is possible for the other npσ 1 u+ and npπ 1 +u states. A number of experimental and theoretical investigations have shown that predissociation of the npσ 1 u+ and npπ 1 +u states is primarily caused by direct or indirect coupling to the B 1 u+ continuum (Glass-Maujean 1979, Glass-Maujean et al 1979, 1984, 1985b, 1987). 2 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al The electron excitation and emission cross sections of the discrete levels of the B 1 u+ and C 1 u states have been determined over a wide energy range from the excitation function measurement by Liu et al (1998) using the Abgrall et al (1993a, 1993b, 1993c) transition probabilities. The B B̄ 1 u+ and D 1 u cross sections for the transitions to the levels below their dissociation limits have been obtained by Glass-Maujean et al (2009) from threshold to 1000 eV. However, the cross sections of other singlet ungerade states, in general, are only available at a few energies, usually at 100 eV (Jonin et al 2000, Glass-Maujean et al 2009). The review article on electron impact cross sections of H2 by Yoon et al (2008) stressed the need for cross sections of higher singlet ungerade states over a wide energy range. A recent work by Liu et al (2009a) has obtained excitation, emission and dissociation cross sections of the B 1 u+ and D 1 u states by using the B 1 u+ and C 1 u shape functions of Liu et al (1998) and theoretically calculated oscillator strengths. In addition, the predissociation cross section of D 1 +u by the B 1 u+ continuum was obtained. Other investigators have also calculated photodissociation of various singlet ungerade states. Dalgarno and Allison (1969) and Glass-Maujean (1986) have calculated the photodissociation cross section of B 1 u+ from rotationless levels of the X 1 g+ (0) state. Glass-Maujean (1986) has also examined vibrational dependence of the Ji = 0 dissociation cross section. Burciaga and Ford (1991) have investigated the C 1 u state vibrational shape resonance arising from the small hump in its potential energy curve and examined the effect of the B 1 u+ −C 1 +u nonadiabatic coupling on the photodissociation cross section of the C 1 u state. Beswick and Glass-Maujean (1987) have considered nonadiabatic coupling between the B 1 u+ and B 1 u+ continua and demonstrated quantum interference between photodissociation cross sections of the B 1 u+ and B 1 u+ states. Cheng et al (1998) have experimentally investigated the photodissociaton resonance near the H(1s)+H(2) threshold and reveal a large variation in the dissociation cross section profiles. Glass-Maujean et al (2007a) have recently calculated photodissociation cross sections of the B B̄ 1 u+ and D 1 u states and have achieved good agreement with their high-resolution measurement. The singlet ungerade states of H2 have been extensively investigated experimentally including electron excitation (Ajello et al 1984, Khakoo and Trajmar 1986, Liu et al 1995, 2000, 2002, 2003, Abgrall et al 1997, 1999, Jonin et al 2000, Dziczek et al 2000, Wrkich et al 2002, Kato et al 2008, Glass-Maujean et al 2009), photoabsorption (Herzberg and Howe 1959, Namioka 1964a, 1964b, Takezawa 1970, Herzberg and Jungen 1972, Dabrowski 1984, Glass-Maujean et al 1984, 1985a, 1985b, 1987, 2007a, 2007b, 2007c, 2008a), photoemission (Roncin et al 1984, Larzillière et al 1985, Abgrall et al 1993a, 1993b, 1993c, 1994, Roudjane et al 2006, 2007), photoionization (Dehmer and Chupka 1976, 1995) and nonlinear laser spectroscopy (Hinnen et al 1994a, 1994b, 1995, Hinnen and Ubachs 1995, 1996, Hogervorst et al 1998, Reinhold et al 1996, 1997, De Lange et al 2001, Koelemeij et al 2003, Greetham et al 2003, Ubachs and Reinhold 2004, Hollenstein et al 2006, Ekey et al 2006, Salumbides et al 2008, 2011, Gabriel et al 2009, Ivanov et al 2011). The simultaneous absorption, dissociation, emission and ionization investigations with synchrotron radiation by Glass-Maujean et al (2007a, 2007b, 2007c, 2008a, 2008b), and corresponding theoretical calculations (Glass-Maujean and Jungen 2009, Glass-Maujean et al 2010) have produced many physical parameters such as predissociation, autoionization and emission yields as well as transition probabilities of singlet ungerade levels. A recent synchrotron photoabsorption investigation using a Fourier transform technique has yielded an accurate predissociation cross section of the D 1 +u state (Dickenson et al 2010). As the simplest neutral molecule, molecular hydrogen has been theoretically investigated by many researchers. The two principal theoretical methods used to calculate the excited states of H2 are traditional ab initio calculations, which deal with a few coupled electronic states at most, and multichannel quantum defect theory (MQDT), which treats the whole family of Rydberg states. Since the pioneering work of Kolos and Wolniewicz (1968), ab initio calculations of the potential energies and transition moments have continued to develop for several decades. Accurate electronic potential energy calculations, including the adiabatic, diagonal nonadiabatic, relativistic and radiative corrections, have been carried out (Dressler and Wolniewicz 1986, Wolniewicz and Dressler 1992, 1994, Wolniewicz 1993, 1995a, 1995b, Wolniewicz et al 1998, Staszewska and Wolniewicz 2002, Wolniewicz and Staszewska 2003a, Wolniewicz 2007). Calculations of the H2 transition moment functions (Wolniewicz 1995c, Dressler and Wolniewicz 1995, Wolniewicz and Staszewska 2003a, 2003b) and nonadiabatic coupling of the first several members of the singlet ungerade Rydberg series have been reported recently (Wolniewicz et al 2006). Accurate adiabatic and nonadiabatic corrections of the X 1 g+ state have been obtained recently by Pachucki and Komasa (2009), and very accurate dissociation energies of the X 1 g+ (vi , Ji ) levels of H2 and D2 have been reported by Komasa et al (2011). Since its first application to interpret a high-resolution H2 photoabsorption spectrum by Herzberg and Jungen (1972), MQDT has been developed to treat autoionization, dissociation and the ro-vibronic structures of singlet and triplet manifolds (Jungen and Atabek 1977, Ross and Jungen 1987, 1994a, 1994b, 1994c, Jungen and Ross 1997, Matzkin et al 2000, Ross et al 2001, Kirrander et al 2007, Glass-Maujean and Jungen 2009, Glass-Maujean et al 2010). Of particular relevance, Abgrall et al (1993a, 1993b, 1993c, 1994, 1997, 2000) have carried out extensive semi-ab initio calculations of the nonadiabatic transition probabilities of the B 1 u+ , C 1 u , B 1 u+ and D 1 u −X 1 g+ band systems. Glass-Maujean et al (2007b, 2007c, 2008a, 2008b) have calculated adiabatic transition probabilities of higher npσ 1 u+ and npπ 1 u states and tested the accuracy with highresolution synchrotron photoabsorption measurements. GlassMaujean et al (2009) also performed nonadiabatic calculations of the B 1 u+ , B B̄ 1 u+ , 5pσ 1 u+ , D 1 +u and D 1 +u states for Ji 3 levels. The transition probabilities, along with dissociation, ionization and emission yields, of the 1 + npπ 1 − u −X g band systems have been obtained with MQDT calculations by Glass-Maujean and Jungen (2009) and Glass-Maujean et al (2010). 3 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al 2. Theory where Eph and Ek are in hartree, D(z) in au, R0 in bohr (a0 ) and mH is the mass of a hydrogen atom (Allison and Dalgarno 1969). Differences in the definition of the Hönl–London factors (see Liu et al 2009a) lead to a degeneracy factor, Gij , usually for – transitions, to appear or not appear in equations (1), (5) and (6). The variation in the definition of Hönl–London factors in the literature have been described by Hansson and Watson (2005). In this work, the definition of Hj i (Jj , Ji ) proposed by Hansson and Watson (2005) is followed. Consequently, Gij is not needed. The discrete and continuum nuclear wavefunctions, χvi ,Ji (R) and χEk ,Jj (R), are obtained by numerical solution of the Schrödinger equation. The potential energy curves consist of Born–Oppenheimer potentials (VBO (R)), plus adiabatic (Vadi (R)), relativistic (Vrel (R)) and radiative (Vrad (R)) corrections: Following the earlier work by Liu et al (2009a), calculations of the photodissociation cross sections are outlined here. Throughout this paper, dissociation strictly refers to the nonradiative process of H2 immediately following the excitation to the continuum levels. Indices i and j denote appropriate levels of the X 1 g+ and singlet ungerade states, respectively. 