ANALYSIS OF ELECTRON-IMPACT EXCITATION AND EMISSION OF THE npσ1Σ+u AND npπ1Πu RYDBERG SERIES OF H2

, , and

Published 2008 December 23 © 2009. The American Astronomical Society. All rights reserved.
, , Citation Michèle Glass-Maujean et al 2009 ApJS 180 38 DOI 10.1088/0067-0049/180/1/38

0067-0049/180/1/38

ABSTRACT

Calculated and recently measured photoabsorption transition probabilities of the H2 npσ1Σ+u and npπ1Πu − X1Σ+g band systems have been examined with high-resolution (Δλ = 95–115 mÅ) electron-impact induced emission spectra obtained previously by Jonin et al. and Liu et al. When localized rovibronic coupling is insignificant, transition probabilities calculated with the adiabatic approximation are found to be generally consistent with experiment. However, in the presence of significant coupling, the transition probabilities obtained from a nonadiabatic calculation of B'1Σ+u, D1Π+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, D'1Π+u, and 5pσ1Σ+u state coupling give better agreement with the experimental spectra. Emission yields obtained by comparison of the calculated and experimental spectra are also consistent with the measured predissociation and autoionization yields. In addition, more accurate excitation and emission cross sections and nonradiative yields have been obtained for a number of the npσ1Σ+u and npπ1Πu states. The results obtained in the present investigation lead to a significantly more accurate calibration of the Cassini UVIS instrument and laboratory spectrometers in the range 790–920 Å. They are also an important step toward an accurate model of extreme ultraviolet H2 auroral and dayglow emissions in the outer planet atmospheres.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Molecular hydrogen emission in the vacuum ultraviolet (VUV) arises in transitions from the 1sσgnpσu 1Σ+u and 1sσgnpπu 1Πu Rydberg series to the ground X1Σ+g state. Although the Lyman- (B1Σ+uX1Σ+g) and Werner- (C1ΠuX1Σ+g) band systems dominate in the far ultraviolet (FUV) region, the contribution from the higher npσu and npπu (n ⩾  3) Rydberg states becomes important in the extreme ultraviolet (EUV) region. The large number of states contributing to emission in the EUV region produces a much more congested and complicated spectrum than is found in the FUV region. Additional decay mechanisms for some excited levels also complicate the EUV emission spectrum. In the absence of collisions, predissociation competes with spontaneous emission to depopulate the levels that lie above the H(1s)+H(2ℓ) dissociation limit (Julienne 1971; Glass-Maujean et al. 1987). Above the first ionization limit, autoionization becomes another mechanism of depopulating the npσu and npπu states (Herzberg & Jungen 1972; Dehmer & Chupka 1976). Both predissociation and autoionization of H2 singlet-ungerade states have been investigated extensively (Dehmer & Chupka 1976; Glass-Maujean et al. 1978; Glass-Maujean 1979; Glass-Maujean 1986; Glass-Maujean et al. 1987, 2007a, 2007b, 2007c, 2008a, 2008b; Guyon et al. 1979; Dehmer & Chupka 1995; Dehmer et al. 1989, 1992; Pratt et al. 1990, 1992, 1994; Stephens & Greene 1994).

Electron-impact excitation of molecular hydrogen is an important process in molecular clouds and outer planet atmospheres. Several observations of Jupiter aurorae with the Hubble Space Telescope (HST) in the FUV region have confirmed the importance of the electron-impact excitation process (Clarke et al. 1994; Trafton et al. 1994; Kim et al. 1995). In addition, the Hopkins Ultraviolet Telescope (HUT) observations of Jupiter aurorae and dayglow in both the FUV and EUV regions have revealed the electron-impact excitation of H2 to various singlet-ungerade states (Feldman et al. 1993; Morrissey et al. 1997; Wolven & Feldman 1998). Analyses of Galileo and Far Ultraviolet Spectroscopic Explorer (FUSE) observations of Jupiter auroral emissions in both the EUV and FUV regions by Ajello et al. (1998) and Gustin et al. (2004) have shown intense H2 emission over a range of molecular hydrogen column densities of 1016–1021 cm2. All observations in the EUV wavelength region show significant emission from high Rydberg (n ⩾  3) states between 760 Å  and 900 Å. However, the lack of reliable excitation and emission cross sections, particularly for the large number of transitions on the blue side of 900 Å, results in difficulty in modeling both experimental and spacecraft observations (Liu et al. 2000). Accurate excitation and emission cross sections of the high Rydberg states, therefore, are important to the interpretation of the outer planet observations in the EUV region.