2.1. Photodissociation cross section Under irradiation by photons with energy of Eph = hcν, the dissociation cross section for excitation from (vi , Ji ) to the continuum level (Ek , Jj ) is 8π 3 ν Hj i (Jj , Ji ) ρ(Ek ) σ (vi , Ji ; Eph ) = 3hc J 2Ji + 1 j × |χEk ,Jj (R)|D(R)|χvi ,Ji (R)|2 , V (R) = VBO (R) + Vadi (R) + Vrel (R) + Vrad (R). (1) where Hj i (Jj , Ji ) and D(R) are the Hönl–London factors and the electric dipole transition moment (Le Roy et al 1976). χvi ,Ji (R) and χEk ,Jj (R) are the radial wavefunctions of initial level i and the continuum level j , respectively. ρ(Ek ) is the density-of-states normalization factor at energy Ek = hcνk above the dissociation limit of state j , with δ(Ek − Ek ) ρ(Ek ) (2) Ek = Eph + E(vi , Ji ) − Vj (R → ∞), (3) χEk ,Jj (R)|χEk ,Jj (R) = The VBO (R), Vadi (R), Vrel (R) and Vrad (R) calculated by Wolniewicz (1993) and Wolniewicz et al (1998) are used for the X 1 g+ potential. The VBO (R) and Vadi (R) of the B 1 u+ , C 1 u , B B̄ 1 u+ , D 1 u and 5pσ 1 u+ states are all taken from ab initio calculations of Staszewska and Wolniewicz (2002), and Wolniewicz and Staszewska (2003a, 2003b). The Vrel (R) of the B 1 u+ and B B̄ 1 u+ states are based on results of Wolniewicz (1995b) and De Lange et al (2001), respectively. Since relativistic corrections for other singlet ungerade states are not available, they are assumed to be R-independent and are derived from corresponding corrections of correlated hydrogen atoms at R → ∞. Specifically, the Vrel (R) of the C 1 u , D 1 u and 5pσ 1 u+ states are −1.673, −1.469 and −1.622 cm−1 , respectively. The Vrad (R) of all the singlet ungerade states are assumed to be equal to the radiative correction of H+2 X 2 g+ calculated by Bukowski et al (1992). The potential energy curves of the B 1 u+ , C 1 u , B 1 u+ , D 1 u , B B̄ 1 u+ , D 1 u and 5pσ 1 u+ states are displayed in figure 1. The photodissociation cross section depends on the amplitude of the continuum radial wavefunction. The converged amplitude is determined by normalization of the wavefunction via equation (4) in the asymptotic region. The cross section is then appropriately scaled by the amplitude. In some cases, propagation of the continuum wavefunction beyond the available ranges of internuclear distance for V (R) and D(R) is required to achieve the convergence. Accurate evaluation of the cross section thus requires appropriate extrapolation of V (R) and D(R) in the asymptotic region. The following extrapolation procedure is used in this work. First, a value of 255 475.6131 cm−1 is added to V (R) in equation (7) so that all potentials are relative to the v = 0 and J = 0 level of the X 1 g+ state. This added value is based on the updated H atom ionization potential and measured dissociation energy of X 1 g+ (vi = 0, Ji = 0) recently obtained by Liu et al (2009b), and is slightly different from the 255 475.612 cm−1 value obtained by Wolniewicz (1995a). V (R) is then extrapolated according to α β γ V (R) = V (∞) + 6 + 8 + 10 . (8) R R R and where Vj (R → ∞) is the asymptotic potential energy of state j . In the present calculation, the internuclear distance, R, is converted to a dimensionless quantity z = R/R0 , where R0 is an arbitrarily selected scale length. The amplitude of the continuum wavefunction is asymptotically normalized to unity: lim χEk ,Jj (z) = sin[kz + ηJj (Ek )], (4) √ where ηJj (Ek ) is the phase shift and k = 2π R0 2μcνk / h, with μ being the reduced mass of H2 . The conversion and √ normalization give a density-of-states factor of ρ(Ek ) = 2R0 2μc/ hνk (states per cm−1 ). The photodissociation cross section in equation (1), in units of Mb, can be re-written as 2 H (J , J ) ji j i ph −2 μν σ (vi , Ji ; Eph ) = 3.2270 × 10 νk J 2Ji + 1 z→∞ j × |χEk ,Jj (z)|D(z)|χvi ,Ji (z)| , 2 (5) −1 where νk and ν are in units of cm , D(z) in debye, μ in unified atomic mass units and R0 in Å (Liu et al 2009a). In atomic units, the photodissociation cross section can be alternatively written as 2 4μEph Hj i (Jj , Ji ) ph σ (vi , Ji ; Eph ) = 25.936 mH Ek J 2Ji + 1 j × |χEk ,Jj (z)|D(z)|χvi ,Ji (z)|2 , (7) (6) 4 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al Table 1. Parameters for asymptotic potentials and transition moments. State αa βa γa V (∞)a,b D(∞)c X 1 g+ B 1 u+ C 1 u B B̄ 1 u+ D 1 u 5pσ 1 u+ 2.861 991E6 8.916 562E11 8.381 670E9 1.650 635E10 9.906 735E9 3.666 206E10 −6.172 744E8 −4.252 095E15 −1.355 295E13 −4.648 399E13 −3.085 039E13 −1.630 490E14 3.400 870E10 5.125 551E18 0.0 3.341 543E16 2.607 558E16 1.858 102E17 36 118.0696 118 377.2327 118 377.2327 133 610.3531 138 941.9834 133 610.2913 – 1.053 498 1.053 498 0.421 875 0 0 The units for α, β and γ are such that V (R) and V (∞) are in cm−1 and R in Å (see equation (8)). b The values of V (R) and V (∞) are relative to the vi = 0 and Ji = 0 level of the X 1 g+ state (see text). c In au, for transition between X 1 g+ and the indicated singlet ungerade state. a The second to fifth columns of table 1 list the values of α, β, γ and V (∞) for the X 1 g+ , B 1 u+ , C 1 u , B B̄ 1 u+ , D 1 u and 5pσ 1 u+ states when the potential is in units of cm−1 and the internuclear distance in Å. Ab initio transition moments, D(R), calculated by Wolniewicz and Staszewska (2003a, 2003b) are utilized for the calculation of photodissociation cross sections. Beyond the available calculated R values, the transition moment is extrapolated with the form 2.2. Electron impact dissociation The electron impact excitation cross sections of this investigation are obtained from the modified Born approximation developed by Shemansky et al (1985a, 1985b). In this formulation, the energy dependence of the cross section is represented by an excitation shape function, which is often obtained from the measurement of the relative emission intensity of a discrete–discrete transition as a function of the impact energy from threshold to a few keV. If the electron excitation is dipole-allowed, application of the Born approximation in the high-energy asymptotic region enables the determination of the absolute value of the excitation cross section by normalizing to the optical oscillator strength. When the shape function is expressed in terms of electron energy in threshold units, an entire electronic band system can be characterized with an emission measurement over a few rovibronic levels. The cross section for excitation (vi , Ji ) to (vj , Jj ) is given by δ , (9) Rn where δ and n are parameters. The last column of table 1 gives the asymptotic values, D(∞), in au for relevant transitions between X 1 g+ and singlet ungerade states. In the absence of nonadiabatic coupling, the calculated photodissociation cross section can be considered as nearly exact. The calculated eigenvalues of the J 27 levels of the X 1 g+ state, when combined with the nonadiabatic correction derived from the polynomial of Wolniewicz (1995a), agree within fractions of 1 cm−1 to the derived experimental values of Dabrowski (1984). When compared with the more accurate theoretical calculation of Komasa et al (2011), the calculated values differ less than 0.1 cm−1 for J 25 levels. However, the difference increases with J, with the largest, ∼0.4 cm−1 , at J = 31 of v = 0. The use of polynomial representation of nonadiabatic correction, which is less accurate at high J, is primarily responsible for the difference in the calculated term values. Moreover, the calculated eigenvalues of the C 1 − u rovibrational levels agree with those measured and calculated by Abgrall et al (1993b, 1993c, 1994) within fractions of 1 cm−1 . The calculated energy values of the D 1 − u levels differ by no more than 1 cm−1 from those measured by Namioka (1964b), Takezawa (1970), Herzberg and Jungen (1972), Dehmer and Chupka (1976) and Roncin and Launay (1994). The photodissociation oscillator strength is related to the photodissociation cross section by ∞ σ ph (vi , Ji ; Ek ) dνk , (10) f (vi , Ji ; j ) = 1.1296 × 10−6 D(R) = D(∞) + σ (vi , vj ; Ji , Jj ) = with C0 Sij (X) = C7 πf (vi , vj ; Ji , Jj ) Sij (X) Eij E 1 1 − 3 X2 X × exp(−mC8 X) + (11) 4 Cm (X − 1) C7 m=1 C5 C6 1 + ln(X), + C7 C7 X (12) and C7 (vi , vj ; Ji , Jj ) = 2π(2Ji + 1) f (vi , vj ; Ji , Jj ), Eij (13) where the cross section σij