While many investigations on electron-impact induced emission of H2 have been carried out, reliable excitation and emission cross sections for the n > 3 Rydberg states are generally not available. Early low-resolution investigations of the electron-impact induced emission spectrum of H2 have been reported by Ajello et al. (1982, 1984, 1988), who also performed crude and inadequate modeling of the emission spectrum with band transition probabilities by Allison & Dalgarno (1970). Liu et al. (1995) and Abgrall et al. (1997, 1999) have shown that the band transition probabilities partitioned by Hönl–London factors are inaccurate. Transition probabilities calculated by Abgrall et al. (1993a, 1993b, 1993c, 1999) accurately reproduce the experimental intensity distribution. Jonin et al. (2000) and Liu et al. (2000) have extended the high-resolution experimental investigations and theoretical modeling into the EUV region. Their investigation found that B1Σ+uX1Σ+g, C1ΠuX1Σ+g, B'1Σ+uX1Σ+g and D1ΠuX1Σ+g transition probabilities of Abgrall et al. (1993a, 1993b, 1993c, 1994) can reproduce the observed relative intensities in wavelength regions where the contribution from n > 3 Rydberg states is negligible. The calculated spectra of Jonin et al. (2000) reproduced 95%–96% of observed H2 emission intensities in the 900−1040 Å region. The remaining 4%–5% intensity differences are attributed to emission from higher (n > 3) Rydberg states, perturbations between the n ⩽ 3 (primarily the B'1Σ+u) states and n > 3 states, and cascade excitation of the low vj level of the B'1Σ+u state via the singlet-gerade states (Liu et al. 2002). Jonin et al. (2000) also obtained experimental estimates of the emission cross sections of the B'1Σ+u, D1Πu, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, D'1Πu, and D''1Πu states. They, however, encountered a number of difficulties, especially for the transitions below 900 Å. Experimentally, significant contributions from the high-Rydberg (n ⩾  3) states make the EUV emission spectrum much more congested. In the absence of reliable theoretical calculations, it is difficult to appropriately partition the overlapping experimental intensity to individual transitions. The general weakness of transitions for these high-Rydberg states and optical thickness of certain resonance transitions also seriously compromise the analytical interpretation. The lack of transition probabilities and oscillator strengths makes it difficult to estimate the self-absorption of resonance transitions. Finally, the emission yields of many states are not accurately known because of dissociation, predissociation, and autoionization. The combined effects lead to very significant errors in estimated cross sections.

We have re-analyzed the high-resolution electron-impact induced emission spectra of Liu et al. (2000) and Jonin et al. (2000) with recently measured and calculated transition probabilities of the npσu1Σ+u and npπu1Πu − X1Σ+g (n > 3) band systems. Some calculated transition probabilities have been reported recently by Glass-Maujean et al. (2007a, 2007b, 2007c, 2008a), along with high-resolution photoabsorption measurements. The present analysis provides further examination of the accuracy of the calculated transition probabilities. Emission yields of various rovibrational levels of the npσu1Σ+u and npπu1Πu states are determined by comparing observed and calculated spectra. The derived emission yields are then compared with the autoionization yields determined by Dehmer & Chupka (1976) and the predissociation yields by Glass-Maujean et al. (1987). Excitation and emission cross sections of these band systems are obtained from the calculated transition probabilities and measured emission yields.

The singlet-ungerade states of the H2 have been studied by various experimental techniques including photo absorption (Herzberg & Howe 1959; Namioka 1964a, 1964b; Takezawa 1970; Herzberg & 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; Roncin & Launay 1994), photoionization (Dehmer & Chupka 1976, 1995), and nonlinear laser spectroscopy (Hinnen et al. 1994a, 1994b, 1995a, 1995b, 1996; Hogervorst et al. 1998; Reinhold et al. 1996, 1997; De Lange et al. 2001; Koelemeij et al. 2003; Greetham et al. 2003; Ubachs & Reinhold 2004; Hollenstein et al. 2006; Ekey et al. 2006). The spectral atlas of Roncin & Launay (1994), in particular, has provided an extensive tabulation of transition frequencies.

Molecular hydrogen has also been extensitively theoretically investigated. Multichannel quantum defect theory (MQDT) was first developed to interpret high-resolution H2 photoabsorption spectra (Herzberg & Jungen 1972) and H2 autoionization (Jungen & Atabek 1977; Ross & Jungen 1987, 1994a, 1994b, 1994c, 1997). Since the pioneer work of Kolos & Wolniewicz (1968), ab initio calculations of the potential energies have been developed for several decades. Accurate calculations, including the adiabatic and the diagonal nonadiabatic corrections (Wolniewicz 1993; Staszewska &Wolniewicz 2002; Wolniewicz & Staszewska 2003a), have been carried out. The calculations of H2 transition moment functions (Wolniewicz & Staszewska 2003a, 2003b) and nonadiabatic coupling of the first several members of the singlet-ungerade Rydberg series have been recently reported (Wolniewicz et al. 2006).

2. EXPERIMENT

The experimental data used in the present analysis were obtained almost nine years ago. A subset of the measured spectra has been reported by Jonin et al. (2000) and Liu et al. (2000). Since the experimental setup was substantially similar to that described by Jonin et al. (2000) and Liu et al. (1995), only a brief overview will be given here.

The experimental system consists of a 3 m spectrometer (Acton VM-523-SG) and an electron collision chamber. Electrons generated by heating a thoriated tungsten filament are magnetically collimated with an axially symmetric magnetic field of ∼100 G and accelerated to a kinetic energy of 100 eV. The accelerated electrons, which move horizontally, collide with a vertical beam of H2 gas formed by a capillary array. The cylindrical interaction region is about 3 mm in length and ∼2 mm in diameter. Optical emission from electron-impact excited H2 is dispersed by the spectrometer equipped with a 1200 grooves mm−1 grating coated with B4C. The spectrometer has an aperture ratio of f/28.8 and a field of view of 3.8 mm (horizontal) by 2.4 mm (vertical). The dispersed radiation is detected with a channel electron multiplier (Galileo 4503) coated with CsI. A Faraday cup was utilized to minimize the backscattered electrons and monitor the beam current.

Three sets of spectra at different resolutions were used in the present analysis. The first set, acquired in the first order with a slit width of 40 μm, an increment of 0.040 Å, and an integration time of 70 s per channel, had a FWHM of ∼0.115 Å. As reported in Jonin et al. (2000), its effective foreground H2 column density was (2.3± 0.6) × 1013 cm−2 and the spectral range was from 800 Å  to 1440 Å. The second set, obtained with a 25 μm slit width, 0.020 Å wavelength increment, and 90 s integration time, had a FWHM of ∼0.095 Å. The H2 foreground column density was estimated to be (15 ± 54) × 1013 cm−2 by a cross comparison of the intensities of strong nonresonance transitions between cross-beam and swarm measurements (Jonin et al. 2000). While its spectral range was from 788 Å to 1100 Å, only the features in the 790–910 Å region are used for the analysis, as the optical thickness was too high for many Lyman- and Werner-band resonance transitions. The third set of data, obtained with a slit width of 80 μm and ∼ 20% higher pressure (18 ×  1013 cm−2), has a resolution of ∼0.24 Å, and was used to examine the weak H2 emissions between 750 Å  and 800 Å.

The wavelength scale of the observed spectrum was established by assuming a uniform grating step size and by using the absolute wavelength of the H Lyman series emissions. The mechanical limitation of the stepping motor, and, more importantly, the slight temperature fluctuation of the spectrometer (± 0.3°C) during the scan resulted in significant slowly varying nonuniform wavelength shifts. The wavelength error was estimated by comparing the observed spectra with the model spectra, utilizing the experimentally derived energy-term values. The largest wavelength error, from the extremes of negative and positive shifts, was found to be ∼0.04 Å. As frequencies of many strong transitions of H2 have been accurately measured in previous studies, the effect of the small-wavelength deviation can be reduced by aligning the observed and model spectra over strong features. Hence, the wavelength shifts did not cause significant problems for the analysis reported in the present paper.

3. THEORY

3.1. Photon Emission Intensity of Electron-Impact Excitation

Steady-state photon emission intensity resulting from direct excitation by a continuous electron beam has been described in detail by Jonin et al. (2000). A brief review will be given here. The volumetric photon emission rate (I) from electron-impact excitation is proportional to the excitation rate and emission-branching ratio:

Equation (1)

where the indexes j and i refer to the upper and lower electronic-state vibrational and rotational levels, respectively, and summation over the missing index is assumed. J and v refer to rotational and vibrational quantum numbers, respectively. A(vj, vi;Jj, Ji) is the Einstein spontaneous transition probability for emission from level (vj, Jj) to level (vi, Ji), and A(vj, Jj) is the total radiative transition probability for level (vj, Jj) (including transition to lower singlet-gerade states and to the continuum levels of the X1Σ+g state). The yield of nonradiative processes is represented by η(vj, Jj). Under the present experimental conditions, collision deactivation is 104–105 times slower than radiative decay, and is, therefore, negligible. The sum of appropriate predissociation, dissociation, and autoionization yields is denoted by η(vj, Jj). Since the experimental condition for certain strong resonance transitions to the X1Σ+g(0) levels is optically thick, the parameter κ(epsilonij, ζi) in Equation (1) accounts for the self-absorption for those resonance transitions. The calculation of κ(epsilonij, ζi) from H2 foreground column density and transition probability has been given by Jonin et al. (2000).

The excitation rate, g(vj, Jj), represents the sum of the excitation rates from the rotational and vibrational levels of the X1Σ+g state. It is proportional to the population of the molecules in the initial level, N(vi, Ji), the excitation cross section (σij), and the electron flux (Fe):

Equation (2)

The excitation cross section σij is calculated from the analytical function (Liu et al. 1998)

Equation (3)

where a0 and Ry are the Bohr radius and Rydberg constant, respectively, and fij is the optical absorption oscillator strength, given by Equation (16). Eij is the threshold energy for the (vi, Ji) → (vj, Jj) excitation, E is the excitation energy, and X = E/Eij. The coefficients Ck/C7 (k = 0 to 6) and C8 are determined by fitting the experimentally measured relative excitation function. For the present work, Ck/C7 (k = 0 to 6) and C8 determined by Liu et al. (1998) for the B1Σ+uX1Σ+g and C1ΠuX1Σ+g band systems of H2 are used for the direct excitation of the other npσ1Σ+u and npπ1Πu − X1Σ+g band systems. The absolute value of the collision strength parameter, C7, of Equation (3), is fixed to the absorption oscillator strength fij by the relation

Equation (4)

In addition to the direct excitation, indirect excitation from the higher levels of singlet-gerade states is also possible. While the cascade from higher levels of the singlet-gerade states results in the indirect excitation of many rovibronic levels of the singlet-ungerade states, a recent time-resolved study of Liu et al. (2002) has shown that it preferentially contributes to the low vj levels of the B1Σ+u and B'1Σ+u states, and the indirect excitation to the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, D'1Πu, or higher states is negligible.

The excitation cross section for a band system in the present study will be defined as the statistical average of the rovibrational cross section components:

Equation (5)

where NT is the total H2 population of the X1Σ+g state. The corresponding emission cross section is then given by

Equation (6)

Since transition probabilities of npσ1Σ+u and npπ1Πu states to the excited singlet-gerade (such as the EF1Σ+g and GK1Σ+g) states are negligible when compared with those to the X1Σ+g state, the emission cross sections of the npσ1Σ+u and npπ1Πu − X1Σ+g band systems can be considered identical to the emission cross sections of the npσ1Σ+u and npπ1Πu states, given by Equation (6).

The present model utilizes the B1Σ+u, C1Πu, and D1Πu − X1Σ+g transition probabilities of Abgrall et al. (1994) and the presently calculated adiabatic npσ1Σ+u and npπ1Πu − X1Σ+g transition probabilities of the higher Rydberg n ⩾  4 states. Significant coupling exists among some of the vj levels of the B'1Σ+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, 5pσ1Σ+u, D1Π+u, and D'1Π+u states. The calculated nonadiabatic transition probabilities are used for these levels. Wherever possible, the experimentally measured term values of Roncin & Launay (1994), Dabrowski (1984), and Takezawa (1970) are utilized to calculate the transition frequencies. To compare with the observed spectrum, the calculated spectrum is convoluted with a triangular instrument function with an appropriate FWHM.