and collision strength parameter C7 are in units of a20 . Both threshold energy, Eij , and electron impact energy, E, are in hartree. X is the excitation energy in threshold energy units (i.e. X = E/Eij ), and f (vi , vj ; Ji , Jj ) is the optical absorption oscillator strength. The coefficients Cm /C7 (m = 0–6) and C8 are normally determined by nonlinear least-squares fitting of the experimentally measured relative excitation function. It is implicitly assumed that the parameters Cm /C7 (m = 0–6) and C8 depend only on the electronic quantum number and are independent of rotational and vibrational quantum numbers. Note that the shape function, Sij (X), characterizes the energy dependence 0 where σ is in Mb and νk is in cm−1 . Under the irradiation of a uniform photon field, the oscillator strength, f (vi , Ji ; j ), apart from a field strength constant, gives a direct measurement of the photodissociation rate. ph 5 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al of the cross section while C7 , which is proportional to the photoabsorption oscillator strength, determines the absolute magnitude of the cross section. For H2 singlet ungerade states, Liu et al (1998) found that the Sij (X) of the H2 B 1 u+ −X 1 g+ and C 1 u −X 1 g+ band systems are identical within their experimental uncertainties. Experimentally measured collision strength parameters of the higher singlet ungerade states are not available. In this study, the Sij (X) obtained by Liu et al (1998) for the B 1 u+ and C 1 u states are used for these high singlet ungerade states. Equation (11) is for a discrete to discrete excitation. Electron excitation cross sections to the continuum levels, σ (vi , Ji ; E), can be obtained by dividing the continuum into very small intervals (1–8 cm−1 ). The integrated photodissociation oscillator strength and averaged Eph over an interval are taken as fij and Eij of equation (11) for the interval to calculate its contribution to the electron excitation cross section. The excitation cross section arising from continuum transitions is then obtained by summing the contribution of each interval. The thermally averaged dissociation cross section for a band system in this study is defined as the population-weighted average of the ro-vibrational cross-section components: 1 σij (E) = σ (vi , Ji ; E)gI (2Ji + 1) QT i,j × exp[−E(vi , Ji )/kT ], Figure 2. Comparison of photodissociation cross sections from the Ji = 3 level of H2 X 1 g+ (vi = 0–7) to the continuum level of the C 1 u state. The cross sections are in units of Mb. Ek refers to the total kinetic energy of the two dissociating hydrogen atoms. Note that the photodissociation cross sections from vibrationally excited levels are much greater than from the vi = 0 level. levels. The present cross sections generally agree with those of Glass-Maujean (1986) within ∼5%. In the high kinetic energy region, where the cross section drops by a factor between 20 and 100 from that in the peak region, the difference can approach ∼15%. Both calculations produce virtually identical photodissociation oscillator strengths for the R(0) branch excitation. The photodissociation cross sections of the R(1) branch transition from X 1 g+ (0) to the B B̄ 1 u+ , D 1 u and 5pσ 1 u+ continua have been presented in graphic form by GlassMaujean et al (2007a, 2008a). The present cross sections agree with those shown by Glass-Maujean et al if their values are multiplied by appropriate Hönl–London factors for the R(1) branch transition. It appears that Glass-Maujean et al (2007a, 2008a) intentionally left out or implicitly assumed a unity value for the Hönl–London factors. The implicit assumption is clear from their definition of the photodissociation cross section (Glass-Maujean et al 2007a), where the Hönl–London factor is absent from their equation for the (v , J )− (E, J ) photoabsorption cross section, σE,J ,J . Moreover, both measured and calculated P- and R-branch discrete transition probabilities in the tables of Glass-Maujean et al (2007b, 2008a, 2008b) are presented in the form of equivalent band transition probabilities (i.e. branch transition probabilities divided by appropriate Hönl–London factors). Figure 2 shows the C 1 u photodissociation cross section by comparing excitation from the Ji = 3 of vi = 0–7 levels of X 1 g+ . It is clear that the cross section has very strong vibrational dependence. Moreover, the C 1 u dissociation cross sections from the vi > 0 levels are significantly larger than that from the vi = 0 level. Thus, any significant vibrational excitation of the X 1 g+ state will lead to a very large enhancement of the C 1 u state dissociation cross section. (14) where QT is the partition function at temperature T, gI is the nuclear spin statistics and k is the Boltzmann constant. If two states have identical shape functions and similar threshold energies, equation (11) shows their relative cross section is in the ratio of the oscillator strengths. As the principal quantum number increases, the npσ 1 u+ and npπ 1 u Rydberg series converge to the X 2 g+ state of H+2 . The energy gap between two adjacent Rydberg states decreases according to (n∗ )−3 , where n∗ is the effective principal quantum number. The threshold energies of ro-vibrational levels of high n states are thus very similar. Moreover, their potential energy curves also become very similar as they are related to the energy of the X 2 g+ state by the standard Rydberg formula (Glass-Maujean and Jungen 2009, Glass-Maujean et al 2009). As section 2.1 shows, the photodissociation oscillator strength is proportional to the photodissociation cross section, which in turn, is proportional to the square of the dipole matrix elements. The square of the dipole matrix element of a Rydberg state is also proportional to (n∗ )−3 (Glass-Maujean et al 2009). Thus, relative electron impact excitation cross sections of high n states, beyond the threshold region, are inversely proportional to the cubic power of the ratio of the effective principal quantum number. This relation is utilized to estimate the dissociation cross section of states higher than D 1 u and 5pσ 1 u+ . 3. Results and discussion 3.1. Photodissociation cross section Glass-Maujean (1986) has also calculated B 1 u+ and C 1 u photodissociation cross sections from Ji = 0 of various vi 6 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al Figure 4. Comparison of photodissociation oscillator strengths from Ji = 1 of the vi = 0–14 levels to various singlet ungerade continua. Note that the oscillator strength of the 5pσ 1 u+ state shown in the figure has been multiplied by a factor of 10. The dominance of the photodissociation via the C 1 u and B 1 u+ continua from excited vi levels is clear. Thin lines connecting the discrete points are to guide viewing. Figure 3. Comparison of photodissociation oscillator strengths of the Ji = 0–31 levels of H2 X 1 g+ (0) to various singlet ungerade continua. The oscillator strength of the 5pσ 1 u+ state has been scaled up by a factor of 10 and values of the B 1 u+ and D 1 u states are from the previous calculation of Liu et al (2009a). Note that the values for both C 1 u and D 1 u increase rapidly with Ji . The dominance of the photodissociation via the C 1 u continuum at high Ji levels is obvious. Thin lines connecting the discrete points are to guide viewing. by the B 1 u+ −C 1 +u coupling and small differences in the two potential energy curves. Burciaga and Ford (1991) also suggested that some resonances in the C 1 u cross section might be observable experimentally. However, the presence of the much larger B 1 u+ cross section in the same energy region makes the detection of the resonances unlikely, at least for excitation from the X 1 g+ (0) level. The potential energy curve of the 5pσ 1 u+ state also has a hump of ∼1316 cm−1 that peaks near R = 5.6 a0 (see figure 1). The photodissociation cross section also shows large and sharp shape resonance features. As noted by GlassMaujean et al (2007a), extensive resonance peaks also exist in the photodissociation of the B B̄ 1 u+ state. In addition to the shape resonances, quasi-resonance transitions that correspond to excitation to the quasi-bound levels also occur in the threshold region. While the energies of these levels are above their singlet ungerade dissociation limits, V (∞), they can be temporarily stabilized by the centrifugal potentials. Since the quasi-resonances of the B 1 u+ and D 1 u cross sections have been discussed in detail elsewhere (Liu et al 2009a), the quasi-resonance of other singlet ungerade states will not be repeated here. While other states, such as B 1 u+ , D 1 u and 5pσ 1 u+ , also show strong vibrational dependence, the increase in the absolute cross sections of these states is not as large as that of the C 1 u state. Singlet ungerade dissociation cross sections also show strong Ji dependence. For a given vi level, photodissociation cross sections very often, but not always, monotonically increase with Ji . Since the oscillator strength directly reflects the cumulative effect of the photodissociation cross section in a uniform photon radiation field,4 the (vi , Ji ) dependence of the photodissociation cross section can be more conveniently shown in terms of ro-vibrational dependence of oscillator strength. Figure 3 shows the rotational dependence of the photodissociation oscillator strength from the X 1 g+ (0) level while figure 4 shows the vibrational dependence from the Ji = 1 level. The potential energy curve of the C 1 u state has a 101.6 cm−1 hump (above V (∞)) which peaks near R = 9.1 a0 . Consequently, the photodissociation cross section of the C 1 u state shows extensive resonance features. Burciaga and Ford (1991) have investigated these C 1 u shape resonances caused by the hump. They also examined the effect of B 1 u+ −C 1 +u nonadiabatic coupling on the photodissociation cross section of the C 1 u state. The present adiabatic calculation has confirmed the presence of these resonances. However, the location and the magnitude of the resonance cross sections differ slightly from those of Burciaga and Ford (1991). The difference is likely caused 3.2. Electron impact dissociation cross sections The shapes of excitation functions for the singlet ungerade continua are implicitly understood to be identical to their discrete counterparts, which, in turn, are assumed to be the same as those of the B 1 u+ −X 1 g+ and C 1 u −X 1 g+ bands obtained by Liu et al (1998). Since the continuum level can be considered as an extension of the discrete level within a common electronic state, the first assumption is expected to be valid. Additionally, it is very difficult, if 4 Note that electron–matter interaction at high impact energy is equivalent to photon–matter interaction with a constant photon field strength with respect to frequency. 7 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al not impossible, on a state-to-state specific basis, to measure the excitation or emission intensity of a discrete-continuum transition as a function of excitation energy. The first assumption is thus necessary. In the case of the second assumption, several experimental observations suggest that the cross section shapes of singlet ungerade states are similar. The high resolution (δλ ∼ 0.1 Å) emission spectra obtained by Jonin et al (2000) and Glass-Maujean et al (2009) have shown that the shapes of the B 1 u+ , D 1 u and D 1 u states are very similar to those of the B 1 u+ and C 1 u states near 100 eV. Emissions from the B B̄ 1 u+ and 5pσ 1 u+ states are either very weak or from levels that are strongly perturbed. While no definitive conclusion can be provided on the similarity of the shape of the B B̄ 1 u+ and 5pσ 1 u+ excitation functions, the observed relative emission intensities do not suggest that their shapes are significantly different from the low-lying states. Although suffering from spectral overlap, excitation function measurements by Ajello et al (1984) at a resolution of δλ = 5 Å from threshold to 350 eV also suggest that the shapes of the B 1 u+ , B 1 u+ , C 1 u and D 1 u excitation functions are similar. The effect of the second assumption on the accuracy of the singlet ungerade dissociation cross sections will be discussed later. Tables 2–4 list electron impact excitation cross sections from the X 1 g+ state to the various singlet ungerade continua from threshold to 1000 eV at temperatures of 300, 1400 and 5000 K, respectively. For comparison, the B B̄ 1 u+ and D 1 u cross sections based on the earlier work of Liu et al (2009a) are also presented. The difference between the excitation and emission cross sections of the B 1 u+ state in table 1 of Liu et al (2009) corresponds to the B 1 u+ dissociation cross section listed in the fourth column of table 2. The D 1 u dissociation cross section in the sixth column of the tables in this work corresponds to the values under the heading σ3p in table 2 of Liu et al (2009)5 . Note that B 1 u+ and D 1 u cross sections at 300, 1000 and 1500 K were tabulated in the earlier work while the dissociation cross sections at 300, 1400 and 5000 K are given in this work. The ninth column under the heading of Others represents the sum of the excitation cross sections into the continuum levels of npσ 1 u+ and npπ 1 u Rydberg states that are higher than those listed in columns 2–8. The cross sections for excitation into these continuum levels are estimated on the basis of the ratio of effective principal quantum number as described in section 2.2. At 300 and 1400 K, the contribution of these other states accounts for less than 6% of the total dissociation cross section. At 5000 K, the contribution is about 7% of the total. The aggregate contribution from other singlet ungerade states that do not belong to the npσ and npπ Rydberg series (such as 4B 1 u+ , 6B1 u+ and V 1 u ) is completely negligible even at 5000 K. Therefore, the total cross section listed in the last column of the tables 2–4 also represents the dissociation cross section via all singlet ungerade continua. While tables 2–4 list the cross sections up to 1000 eV excitation energy, they can be easily extended to the non-relativistic Born limit. Beswick and Glass-Maujean (1987) investigated the effect of nonadiabatic coupling between B 1 u+ and B 1 u+ around R ∼ 15 a0 on the photodissociation cross sections of these two states. The calculations show an oscillating behaviour of partial cross sections, with the maxima of the B 1 u+ corresponding to the minima of the B 1 u+ and vice versa, although the sum of the two partial cross sections is identical to the sum of the partial cross sections obtained without nonadiabatic coupling. This effect is caused by couplinginduced transitions between the two states. Nonadiabatic coupling is expected to have a smaller effect on individual state electron dissociation cross sections than the corresponding photodissociation cross sections. This is because the oscillator strength is an integration of photodissociation cross section over Ek and the effect of the oscillation in each partial cross section tends to cancel out over a wide Ek range. In particular, the total electron impact dissociation cross section should not be affected by nonadiabatic coupling. There are three primary sources of uncertainty for the individual state cross sections. First is experimental error of the B 1 u+ −X 1 g+ and C 1 u −X 1 g+ excitation functions. The estimated error of the B 1 u+ and C 1 u discrete excitation cross sections above threshold (>17 eV) is 16%, while the relative error between the B 1 u+ and C 1 u cross sections is 10% (Liu et al 1998). The second source of error is the possibility that the particular excitation shape function is somewhat different from those of the B 1 u+ and C 1 u states. Based on the experimental observations mentioned previously, an upper error limit of 10% can be attributed to the possible difference in the shape of the excitation functions. The final source of the error, ∼6%, represents the imprecision of the oscillator strength, which arises from inaccuracy of the transition moments and the neglect of nonadiabatic coupling among the singlet ungerade states. On the basis of the square root of the sum of squares for the individual components, the error in the B 1 u+ and C 1 u dissociation cross section is about 17% when the excitation energy is above 17 eV. The corresponding error of the B 1 u+ , D 1 u , B B̄ 1 u+ , D 1 u and 5pσ 1 u+ states is estimated to be 20% at electron impact energy >20 eV. The relative error among the dissociation cross sections of the B 1 u+ , C 1 u , B 1 u+ , D 1 u , B B̄ 1 u+ , D 1 u and 5pσ 1 u+ states, however, is less than 15%. Estimation of the higher npσ 1 u+ and npπ 1 u continua with the ratio of the effective principal quantum numbers introduces an additional source of error for these cross sections. If a generous value of 10% is assumed for this additional error, the overall error of the cross sections listed in the ninth column of tables 2–4 is 22%. The overall error for the total dissociation cross section of the singlet ungerade continua, however, is lower for two reasons. First, the B 1 u+ and C 1 u cross sections, which have lower error limits, are significantly greater than the sum of the higher npσ 1 u+ and npπ 1 u states, which has a higher error limit. In addition, the electronic transition moments of the B 1 u+ −X 1 g+ , C 1 u −X 1 g+ and B 1 u+ −X 1 g+ band systems are presumably more accurate than those of the higher band systems, and nonadiabatic coupling, while resulting in the repartitioning of the photodissociation cross sections (and 5 Although the D 1 u state is also designated as 3pπ 1 u , which is true for small R, it actually correlates with the H(1s)+H(3d) configuration in the separate atom limit. 