3.2. Calculation of the npσ1Σ+u and npπ1Πu − X1Σ+g Transition Probabilities

3.2.1. Adiabatic Calculation of Higher npσ1Σu+ and npπ1ΠuX1Σ+g Band Systems

The potential curves of the B1Σ+u, B'1Σ+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, and 5pσ1Σ+u states (Staszewska & Wolniewicz 2002), the C1Πu, D1Πu, and D'1Πu states (Wolniewicz & Staszewska 2003b), and the X1Σ+g state (Wolniewicz 1993), are known theoretically with adiabatic and nonadiabatic corrections. The dipole transition moment functions of the transitions to the X1Σ+g state have been tabulated by Wolniewicz & Staszewska (2003a, 2003b). For higher members of the Rydberg series, the ab initio potential energy curves are unavailable. The quantum defect theory developed by Jungen & Atabek (1977) allows the determination of the Born–Oppenheimer potential energy with large adiabatic corrections through the classical formula

Equation (7)

where δ is the quantum defect, and $V_{H_2^+}(R)$ is the potential energy curve of H2 X2Σ+g, to which the Rydberg series converge.

When the calculated dipole transition moment function, DnΛ(R), is not available, it can be approximated from the known function of the 4pΛ− X transition, using the quantum defect δΛ:

Equation (8)

Numerical integration using the Numerov method is employed to solve the Schrödinger equation to obtain eigenvalues and eigenfunctions. The spontaneous transition probability, A(vj, vi; Jj, Ji), is given by (Hilborn 1982)

Equation (9)

where ωij$\big[\!\!=\!\!\big(E_{v_j, J_j} - E_{v_i, J_i}\big)\big/\hbar\big]$ is the angular frequency for the transition ji, χ(R) is the vibrational wavefunction with the rotational energy correction, and $\mathcal{H}_{ji}(J_j,J_i)$ is the Hönl–London factor as defined by Hansson & Watson (2005).

Equation (9) can be rewritten as

Equation (10)

where A(vj, vi; Jj, Ji) is in s−1, $\big(E_{v_j, J_j} - E_{v_i, J_i}\big)$ is in hartree, and the dipole transition moment function, D(R) is in bohr.

3.2.2. Nonadiabatic Calculation of the B'1Σ+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, 5pσ1Σ+u, D1Π+u, and D'1Π+u − X1Σ+g Band Systems

The small mass of the H2 often prevents a satisfactory reproduction of the experimental measurement using the Born–Oppenheimer approximation even if the adiabatic and diagonal nonadiabatic coupling corrections are applied. Coupling between the different electronic states must be included. While nonadiabatic coupling often appears as positional shifts of observed energy levels, the deviation of spectral intensities is even more apparent. Nonadiabatic coupling is of two types: rotational interaction mixing the 1Σ+u and 1Π+u states, and vibrational interaction mixing states with the same symmetry. Following the work and notation of Wolniewicz et al. (2006), the matrix elements between two coupling states for rotational mixing can be written:

Equation (11)

Likewise, for vibrational coupling,

Equation (12)

These matrix elements for the B'1Σ+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, D1Πu, and D'1Πu states have been tabulated as functions of the internuclear distance R by Wolniewicz et al. (2006).

In the present work, nonadiabatic coupling is assumed to be so weak that it can be treated as a perturbation. Consequently, the coupling matrix elements can be evaluated in an adiabatic vibrational basis set. The present adiabatic vibrational basis set, having 122 terms, includes the full vibrational progressions of the B'1Σ+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, 5pσ1Σ+u, D1Π+u, and D'1Π+u states. Distinguishing between 1Σ+u and 1Π+u symmetry, the nonadiabatic eigenfunction, Φj, expressed in terms of the adiabatic vibrational basis, ϕΛk, is

Equation (13)

The transition probability from the level j of the singlet-ungerade state to the level i of the X1Σ+g state is then

Equation (14)

where Dji is the dipole matrix element of the transition from level j to the ground state i and is given by

Equation (15)

In Equation (14) epsilon is 1 for a P-branch transition and −1 for an R-branch transition (Vigué et al. 1983; Glass-Maujean & Beswick 1989).

The dipole moment functions of the D1ΠuX1Σ+g, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ − X1Σ+g, and 5pσ1Σ+u − X1Σ+g band systems have been calculated and tabulated by Wolniewicz & Staszewska (2003a, 2003b).

The absorption line oscillator strength, f(vi, vj; Ji, Jj), of Equations (3) and (4) is related to the line transition probability, A(vj, vi; Jj, Ji), of Equations (9) and (14) by Abgrall & Roueff (2006)

Equation (16)

where v(vi, vj; Ji, Jj) refers to the transition wavenumber in reciprocal centimeters.

4. ANALYSIS

4.1. Predissociation and Autoionization

In addition to spontaneous emission, nonradiative processes such as predissociation and autoionization occur for some levels. Figure 1 shows the experimental energy term values of the lowest Jj levels of some npσ1Σ+u(vj) and npπ1Π+u(vj) states, along with the positions of the H(1s)+H(2ℓ) dissociation limit and the first H2 ionization potential. In general, any npσ1Σ+u or npπ1Π+u levels above the dissociation limit can be predissociative. Any npσ1Σ+u or npπ1Πu levels above the ionization potential can autoionize.

Figure 1.

Figure 1. Experimental energy term values for Jj = 0 of the npσ1Σ+u(vj) state and Jj = 1 of the npπ1Π+u(vj) state. Note that the inner-well vibrational quantum number is used for the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ state. The positions of the H(1s)+H(2ℓ) continuum and H2 first ionization potential (Jj = 0) are indicated by the dotted lines. The H(1s)+H(3ℓ) and H(1s)+H(4ℓ) continua, at 133610.35 and 138941.96 cm−1, respectively, are beyond the scale of the figure. All energy values are relative to the Ji = 0 and vj = 0 level of the X1Σ+g state.