8 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al Table 2. Electron impact dissociation cross sections via singlet ungerade continua (T = 300K).a,b E (eV) B 1 u+ C 1 u B 1 u+ c B B̄ 1 u+ D 1 u c D 1 u 5pσ 1 u+ Othersd Total 14.8 15 15.5 16 16.5 17 17.5 18 18.5 19 19.5 20 22.5 25 27.5 30 35 40 45 50 60 70 80 90 100 150 200 250 300 350 400 450 500 550 600 650 700 750 800 850 900 950 1000 0.00 0.02 0.08 0.15 0.22 0.28 0.35 0.41 0.47 0.52 0.57 0.62 0.82 0.98 1.10 1.19 1.32 1.39 1.43 1.45 1.45 1.43 1.39 1.35 1.31 1.14 1.02 0.92 0.84 0.77 0.71 0.66 0.62 0.58 0.55 0.52 0.49 0.47 0.45 0.43 0.41 0.39 0.38 0.01 0.04 0.20 0.38 0.56 0.74 0.90 1.06 1.21 1.35 1.48 1.61 2.14 2.55 2.86 3.10 3.43 3.63 3.73 3.78 3.78 3.71 3.62 3.51 3.41 2.97 2.65 2.40 2.18 2.00 1.85 1.72 1.61 1.51 1.42 1.34 1.28 1.21 1.16 1.11 1.07 1.02 0.99 0.03 0.14 0.69 1.40 2.16 2.89 3.60 4.27 4.90 5.50 6.07 6.60 8.89 10.63 11.98 13.01 14.43 15.27 15.74 15.96 15.98 15.71 15.32 14.88 14.45 12.61 11.25 10.17 9.27 8.51 7.86 7.31 6.83 6.41 6.04 5.71 5.42 5.16 4.93 4.72 4.53 4.35 4.19 0.00 0.00 0.00 0.00 0.00 0.01 0.03 0.06 0.09 0.11 0.14 0.16 0.26 0.33 0.40 0.44 0.51 0.55 0.58 0.60 0.61 0.60 0.59 0.58 0.57 0.50 0.45 0.41 0.37 0.34 0.32 0.30 0.28 0.26 0.25 0.23 0.22 0.21 0.20 0.19 0.19 0.18 0.17 0.00 0.00 0.00 0.00 0.00 0.01 0.04 0.06 0.09 0.11 0.14 0.16 0.26 0.34 0.40 0.45 0.52 0.57 0.59 0.61 0.62 0.62 0.61 0.59 0.58 0.51 0.46 0.42 0.38 0.35 0.33 0.30 0.28 0.27 0.25 0.24 0.23 0.22 0.21 0.20 0.19 0.18 0.18 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.01 0.02 0.03 0.04 0.04 0.08 0.11 0.13 0.15 0.18 0.19 0.20 0.21 0.21 0.21 0.21 0.21 0.20 0.18 0.16 0.15 0.13 0.12 0.11 0.11 0.10 0.09 0.09 0.08 0.08 0.08 0.07 0.07 0.07 0.06 0.06 0.00 0.00 0.00 0.00 0.00 0.01 0.03 0.05 0.07 0.09 0.11 0.13 0.22 0.29 0.34 0.38 0.44 0.48 0.50 0.51 0.52 0.52 0.51 0.50 0.49 0.43 0.39 0.35 0.32 0.30 0.28 0.26 0.24 0.23 0.21 0.20 0.19 0.18 0.17 0.17 0.16 0.15 0.15 0.00 0.00 0.00 0.00 0.00 0.01 0.04 0.09 0.15 0.21 0.28 0.33 0.55 0.72 0.86 0.97 1.12 1.22 1.28 1.32 1.35 1.34 1.32 1.29 1.26 1.11 1.00 0.91 0.83 0.77 0.71 0.66 0.62 0.58 0.55 0.52 0.50 0.47 0.45 0.43 0.42 0.40 0.39 0.04 0.21 0.96 1.93 2.93 3.95 4.98 6.00 6.99 7.92 8.82 9.65 13.22 15.96 18.07 19.71 21.96 23.30 24.07 24.45 24.54 24.15 23.57 22.92 22.26 19.44 17.36 15.71 14.33 13.17 12.17 11.32 10.57 9.93 9.36 8.85 8.41 8.01 7.65 7.32 7.02 6.75 6.50 Unit is 10−19 cm2 . The absolute error of the B 1 u+ and C 1 u states is 17% for E 17 eV. The corresponding error for the B 1 u+ , D 1 u , B B̄ 1 u+ , D 1 u and 5pσ 1 u+ states is less than 20% when E is above 20 eV. The error for the total dissociation cross section is ∼19% also for E > 20 eV. See text for discussion of various errors. Cross section entries are intentionally shown in two digits after the decimal point to allow meaningful comparison of the magnitude of various states and of different temperature. c From Liu et al (2009a). d Sum of npσ 1 u+ with n 6 and npπ 1 u with n 5. a b thus oscillator strengths) among the interacting states, does not alter the total oscillator strength (Beswick and Glass-Maujean 1987). In other words, the ∼6% error, attributed to possible inaccuracy of transition moments and the negligence of nonadiabatic coupling, does not apply to the total dissociation cross section. For this reason, the error limit for the total dissociation cross section is ∼19%, slightly smaller than the limit of individual states such as B 1 u+ , D 1 u , B B̄ 1 u+ , D 1 u and 5pσ 1 u+ . 3.3. Discussion A common feature of photoexcitation from the X 1 g+ state to the singlet ungerade continua is that the dissociation cross sections and oscillator strengths very often, but not always, increase with the initial rotational quantum number. The dependence of the dipole transition moment on internuclear distance may be partially responsible for the increase. In addition, the small reduced mass of H2 leads to a very large rotational constant. Consequently, the number of bound or discrete vibrational levels decreases fairly rapidly with Jj . 9 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al Table 3. Electron impact dissociation cross sections via singlet ungerade continua (T = 1400K).a,b E (eV) B 1 u+ C 1 u B 1 u+ c B B̄ 1 u+ D 1 u c D 1 u 5pσ 1 u+ Othersd Total 14.5 14.7 15 15.5 16 16.5 17 17.5 18 18.5 19 19.5 20 22.5 25 27.5 30 35 40 45 50 60 70 80 90 100 150 200 250 300 350 400 450 500 550 600 650 700 750 800 850 900 950 1000 0.01 0.01 0.04 0.12 0.21 0.29 0.37 0.45 0.52 0.59 0.65 0.71 0.77 1.01 1.20 1.34 1.45 1.60 1.69 1.73 1.75 1.75 1.72 1.67 1.62 1.58 1.37 1.22 1.11 1.01 0.92 0.85 0.79 0.74 0.69 0.65 0.62 0.59 0.56 0.53 0.51 0.49 0.47 0.45 0.01 0.03 0.11 0.31 0.54 0.76 0.97 1.17 1.36 1.54 1.71 1.87 2.02 2.66 3.15 3.52 3.81 4.20 4.43 4.55 4.61 4.60 4.51 4.40 4.27 4.14 3.61 3.21 2.90 2.64 2.43 2.24 2.08 1.94 1.82 1.72 1.62 1.54 1.47 1.40 1.34 1.29 1.24 1.19 0.01 0.05 0.26 0.90 1.68 2.49 3.28 4.04 4.75 5.43 6.07 6.68 7.25 9.69 11.55 12.98 14.08 15.59 16.47 16.95 17.19 17.19 16.89 16.46 15.99 15.52 13.53 12.07 10.91 9.94 9.13 8.43 7.83 7.31 6.86 6.47 6.12 5.81 5.53 5.28 5.06 4.85 4.66 4.49 0.00 0.00 0.00 0.00 0.00 0.00 0.02 0.05 0.08 0.11 0.13 0.16 0.19 0.30 0.39 0.46 0.51 0.59 0.63 0.66 0.68 0.69 0.69 0.68 0.66 0.64 0.57 0.51 0.46 0.42 0.39 0.36 0.34 0.32 0.30 0.28 0.26 0.25 0.24 0.23 0.22 0.21 0.20 0.20 0.00 0.00 0.00 0.00 0.00 0.00 0.02 0.05 0.08 0.12 0.15 0.18 0.20 0.32 0.42 0.49 0.55 0.63 0.68 0.71 0.73 0.74 0.74 0.72 0.71 0.69 0.60 0.54 0.49 0.45 0.42 0.39 0.36 0.34 0.32 0.30 0.28 0.27 0.26 0.24 0.23 0.23 0.22 0.21 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.01 0.02 0.04 0.05 0.06 0.10 0.13 0.16 0.18 0.21 0.23 0.24 0.25 0.26 0.26 0.25 0.25 0.24 0.21 0.19 0.17 0.16 0.15 0.14 0.13 0.12 0.11 0.11 0.10 0.10 0.09 0.09 0.08 0.08 0.08 0.07 0.00 0.00 0.00 0.00 0.00 0.00 0.01 0.03 0.06 0.08 0.11 0.13 0.15 0.24 0.31 0.37 0.41 0.48 0.52 0.54 0.55 0.56 0.56 0.55 0.54 0.52 0.46 0.41 0.38 0.34 0.32 0.30 0.28 0.26 0.24 0.23 0.22 0.21 0.20 0.19 0.18 0.17 0.17 0.16 0.00 0.00 0.00 0.00 0.00 0.00 0.01 0.04 0.10 0.17 0.24 0.32 0.38 0.62 0.81 0.96 1.08 1.25 1.36 1.42 1.46 1.49 1.49 1.46 1.43 1.39 1.23 1.10 1.00 0.92 0.85 0.79 0.73 0.69 0.64 0.61 0.58 0.55 0.52 0.50 0.48 0.46 0.44 0.43 0.03 0.10 0.42 1.33 2.43 3.55 4.69 5.83 6.97 8.06 9.09 10.09 11.01 14.94 17.96 20.28 22.07 24.54 26.00 26.82 27.23 27.29 26.85 26.19 25.46 24.72 21.58 19.26 17.42 15.89 14.59 13.49 12.53 11.71 10.99 10.36 9.80 9.31 8.86 8.47 8.10 7.77 7.47 7.20 Unit is 10−19 cm2 . The absolute error of the B 1 u+ and C 1 u states is 17% for E 17 eV. The corresponding error for the B 1 u+ , D 1 u , B B̄ 1 u+ , D 1 u and 5pσ 1 u+ states is less than 20% for E 19 eV. The error for the total dissociation cross section is ∼19% for E > 20 eV. See text for discussion of various errors. Cross section entries are intentionally shown in two digits after the decimal point to allow a meaningful comparison of the magnitude of various states and different temperature. c Based on the photodissociation cross section of Liu et al (2009a). d Sum of npσ 1 u+ with n 6 and npπ 1 u with n 5. a b sections. Nevertheless, the thermally averaged cross sections shown in tables 2–4 are a convenient way to illustrate the general trend of the cross section variation with ro-vibrational quantum number. All individual state cross sections listed in tables 2–4 increase with temperature. The increase is largely a consequence of the rotational and vibrational dependence of the dipole matrix elements as illustrated in the variation of the photodissociation oscillator strengths with Ji and vi in figures 3 and 4. In general, the increase of the Jj quantum These missing discrete vibrational states for the higher Jj levels are shifted into the continuum and thus contribute to the dissociation cross section and oscillator strength. The population distribution of H2 has been found to be highly non-LTE in cometary atmospheres (Liu et al 2007) at the top of the thermosphere of Saturn (Shemansky et al 2009), in certain regions of the ISM (Habart et al 2005) and in low-density hydrogen plasma environments. The modelling of these observations generally requires state-to-state cross 10 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al Table 4. Electron impact dissociation cross sections via singlet ungerade continua (T = 5000K).a,b E (eV) B 1 u+ C 1 u B 1 u+ c B B̄ 1 u+ D 1 u c D 1 u 5pσ 1 u+ Othersd Total 14 14.2 14.4 14.7 15 15.5 16 16.5 17 17.5 18 18.5 19 19.5 20 22.5 25 27.5 30 35 40 45 50 60 70 80 90 100 150 200 250 300 350 400 450 500 550 600 650 700 750 800 850 900 950 1000 0.55 0.70 0.86 1.14 1.44 1.94 2.42 2.89 3.33 3.74 4.13 4.49 4.84 5.16 5.47 6.75 7.71 8.43 8.98 9.69 10.08 10.26 10.31 10.19 9.93 9.63 9.32 9.03 7.82 6.94 6.24 5.66 5.18 4.77 4.42 4.12 3.86 3.63 3.44 3.26 3.10 2.96 2.83 2.72 2.61 2.51 0.97 1.25 1.58 2.14 2.76 3.82 4.86 5.84 6.78 7.66 8.49 9.27 10.00 10.69 11.35 14.09 16.15 17.70 18.88 20.42 21.26 21.66 21.78 21.56 21.03 20.40 19.75 19.13 16.58 14.72 13.24 12.02 10.99 10.13 9.39 8.75 8.21 7.73 7.30 6.93 6.60 6.29 6.02 5.78 5.55 5.34 0.27 0.37 0.50 0.79 1.20 2.09 3.07 4.06 5.01 5.93 6.79 7.60 8.37 9.10 9.79 12.70 14.91 16.60 17.89 19.64 20.64 21.17 21.39 21.32 20.89 20.32 19.72 19.12 16.64 14.82 13.37 12.17 11.15 10.29 9.56 8.92 8.37 7.88 7.46 7.08 6.74 6.43 6.16 5.91 5.68 5.47 0.00 0.00 0.00 0.01 0.01 0.03 0.05 0.11 0.18 0.26 0.34 0.42 0.49 0.56 0.63 0.92 1.14 1.31 1.45 1.64 1.75 1.82 1.86 1.88 1.86 1.82 1.77 1.72 1.51 1.35 1.23 1.12 1.03 0.95 0.89 0.83 0.78 0.74 0.70 0.66 0.63 0.60 0.58 0.55 0.53 0.51 0.01 0.01 0.01 0.02 0.04 0.08 0.17 0.29 0.44 0.60 0.75 0.90 1.04 1.17 1.29 1.83 2.24 2.56 2.81 3.16 3.36 3.49 3.55 3.58 3.53 3.45 3.36 3.26 2.86 2.55 2.31 2.11 1.94 1.80 1.67 1.56 1.47 1.38 1.31 1.24 1.19 1.13 1.08 1.04 1.00 0.96 0.00 0.00 0.00 0.00 0.00 0.01 0.02 0.05 0.08 0.13 0.19 0.24 0.29 0.34 0.38 0.58 0.73 0.85 0.95 1.08 1.16 1.21 1.24 1.26 1.25 1.22 1.19 1.16 1.02 0.91 0.83 0.76 0.70 0.65 0.60 0.56 0.53 0.50 0.47 0.45 0.43 0.41 0.39 0.38 0.36 0.35 0.00 0.00 0.00 0.00 0.00 0.01 0.02 0.05 0.09 0.14 0.19 0.24 0.28 0.33 0.37 0.55 0.69 0.80 0.89 1.01 1.08 1.13 1.15 1.17 1.16 1.13 1.11 1.08 0.94 0.85 0.77 0.70 0.65 0.60 0.56 0.52 0.49 0.46 0.44 0.42 0.40 0.38 0.36 0.35 0.33 0.32 0.00 0.00 0.00 0.00 0.00 0.01 0.03 0.09 0.24 0.41 0.62 0.82 1.00 1.16 1.32 1.98 2.50 2.90 3.21 3.66 3.93 4.10 4.19 4.25 4.21 4.12 4.02 3.91 3.43 3.07 2.79 2.55 2.35 2.18 2.03 1.90 1.78 1.68 1.59 1.51 1.44 1.38 1.32 1.27 1.22 1.17 1.79 2.33 2.97 4.11 5.46 7.98 10.64 13.37 16.15 18.86 21.49 23.97 26.31 28.51 30.59 39.40 46.08 51.17 55.06 60.29 63.27 64.83 65.48 65.19 63.84 62.09 60.23 58.41 50.81 45.22 40.77 37.09 33.99 31.36 29.11 27.17 25.49 24.01 22.71 21.55 20.52 19.59 18.75 17.98 17.29 16.65 Unit is 10−19 cm2 . The absolute error of the B 1 u+ and C 1 u states is 17% for E 17 eV. The corresponding error for the B 1 u+ , D 1 u , B B̄ 1 u+ , D 1 u and 5pσ 1 u+ states is less than 20% for E 19 eV. The error for the total dissociation cross section is ∼19% for E > 20 eV. See the text for discussion of various errors. c Based on the photodissociation cross section of Liu et al (2009a). d Sum of npσ 1 u+ with n 6 and npπ 1 u with n 5. a b number pushes more vibrational levels from the discrete region to the continuum region, which leads to the increase in the dissociation cross section. The increased temperature also leads to more population at higher Ji and vi levels. These higher Ji and vi levels have smaller threshold energies (Eij ) to the continuum levels than the lower levels, which results in a larger dissociation cross section in the threshold energy region. It can be noted that 1400 K is on the higher end of the range for the Jovian thermosphere, while 5000 K may be close to the temperature of some hydrogen plasmas. When the population distribution does not deviate significantly from LTE, the averaged cross sections in tables 2–4 provide a quick way to estimate the loss of H2 by singlet ungerade continua. Unless a very large population is present in very high Ji (>23) levels, dissociation through excitation from the vi = 0 level takes place primarily via the B 1 u+ continuum. At room 11 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al ∼63%, respectively, greater than the tabulated ones. Finally, the good agreement between Borges et al (1998) and present cross sections in the high-energy region, to a very large extent, reflects the good agreement in the dissociation oscillator strengths and electronic transition moments. Dissociation by excitation to the continuum levels is not the only path for the fragmentation of H2 through the singlet ungerade states. Autoionization, dissociative emission, predissociation of the discrete levels of singlet ungerade states and dissociative ion-pair formation are additional mechanisms for the break up of H2 . Dissociative emission arises from excitation of H2 to discrete singlet ungerade levels, followed by spontaneous emission to the continuum levels of the X 1 g+ state. In general, the dissociative emission from the B 1 u+ state far dominates over the emission from other states. The C 1 u , B 1 u+ and D 1 u states contribute marginally to the continuum emission (Abgrall et al 1997). For this reason, dissociative emission is usually referred to as the Lyman continuum (Stephens and Dalgarno 1972). Other npσ 1 u+ and npπ 1 u states contribute very little because they have much smaller excitation cross sections and because they have either poor overlap with the X 1 g+ continuum or their emission branching ratio to the X 1 g+ state is significantly reduced by predissociation and autoionization. As stated above, the predissociation of the singlet ungerade levels is largely caused by the coupling with the B 1 u+ continuum. The D 1 +u levels above the H(1s)+H(2) limit are directly predissociated by the B 1 u+ continuum (GlassMaujean et al 1979). The predissociation of npσu 1 u+ (n > 4) takes place by homogeneous coupling with the B 1 u+ continuum levels (Glass-Maujean 1979). The predissociation of npπu 1 +u (n > 3) takes place by either npπu 1 +u −D 1 +u homogeneous coupling followed by D 1 +u −B 1 u+ Coriolis coupling or npπu 1 +u − npσu 1 u+ Coriolis coupling followed by npσu 1 u+ −B 1 u+ homogeneous coupling. The D 1 − u and 1 + npπu 1 − states are not coupled to the B state or other u u npσu 1 u+ states. They can only couple to a dissociating 1 − u state. Among the npπu 1 − u states below the H(1s)+H(n = 1 − 3) limit, C 1 − u is the only dissociative u state. Since 1 − npπu u states are only weakly coupled to the C 1 − u state, their predissociation rates are negligibly small. Autoionization arises from the coupling of npσ 1 u+ and npπ 1 u to the X 2 g+ ionization continuum. Since the classical work of Dehmer and Chupka (1976), significant progress has been made towards obtaining the autoionization yield of many singlet ungerade levels. In particular, recent experimental and theoretical investigations by Glass-Maujean et al (2007b, 2008a, 2008b, 2009, 2010) and by Glass-Maujean and Jungen (2009) have made it possible to reliably calculate predissociation and autoionization cross sections of major singlet ungerade rovibrational levels that are accessible at room temperature. When combined with the dissociation cross section obtained in this work, a reliable estimate of the total cross section of H2 fragmentation through singlet ungerade excitation, by photon or electron, will be obtained. Excitation to ion-pair states and to Rydberg series that converge to the ion-pair states is further fragmentation channel of H2 . An important feature of these Rydberg series is a much smaller