Standard image High-resolution image

The extensive experimental and theoretical work of Glass-Maujean (1986) and coworkers (Glass-Maujean et al. 1978, 1979, 1984, 1985a, 1987) have shown that the predissociation of the npσu1Σ+u (n > 3) and npπu 1Πu (n ⩾  3) states primarily takes place via coupling to the continuum levels of the B'1Σ+u state. The rate of predissociation differs drastically depending on the orbit symmetries and relative energy separations. For instance, the D1Π+u and B'1Σ+u states are strongly coupled by Coriolis interaction. Thus, the D1Π+u rovibrational levels that lie above the H(1s)+H(2ℓ) limit are predissociated very rapidly. The lifetimes of the Jj = 2 of the vj = 3 − 11 levels of the D1Π+u state were determined to be (3.7−5.9) × 10−13 s (Glass-Maujean et al. 1979, 1985b), which are 10,000 times shorter than their expected fluorescence lifetime, (2−4) × 10−9 s (Glass-Maujean et al. 1985c; Abgrall et al. 1994). In contrast, the D1Πu state and other higher npπ1Πu states are not coupled to the B'1Σ+u or other npσ 1Σ+u states. These states can only couple to a dissociating 1Πu state. For the npπ1Πu states below the H(1s)+H(n = 3) limit, C1Πu is the only dissociative 1Πu state. Since npπu1Πu states are only weakly coupled to the C1Πu state, their predissociation rates are negligibly small. Above the H(1s)+H(n = 3) limit, npπ 1Πu levels can also be dissociated by the D1Πu continuum. Glass-Maujean et al. (2007c) recently found that the coupling between the D'1Πu state and D1Πu continuum is fairly efficient. Moreover, the predissociation of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and npσ 1Σ+u (n > 4) states, in general, takes place by homogeneous coupling with the B'1Σ+u continuum levels (Glass-Maujean et al. 1978). Due to the difference in the npσ 1Σ+u − B'1Σ+u continuum Franck–Condon overlap integrals, variations in the predissociation rates are expected. The predissociation of the npπ 1Π+u (n > 3) arises from either npπ1Π+u − D1Π+u homogeneous coupling followed by D1Π+u − B'1Σ+u Coriolis coupling or npπ1Π+u − npσ 1Σ+u Coriolis coupling followed by npσ 1Σ+u − B'1Σ+u homogeneous coupling (Glass-Maujean 1979). Levels below the H(1s)+H(2ℓ) limit cannot be predissociated. In particular, the vj = 0 levels of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and D'1Πu states lie below the limit, and spontaneous emission is the only decay mechanism.

The ionization energy of the vi = 0 and Ji = 0 level of the X1Σ+g state of H2 is 124,417.507 cm−1. For the low Jj levels of the of $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and D'1Πu states, autoionization is energetically not possible unless vj is ⩾4 for the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ state and vj⩾ 3 for the D'1Πu state (see Figure 1). Furthermore, vibrational autoionization of H2 has a tendency to proceed with a small change in vibrational quantum number (i.e., Δv = v+vj) and the autoionization rate rapidly decreases for processes with large Δv changes (Dehmer & Chupka 1976; O'Halloran et al. 1988, 1989). Autoionization of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ state requires a fairly large change in Δv and its efficiency is low (⩽ 5%; Dehmer & Chupka 1976). Autoionization rates for the vj > 4 levels of the D'1Πu state is significant and the efficiency for the low Jj levels of these bands has been measured by Dehmer & Chupka (1976).

4.2. Spectral Analysis

The examination of the relative accuracy of the calculated transition probabilities rests on the relationship of the photoemission intensity to the emission-branching ratio via Equation (1) and the oscillator strength via Equation (3). The first set of spectra with an H2 foreground column density of (2.3± 0.6) × 1013 cm−2 is optically thin for all except a few strong Werner-band resonance transitions. In principle, it gives accurate relative intensities. In practice, spectral intensities and signal-to-noise ratios for most transitions on the blue side of 850 Å are too weak for reliable relative intensity measurement. In the present analysis, examination of the relative accuracy of the transition probabilities and derivation of nonradiative yields of the npσ 1Σ+u (n > 3) and npπ 1Πu (n > 3) states are primarily carried out in the 790−900 Å region of the second set of spectra, although transitions in 900−1180 Å region of the first set are also utilized to confirm the derived nonradiative yields. The second set of spectra has a foreground column density of ∼15 × 1013 cm−2. The largest self-absorption in the 790−900 Å region is ∼28%, for the Q(1) and R(1) transitions of the D1Πu(2)−X1Σ+g(0) band. Comparison of relative intensities between the first and second sets of spectra in the (2, 0) and (1, 0) band regions of the D1ΠuX1Σ+g band system verified that the self-absorption model described in Jonin et al. (2000) reliably accounts for the intensity.