Rydberg constant Table 5. Comparison of electron impact dissociation cross sections of the B 1 u+ and C 1 u statesa B 1 u+ E (eV) Present Borges et al 100 200 300 400 500 800 1000 1.31 1.02 0.839 0.711 0.617 0.446 0.379 1.60 1.08 0.843 0.686 0.591 0.485 0.347 a b C 1 u b Present Borges et al b 3.41 2.65 2.18 1.85 1.61 1.16 0.986 4.66 3.17 2.45 2.01 1.71 1.21 1.03 Unit is 10−19 cm2 . Borges et al (1998). temperature, dissociation via the B 1 u+ continuum dominates dissociation through other singlet ungerade continua. Since the predissociation of the singlet ungerade states primarily occurs by direct or indirect coupling with the B 1 u+ continuum, predissociation gives additional importance of the B 1 u+ continuum. However, for excitation from a vibrationally excited level, dissociation via the B 1 u+ and C 1 u continua is more important than via the B 1 u+ continuum (see figure 4). At 5000 K, approximately 1/3 of H2 population is at the vi > 0 levels. Table 4 shows that the dissociation cross section of the C 1 u is comparable to that of the B 1 u+ state. If the H2 population, along with the photon and electron energy distributions, is known, the cross sections obtained in this work make it possible to accurately evaluate the conversion of photon and electron energies into the kinetic energy of hydrogen atoms by way of singlet ungerade continua. While the energy transfer from electrons and photons to hydrogen atoms can take place by other dissociation mechanisms, the cross sections obtained in this work represent a set of basic physical parameters for a quantitative understanding of the heating of H2 -dominated atmospheres by solar radiation and photoelectrons. Unlike the B 1 u+ state, not many theoretical calculations have been carried out for the electron dissociation cross sections of the other singlet ungerade states. Table 5 compares the B 1 u+ and C 1 u dissociation cross section obtained in this work for T = 300 K to those calculated by Borges et al (1998) on the basis of first Born approximation. Except for B 1 u+ and C 1 u cross sections at 100 eV and C 1 u cross section at 200 eV, all other calculated values agree with the present values within 12%. At 100 eV, the calculated B 1 u+ and C 1 u cross sections are about 22% and 37% larger than the present cross sections. Similarly, the C 1 u cross section is nearly 20% greater than the present value at 200 eV. The large difference at 100 and 200 eV is consistent with fact that the Born approximation almost always overestimates the lowenergy cross sections. In the case of the dissociation cross sections of the B 1 u+ state, which have been calculated by many theoretical methods, Liu et al (2009a) found that the Borges et al (1998) cross sections have the best agreement with the values listed in the fourth column of table 2, though the calculated values at 100 and 200 eV are also ∼38% and 12 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al and Bohr radius (by a factor of ∼918.6) than that of H or H2 . Because they are stable only at large internuclear distances (R > 12a0 ), only states with n greater than ∼100 have been observed. Although many investigations have been performed (Chupka et al 1975, Kung et al 1986, Pratt et al 1992, Vieitez et al 2008, 2009, Ekey and McCormack 2011), the magnitude of ion-pair production cross section is not well known. However, since the formation of Rydberg series that converge to ion-pair states takes place in the long-range regime (R > 12a0 ) (Vieitez et al 2008, Vieitez 2010), the dissociation cross section of this channel from the low vi levels is probably very small. Photodissociation of H2 in the ISM primarily takes place by dissociative emission to the X 1 g+ continuum, especially by the B 1 u+ −X 1 g+ continuum. However, in the case of dissociation by the singlet ungerade continua, the variation of the B 1 u+ , C 1 u and B 1 u+ dissociation cross sections with vi has a profound implication for the photodissociation of H2 in the ISM. In contrast to the normal laboratory condition, photodissociation of H2 in the ISM, even in a very cold environment, is more likely dominated by the C 1 u and B 1 u+ continua than by the B 1 u+ continuum transition. The primary reason is the shielding of the interstellar photon radiation field by the ionization continuum of the H atom. The strength of the field becomes very weak once the frequency is higher than 109 679 cm−1 . Since dissociation of H2 from vi = 0 and Ji = 0 requires a minimum photon energy of 118 377 cm−1 , photodissociation requires an internal energy of more than 8698 cm−1 . This means that any significant photodissociation must take place from following ro-vibrational levels: (a) vi = 0 and Ji 13, (b) vi = 1 and Ji 9, (c) vi = 2 and Ji 3 and (d) vi 3. For the cases (c)–(d), figure 4 suggests that B 1 u+ , and, especially, C 1 u are more important than the B 1 u+ continuum. Only when most of dissociated H2 is originally between Ji = 13 and Ji = 22 of the vi = 0 level will the B 1 u+ dissociation be dominant. Even in this case, the contribution of the B 1 u+ and C 1 u continua is very significant. As noted, the uncertainty of the electron impact dissociation cross sections beyond threshold is 16–20%. A substantial portion of this uncertainty is due to the lack of measured excitation functions for the B 1 u+ , D 1 u , B B̄ 1 u+ , D 1 u and 5pσ 1 u+ states and the uncertainty in the measured Lyman and Werner band excitation functions of Liu et al (1998). A reduction of error to 14–15% is achievable by measuring the excitation functions of B 1 u+ , D 1 u , B 1 u+ , D 1 u and 5pσ 1 u+ , along with a few small improvements in the electron gun and data acquisition as described by Young et al (2010). the B 1 u+ continuum, and, especially, the C 1 u continuum becomes more important. The calculated continuum oscillator strengths, along with previously published Lyman and Werner band excitation functions, are utilized to determine electron impact dissociation cross sections through the continuum levels of the singlet ungerade states. In addition to the state-to-state cross sections, thermally averaged cross sections from threshold to 1000 eV at various temperatures are presented for the first time. Significant ro-vibrational dependence of the photodissociation cross sections and oscillator strengths results in temperature dependence in the electron dissociation cross sections. Acknowledgments The authors wish to thank Professor Lutoslaw Wolniewicz for making results of his ab initio calculations accessible. The analysis described in this paper was carried out at Space Environment Technologies (SET) and Jet Propulsion Laboratory (JPL), California Institute of Technology. Work performed at SET is supported by the Cassini UVIS contract with the University of Colorado, by NSF AGS-0938223, and NASA Cassini Data Analysis program. XL, PVJ and CPM acknowledge financial support through NASA’s Planetary Atmospheres Research programs. References Abgrall H, Roueff E and Drira I 2000 Astron. Astrophys. Suppl. Ser. 141 297–300 Abgrall H, Roueff E, Launay F and Roncin J-Y 1994 Can. J. Phys. 72 856–65 Abgrall H, Roueff E, Launay F, Roncin J-Y and Subtil J-L 1993a J. Mol. Spectrosc. 157 512–23 Abgrall H, Roueff E, Launay F, Roncin J-Y and Subtil J-L 1993b Astron. Astrophys. Suppl. Ser. 101 273–322 Abgrall H, Roueff E, Launay F, Roncin J-Y and Subtil J-L 1993c Astron. Astrophys. Suppl. Ser. 101 323–62 Abgrall H, Roueff E, Liu X and Shemansky D E 1997 Astrophys. J. 481 557–66 Abgrall H, Roueff E, Liu X, Shemansky D E and James G K 1999 J. Phys. B: At. Mol. Opt. Phys. 32 3813–38 Ajello J M, Shemansky D E, Kwok T L and Yung Y L 1984 Phys. Rev. A 29 636–53 Allison A C and Dalgarno A 1969 At. Data 5 92–102 Beswick J A and Glass-Maujean M 1987 Phys. Rev. A 35 3339–49 Borges I Jr, Jalbert G and Bielschowsky C E 1998 Phys. Rev. A 57 1025–32 Bukowski R, Jenziorski B, Moszynski R and Kolos W 1992 Int. J. Quantum Chem. 42 287–319 Burciaga J R and Ford A L 1991 J. Mol. Spectrosc. 