Analysis was initiated by adding the D'1Πu − X1Σ+g transition to an existing model of the B1Σ+uX1Σ+g, C1ΠuX1Σ+g, B'1Σ+uX1Σ+g, and D1ΠuX1Σ+g band systems (Jonin et al. 2000). Since predissociation of the D'1Πu levels below the H(1s)+H(n = 3) dissociation limit is negligibly small, only autoionization was considered. Figure 1 shows that autoionization of D'1Πu is negligible for vj⩽ 3 vibrational levels. So, a comparison of the relative spectral intensity of calculated and observed D'1Πu − X1Σ+g transitions can be performed straightforwardly. It was found that the calculated spectrum can reproduce the relative intensities of observed D'1Πu − X1Σ+g transitions from 790 Å to 1100 Å if the emission yields of 0.62, 0.62, and 0.35 are applied to the Jj = 1, 2, and 3 levels of the D'1Πu(4) state and the calculated adiabatic transition probabilities of the Q(1) transitions are reduced by 48%, as suggested by the measurement of Glass-Maujean et al. (2007c). Since the autoionization yield of the Q(1) transition is 8%–17% (Dehmer & Chupka 1976; M. Glass-Maujean et al. 2008, in preparation), the emission yield implies a dissociation yield of 0.2–0.3 for the Jj = 1 level of the D'1Πu(4) state, which is consistent with an upper limit of 0.3 obtained by Glass-Maujean et al. (1987). In particular, emission in the 790−850 Å wavelength region is dominated by the Q-branch transitions of the D1Πu and D'1Πu states. The experimental spectrum provides a good test of consistency among the transition probabilities of the D1ΠuX1Σ+g and D'1ΠuX1Σ+g band systems. The good agreement in the relative intensities between the model and observed spectra shows that the calculated transition probabilities of the D'1Πu − X1Σ+g band system are consistent with those of D1ΠuX1Σ+g.

Emission features on the blue side of 750−790 Å are generally very weak, and can only be practically measured with 80 μm slit widths (∼0.24 Å resolution). Most of the lines are from the Q-branch lines of the D1ΠuX1Σ+g and D'1ΠuX1Σ+g bands. The line at ∼784.04 Å, which arises from the Q(1) transition of the D'1Πu(5)−X1Σ+g(0) band, is the strongest feature in the 750–790 Å region. As noted by Liu et al. (2000), the observed emission intensities of the Q(2) and Q(3) transitions are much weaker than those expected from a simple population difference of Ji = 1, 2, and 3 levels at room temperature. A recent combination of absorption, ionization, and dissociation measurements by M. Glass-Maujean et al. (2008, in preparation) have shown a sharp increase in ionization efficiency going from the Jj = 1 to the 2 and 3 levels. The R(1) and P(3) transitions of the D'1Π+u(5) level, with a 7%–8% emission yield, are weak but observable.

The establishment of transition probabilities for the D'1ΠuX1Σ+g band allows the determination of variation of the overall relative sensitivity of the spectrometer. The inclusion of the D'1ΠuX1Σ+g emission in the model spectrum and utilization of a higher-resolution experimental spectrum (0.095 Å versus 0.115 Å) with significantly better signal-to-noise ratio produced a somewhat more accurate and flatter sensitivity curve in the 800–850 Å region than that reported by Jonin et al. (2000). After the improved sensitivity curve is applied to the observed spectrum, the Q-branch transitions of the 5pπ1Πu, 6pπ1Πu, 7pπ1Πu, 8pπ1Πu, and 10pπ1Πu states are introduced into the model. Glass-Maujean et al. (2007c) have shown that the measured transition-probability values of a number of Q-branch lines differ significantly from calculated values. Except for the Q(1) transitions of D1Πu(6) and D'1Πu(3) levels, the transition probabilities of all other Q-branch lines used in the model have been adjusted to their experimental values. The P- and R-branch transitions of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, 5pσ1Σ+u, 6pσ1Σ+u, 7pσ1Σ+u, and npπ1Π+u (n = 4–8) state are then added into the model. Because of the adiabatic nature of the calculation for many of these states, the transition probabilities of the P- and R-branches are essentially obtained from the calculated band transition probabilities partitioned by Hönl–London factors. Such an approximation is not expected to be valid in the presence of rovibronic coupling. For low vj levels of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and D'1Π+u states and all discrete vj level of the B'1Σ+u and D1Π+u states, the nonadiabatic calculation based on the ab initio results of Wolniewicz et al. (2006) described in Section 3.2.2 was used (see Section 5.1). The emission yields of some rovibrational levels of these states can be determined by comparing synthetic and calibrated experimental spectra.

5. RESULTS

Except for the few isolated regions noted below, model spectra in the region above 900 Å obtained using the Abgrall et al. (1993a, 1993b, 1993c, 1994, 1997) B1Σ+u, C1Πu, B'1Σ+u, and D1ΠuX1Σ+g transition probabilities agrees with experimental observation (Jonin et al. 2000). In the "strong" emission regions of the vj = 0 and 1 levels of the B'1Σ+u state, the calculated intensities, obtained by considering only direct excitation, are 20%−35% weaker than their experimental counterparts. As noted by Jonin et al. (2000), these regions include 976−982 Å of the (0, 2) band, 1012−1018 Å of the (0, 3) band, 1029−1035 Å of the (1, 4) band, and 1064−1070 Å of the (1, 5) band. The Liu et al. (2002) time-resolved measurements have shown that the preferential cascade excitation of the vj = 0 and 1 levels of the B'1Σ+u state via higher singlet-gerade states is at least partially responsible for the enhancement in the experimental intensities. Good agreement between the calculated and measured intensities is obtained using 35% and 20% of direct excitation rates for the cascade excitation rates to the vj = 0 and 1 levels. In comparision with the n = 2 and 3 states, transitions from n ⩾  4npσ1Σ+u and npπ1Πu states in the 900−1050 Å region are generally weak but noticeable. The use of additional transition probabilities of the higher states, adiabatic or nonadiabatic, naturally leads to better agreement between model and experimental spectra than those shown in Figure 3 of Jonin et al. (2000).

Figure 2 compares the experimental and model spectra in the range 790–850 Å. The model utilizes the transition probabilities of the B1Σ+u, C1Πu, and D1Πu − X1Σ+g band systems calculated by Abgrall et al. (1993a, 1993b, 1993c, 1994), nonadiabatic transition probabilities of the B'1Σ+u, D1Π+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, 5pσ1Σ+u, and D'1Π+uX1Σ+g band systems, and adiabatic transition probabilities of the 6pσ1Σ+u, 7pσ1Σ+u, D'1Πu, D''1Πu, 7pπ1Πu, and 8pπ1Πu − X1Σ+g band systems obtained in the present work. Transitions involving higher Rydberg (n ⩾  9) states are neglected in the model. Spectral assignments for some transitions, including those neglected by the model, are indicated. Question marks in the figure indicate that assignment has not been positively established.

Figure 2.
Standard image High-resolution image
Figure 2.
Standard image High-resolution image
Figure 2.

Figure 2. Comparison of experimental (solid trace) and model (dot trace) spectra in the range 790 Å to 850 Å. The model uses transition probabilities of the B1Σ+u, C1Πu, and D1Πu − X1Σ+g band systems calculated by Abgrall et al. (1993a, 1993b, 1993c, 1994), nonadiabatic transition probabilities of the B'1Σ+u, D1Π+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, 5pσ1Σ+u, and D'1Π+u − X1Σ+g band systems, and adiabatic transition probabilities of the 6pσ1Σ+u, 7pσ1Σ+u, D'1Πu, D''1Πu, 7pπ1Πu, and 8pπ1Πu − X1Σ+g band systems calculated in the present work. Spectral assignments of some transtions are indicated.

Standard image High-resolution image

5.1. Nonadiabatic Coupling of the B'1Σ+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, D1Π+u, and D'1Π+u States

Nonadiabatic coupling among B'1Σ+u(4), $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0), D1Π+u(2), and D'1Π+u(0) levels results in significant differences in the calculated spectrum. Figure 3 compares the relative intensities of experimental and calculated spectra in the region of B'1Σ+u(4), $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0), and D1Πu(2)−X1Σ+g(0) transitions. Although experimental spectrum (solid trace) has been calibrated using the procedures described by Liu et al. (1995) and Jonin et al. (2000), the variation of instrumental sensitivity, which is less than 1.7% over the region shown, is completely negligible. The model spectrum (dot trace, Model 1) in the top panel of Figure 3 was calculated from the B1Σ+u, C1Πu, B'1Σ+u, and D1ΠuX1Σ+g transition probabilities of Abgrall et al. (1994) and the present adiabatic transition probabilities of higher (n ⩾  4) npσ1Σ+u and npπ1Πu series. Apart from the B'1Σ+u − D1Π+u coupling that was taken into account in the calculation of Abgrall et al. (1994), coupling among B'1Σ+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, D1Π+u, and D'1Π+u levels was not considered.

Figure 3.

Figure 3. Comparison of observed (solid trace) and calculated (dot trace) spectra near the D'1Πu(2), B'1Σ+u(4), and $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0) − X1Σ+g(0) band transition region. The calculated spectrum (Model 1) in the top panel was calculated with the B'1Σ+uX1Σ+g and D1ΠuX1Σ+g transition probabilities of Abgrall et al. (1994) and adiabatic transition probabilities of the present work. Except for the partial B'1Σ+u − D1Π+u interaction, nonadiabatic coupling among the B'1Σ+u, D1Π+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, D'1Π+u, and 5pσ1Σ+u states was neglected in the top panel. The dot trace in the bottom panel (Model 2) was obtained identically except for the use of the nonadiabatic transition probabilities for the B'1Σ+u, D1Π+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, 5pσ1Σ+u, and D'1Π+uX1Σ+g transitions. Transitions are labeled as Ji(vj, viJβ, where i and j refer to the lower and upper states, β is electronic designation of singlet-ungerade states, and ΔJ = −1, 0, and +1 correspond to P, Q, and R transitions, respectively.

Standard image High-resolution image

Transitions in Figure 3 are labeled in terms of Ji(vj, viJ β, where i and j refer to the lower and upper states, β is the electronic designation of singlet-ungerade states, and ΔJ = −1, 0 and +1 correspond to P, Q, and R transitions, respectively. The ab initio calculation by Wolniewicz et al. (2006) indicates that the eigenvalues of the Jj = 1 levels of the B'1Σ+u(4) and $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0) states in Abgrall et al. (1994) and Takezawa (1970) need to be interchanged. The model calculation and labeling shown in the figure conform to the indication. The alternative assignment would result in more significant difference at the P(2) and R(0) transitions. In any case, the top panel of Figure 3 shows that the adiabatic model generally underestimates the intensity of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0) level. Moreover, the relative intensity between the R(Jj − 1) and P(Jj + 1) transitions also differs significantly from experiment. Similar discrepancy is apparent in the spectral regions involving transitions to the vi = 1 and 2 levels of the X1Σ+g state. The calculated intensities for the R(1) line of the D'1Π+u(0)−X1Σ+g(1) band at 879.13 Å (not shown) is also too strong.

The bottom panel of Figure 3 compares the observed spectrum (solid trace) with model spectrum obtained from nonadiabatic coupling calculation (dot trace, Model 2). Specifically, the transition probabilities of the P and R branches of the B'1Σ+u, D1Π+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, D'1Π+u, and 5pσ1Σ+u states, used in the top panel, have been replaced by their counterparts obtained from nonadiabatic calculation. The use of nonadiabatic transition probabilities clearly results in better agreement between the calculated and observed spectra. Except for the R(0) and P(2) lines of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0) and B'1Σ+u(4) levels, and the R(1) and P(3) lines from of the D1Πu(2) level, all other discrepancies shown in the top panel have been removed. The disagreement between the nonadiabatic model and observation in the R(1) and P(3) lines of the D1Πu(2)−X1Σ+g(2) band, and the R(0), R(1), and P(2) lines of the B'1Σ+u(4) and $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0)−X1Σ+g(1) bands, while reduced significantly, is still larger than the experimental error. The nonadiabatic transition probabilities also give rise to much better agreement for the R(1) transition of the D'1Πu(0)−X1Σ+g(1) band.