149 1–14 Cheng C H, Kim J T, Eyler E E and Melikechi N 1998 Phys. Rev. A 57 949–57 Chung Y M, Lee E-M, Masuoka T and Samson J A R 1993 J. Chem. Phys. 99 885–9 Chupka W A, Dehmer P M and Jivery W T 1975 J. Chem. Phys. 63 3929–44 Corrigan S J 1965 J. Chem. Phys. 43 4381–6 Dabrowski I 1984 Can. J. Phys. 62 1639–64 Dalgarno A and Allison A C 1969 J. Geophys. Res. 74 4178–80 Dehmer P M and Chupka W A 1976 J. Chem. Phys. 65 2243–73 Dehmer P M and Chupka W A 1995 J. Phys. Chem. 99 1686–99 De Lange A, Hogervorst W and Ubachs W 2001 Phys. Rev. Lett. 86 2988–91 4. Conclusions Photodissociation cross sections and oscillator strengths for excitations from ro-vibrational levels of the X 1 g+ state to the continuum levels of the H2 singlet ungerade states have been obtained. At room temperature, photodissociation is dominated by the excitation to the B 1 u+ continuum. However, once H2 is vibrationally excited, excitation via 13 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al Kato H, Kawahara H, Hoshino H, Tanaka H, Campbell L and Brunger M J 2008 Phys. Rev. A 77 062708 Khakoo M A and Trajmar S 1986 Phys. Rev. A 34 146–56 Kirrander A, Fielding H H and Jungen C 2007 J. Chem. Phys. 127 164301 Kung A H, Page R H, Larkin R J, Shen Y R and Lee Y T 1986 Phys. Rev. Lett. 56 328–31 Koelemeij J C J, de Lange A and Ubachs W 2003 Chem. Phys. 287 349–54 Kolos W and Wolniewicz L 1968 J. Chem. Phys. 48 3672–80 Komasa J, Piszczatowski K, Lach G, Przybytek M, Jeziorski B and Pachucki K 2011 J. Chem. Theory Comput. 7 3105–115 Larzillière M, Launay F and Roncin J-Y 1985 Can. J. Phys. 63 1416 Le Roy R J, MacDonald R G and Burns G 1976 J. Chem. Phys. 65 1485 Liu X, Ahmed S M, Multari R A, James G K and Ajello J M 1995 Astrophys. J. Suppl. Ser. 101 375–99 Liu X, Johnson P V, Malone C P, Young J A, Shemansky D E and Kanik I 2009a J. Phys. B: At. Mol. Opt. Phys. 42 185203 Liu X, Shemansky D E, Abgrall H, Roueff E, Ahmed S M and Ajello J M 2003 J. Phys. B: At. Mol. Opt. Phys. 36 173–96 Liu X, Shemansky D E, Abgrall H, Roueff E, Dziczek D, Hansen D L and Ajello J M 2002 Astrophys. J. Suppl. Ser. 138 229–45 Liu X, Shemansky D E, Ahmed S M, James G K and Ajello J M 1998 J. Geophys. Res. 103 26739–58 Liu X, Shemansky D E, Ajello J M, Hansen D L, Jonin C and James G K 2000 Astrophys. J. Suppl. Ser. 129 267–80 Liu X, Shemansky D E, Hallett J T and Weaver H A 2007 Astrophys. J. Suppl. Ser. 169 458–71 Liu J, Salumbides E J, Hollenstein U, Koelemeij J C J, Eikema K S E, Ubachs W and Merkt F 2009b J. Chem. Phys. 130 174306 Matzkin A, Jungen Ch and Ross S C 2000 Phys. Rev. A 62 062511 Namioka T 1964a J. Chem. Phys. 40 3145–65 Namioka T 1964b J. Chem. Phys. 41 2141–52 Pachucki K and Komasa J 2009 J. Chem. Phys. 130 164113 Pratt S T, McCormack E F, Dehmer J L and Dehmer P M 1992 Phys. Rev. Lett. 68 584–7 Reinhold E, Hogervorst W and Ubachs W 1996 J. Mol. Spectrosc. 180 156–63 Reinhold E, Hogervorst W and Ubachs W 1997 Phys. Rev. Lett. 78 2543–6 Roncin J-Y and Launay F 1994 Atlas of the Vacuum Ultraviolet Emission Spectrum of Molecular Hydrogen (Journal of Physical and Chemical Reference Data, Monograph no 4) (Washington, DC: American Chemical Society) Roncin J-Y, Launay F and Larzillière M 1984 Can. J. Phys. 62 1686–705 Roudjane M, Launay F and Tchang-Brillet W-U L 2006 J. Chem. Phys. 125 214305 Roudjane M, Tchang-Brillet W-U L and Launay F 2007 J. Chem. Phys. 127 054307 Ross S C and Jungen C 1987 Phys. Rev. Lett. 59 1297–30 Ross S C and Jungen C 1994a Phys. Rev. A 49 4353–63 Ross S C and Jungen C 1994b Phys. Rev. A 49 4364–77 Ross S C and Jungen C 1994c Phys. Rev. A 50 4618–28 Ross S C, Jungen C and Matzkin A 2001 Can. J. Phys. 79 561–88 Salumbides E J, Bailly D, Khramov A, Wolf A L, Eikema K S E, Vervloet M and Ubachs W 2008 Phys. Rev. Lett. 101 223001 Salumbides E J, Dickenson G D, Ivanov T I and Ubachs W 2011 Phys. Rev. Lett. 107 043005 Shemansky D E, Ajello J M and Hall D T 1985a Astrophys. J. 296 765–73 Shemansky D E, Ajello J M, Hall D T and Franklin B 1985b Astrophys. J. 296 774–83 Shemansky D E, Liu X and Melin H 2009 Planet. Space Sci. 57 1659–70 Staszewska G and Wolniewicz L 2002 J. Mol. Spectrosc. 212 208–12 Dickenson G D et al 2010 J. Chem. Phys. 133 144317 Dressler K and Wolniewicz L 1986 J. Chem. Phys. 85 2821–30 Dressler K and Wolniewicz L 1995 Ber. Bunsenges. Phys. Chem. 99 246–50 Dziczek D, Ajello J M, James G K and Hansen D L 2000 Phys. Rev. A 61 4702–6 Ekey R C Jr, Marks A and McCormack E F 2006 Phys. Rev. A 73 023412 Ekey R C Jr and McCormack E F 2011 Phys. Rev. A 84 020501 Gabriel O, van den Dungen J J A, Roueff E, Abgrall H and Engeln R 2009 J. Mol. Spectrosc. 253 64–72 Glass-Maujean M 1979 Chem. Phys. Lett. 68 320–3 Glass-Maujean M 1986 Phys. Rev. A 33 342–5 Glass-Maujean M, Breton J and Guyon P M 1979 Chem. Phys. Lett. 63 591–4 Glass-Maujean M, Breton J and Guyon P M 1984 Chem. Phys. Lett. 112 25–8 Glass-Maujean M, Breton J and Guyon P M 1985a J. Chem. Phys. 83 1468–70 Glass-Maujean M, Breton J and Guyon P M 1987 Z. Phys. D 5 189–201 Glass-Maujean M, Breton J, Thieblemont B and Ito K 1985b Phys. Rev. A 32 947–51 Glass-Maujean M and Jungen Ch 2009 J. Phys. Chem. A 113 13124–32 Glass-Maujean M, Jungen Ch, Reichardt G, Balzer A, Schmoranzer H, Ehresmann A, Haar I and Reiss P 2010 Phys. Rev. A 82 062511 Glass-Maujean M, Klumpp S, Werner L, Ehresmann A and Schmoranzer H 2007a J. Phys. B: At. Mol. Opt. Phys. 40 F19–26 Glass-Maujean M, Klumpp S, Werner L, Ehresmann A and Schmoranzer H 2007b J. Chem. Phys. 126 144303 Glass-Maujean M, Klumpp S, Werner L, Ehresmann A and Schmoranzer H 2007c Mol. Phys. 105 1535–42 Glass-Maujean M, Klumpp S, Werner L, Ehresmann A and Schmoranzer H 2008a J. Mol. Spectrosc. 249 51–9 Glass-Maujean M, Klumpp S, Werner L, Ehresmann A and Schmoranzer H 2008b J. Chem. Phys. 128 94312 Glass-Maujean M, Liu X and Shemansky D E 2009 Astrophys. J. Suppl. Ser. 180 38–53 Greetham G M, Hollenstein U, Seiler R, Ubachs W and Merkt F 2003 Phys. Chem. Chem. Phys. 5 2528–34 Habart E, Walmsley M, Verstraete L, Cazaux S, Maiolino R, Cox P, Boulanger F and Pineau des Forets G 2005 Space Sci. Rev. 119 71–91 Hansson A and Watson J K G 2005 J. Mol. Spectrosc. 233 169–73 He Z X, Cutler J N, Southworth S H, Hughey L R and Samson J A R 1995 J. Chem. Phys. 103 3912–16 Herzberg G and Howe L L 1959 Can. J. Phys. 37 636–59 Herzberg G and Jungen Ch 1972 J. Mol. Spectrosc. 41 425–86 Hinnen P C, Hogervorst W, Stolte S and Ubachs W 1994a Appl. Phys. B 59 307–10 Hinnen P C, Hogervorst W, Stolte S and Ubachs W 1994b Can. J. Phys. 72 1032–42 Hinnen P C, Werners S E, Stolte S, Hogervorst W and Ubachs W 1995 Phys. Rev. A 52 4425–33 Hinnen P C and Ubachs W 1995 Chem. Phys. Lett. 240 351–6 Hinnen P C and Ubachs W 1996 Chem. Phys. Lett. 254 32–9 Hogervorst W, Eikema K S E, Reinhold E and Ubachs W 1998 Nucl. Phys. A 631 353–62 Hollenstein U, Reinhold E, de Lange C A and Ubachs W 2006 J. Phys. B: At. Mol. Opt. Phys. 39 L195–201 Ivanov T I, de Lange C A and Ubachs W 2011 J. Chem. Phys. 134 054309 Jonin C, Liu X, Ajello J M, James G K and Abgrall H 2000 Astrophys. J. Suppl. 129 247–66 Jungen Ch and Atabek O 1977 J. Chem. Phys. 66 5584–609 Jungen Ch and Ross S 1997 Phys. Rev. A 55 2503–6 14 J. Phys. B: At. Mol. Opt. Phys. 45 (2012) 015201 X Liu et al Stephens T L and Dalgarno A 1972 J. Quant. Spectrosc. Radiat. Transfer 12 569 Takezawa S 1970 J. Chem. Phys. 52 2575–90 Tate J T and Smith P T 1932 Phys. Rev. 39 270–7 Ubachs W and Reinhold E 2004 Phys. Rev. Lett. 92 101302 Vieitez M O 2010 XUV laser studies of Rydberg-valence states in N2 and H+ H− heavy Rydberg states PhD thesis Vrije Universiteit Vieitez M O, Ivanov T I, Reinhold E, de Lange C A and Ubachs W 2008 Phys. Rev. Lett. 101 163001 Vieitez M O, Ivanov T I, Reinhold E, de Lange C A and Ubachs W 2009 J. Phys. Chem. A 113 13237–45 Wolniewicz L 1993 J. Chem. Phys. 99 1851–68 Wolniewicz L 1995a J. Chem. Phys. 103 1792–9 Wolniewicz L 1995b Chem. Phys. Lett. 233 647–50 Wolniewicz L 1995c Chem. Phys. Lett. 233 644–6 Wolniewicz L 2007 Mol. Phys. 105 1497–503 Wolniewicz L and Dressler K 1992 J. Chem. Phys. 96 6053–64 Wolniewicz L and Dressler K 1994 J. Chem. Phys. 100 444–51 Wolniewicz L, Orlikowski T and Staszewska G 2006 J. Mol. Spectrosc. 238 118–26 Wolniewicz L, Simbotin I and Dalgarno A 1998 Astrophys. J. Suppl. Ser. 115 293–313 Wolniewicz L and Staszewska G 2003a J. Mol. Spectrosc. 217 181–5 Wolniewicz L and Staszewska G 2003b J. Mol. Spectrosc. 220 45–51 Wrkich J, Mathews D, Kanik I, Trajmar S and Khakoo M A 2002 J. Phys. B: At. Mol. Opt. Phys. 35 4695–709 Yoon J-S, Song M-Y, Han J-M, Hwang S H, Chang W-S, Lee B J and Itikawa Y 2008 J. Phys. Chem. Ref. Data 37 913–31 Young J A, Malone C P, Johnson P V, Ajello J M, Liu X and Kanik I 2010 J. Phys. B: At. Mol. Opt. Phys. 43 135201 15