5.2. Emission Yields

The top entries of the Tables 1 and 2 list predissociation yields for the Jj = 0–3 of vj = 1–4 levels of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and Jj = 1–3 of vj = 1–5 levels of the D'1Π+u states. In Table 1, vj refers to the vibrational quantum number of the B''1Σ+u (inner well) state. The vj = 4 level of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ state lies above the first ionization potential of H2. While autoionization is possible, the autoionization rate is apparently negligibly slow in comparison with predissociation (Dehmer & Chupka 1976). The vj = 0 levels of both $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and D'1Πu states are below the B'1Σ+u continuum; their emission yields are unity. The error in the predissociation yields in Tables 1 and 2 are estimated to be ∼8% (i.e., ± 0.08). Within the experimental error, the emission yields of the vj = 1, 2, and 3 levels of the D'1Πu state are found to be near unity, consistent with the negligible predissociation yields determined by Glass-Maujean et al. (1987). Predissociation yields for other higher rovibrational levels of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and D'1Π+u state cannot be reliably determined in the present work. In the final analysis, the emission yields for these higher levels used in the synthetic spectrum were calculated from autoionization yields obtained by Dehmer & Chupka (1976) and predissociation yields reported by Glass-Maujean et al. (1987) or from the Glass-Maujean et al. (2007a, 2007b, 2007c, 2008a, 2008b) line-width measurements. The transition from the D'1Πu and the inner well of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ states to the lower excited singlet-gerade states and the continuum levels of the X1Σ+g are very weak. Predissociation yields for the levels listed in these tables are thus derived considering only the contribution of the discrete transition to the X1Σ+g state to the total transition probability, A(vj, Jj), of Equation (1).

Table 1. Nonradiative Yields of Some Rovibrational Levels of the npσ1Σ+u Statesa,b

Jj vj = 0 vj = 1 vj = 2 vj = 3 vj = 4
B''1Σ+u   0   0.90(>0.8) >0.98 (>0.8) 0.93 (>0.8) >0.95(>0.8)
 1   0.75(>0.5) >0.97(>0.7) 0.95(>0.8) >0.95(⋅⋅⋅)
 2   0.85(>0.6) >0.95 (>0.7) 0.95 (⋅⋅⋅) >0.93 (>0.9)
 3   0.80(>0.6) >0.92 (⋅⋅⋅) ⋅⋅⋅(>0.7) >0.95(⋅⋅⋅)
5pσ1Σ+u  0  0.5(0.65± 0.15) 0.15(⋅⋅⋅)  0.95(1.0± 0.1)e >0.98(1.0± 0.1)e  
1 0.35(0.50± 0.15)   0.3(0.3± 0.1) >0.90(1.0± 0.1)e >0.99(>0.6)  
2 0.80(0.65± 0.15) 0.30c(0.6± 0.1)   0.98(1.0± 0.1)e >0.95(1.0± 0.1)e  
3 0.55(>0.5)   0.3(0.5± 0.2) >0.93(1.0± 0.1)e >0.95(1.0± 0.1)e  
4 ⋅⋅⋅ ⋅⋅⋅    0.9(1.0± 0.1)e ⋅⋅⋅  
6pσ1Σ+u   0 0d 0.55d(0.5± 0.5)      
1 0d(<0.1) 0.50d(0.55± 0.15)      
2 0d(<0.1) 0.55d(0.55± 0.15)      
3 0.l5d(0.2± 0.1) 0.60d(0.6± 0.3)      
4 0.20 0.65      

Notes. aThe estimated error limit for the present yield is 8% (i.e., ± 0.08). Note the vj of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ state refers to the vibrational quantum number of the inner well (B''1Σ+u) state. bUnless noted otherwise, values in parentheses refer to predissociation yields obtained by Glass-Maujean et al. (1987). cSee Section 5.2 for the cause of the large difference between the two sets of data. dObtained after the adjustments have been made on the calculated P- and R-branch transition probabilities to be consistent with observed relative emission intensities. These levels are perturbed, see Section 5.2. At the present time, the nonadiabatic perturbations of these levels cannot be calculated but they are estimated to be very strong. eFrom Glass-Maujean et al. (2008a).

Download table as:  ASCIITypeset image

Table 2. Nonradiative Yields of Some Rovibrational Levels of the npπ1Πu Statea,b

Jj   vj = 0 vj = 1 vj = 2 vj = 3 vj = 4 vj = 5
D'1Π+u 1   0.88(0.65± 0.15)  0.78(0.6± 0.1) 0.92(0.82± 0.05) >0.97(>0.82)  >0.95(0.93± 0.05)
  2   0.88(⋅⋅⋅)  0.93(0.85± 0.05) 0.95(0.88± 0.05) >0.97(>0.89)   0.92(0.88± 0.10)
  3   0.80(0.9± 0.1)  0.95(0.95± 0.05) 0.90 (0.88± 0.08) >0.9(>0.74± 0.1d) >0.95(>0.5± 0.2d)
  4   >0.70 ⋅⋅⋅ >0.87 >0.91 >0.96
D'1Πuc 1         0.38(0.20d) 0.14(0.14)
  2         0.38(0.26d) >0.87(0.93d)
  3         >0.65(0.68d) >0.91(0.98d)
D''1Π+u 1 0.15(<0.15) 0.4(0.4± 0.2) >0.98(1.00) >0.92(0.97)    
  2 0.2(<0.1) 0.8(0.7± 0.2) >0.98(1.00± 0.05) >0.95(0.95)    
  3 0.13 0.55 >0.98 >0.75    
D''1Πu 1 0.05(<0.03) 0.06(<0.10± 0.05) >0.98(⩽ 0.99) 0.85    
  2 0.15(<0.3) 0.0 >0.95 >0.4    
  3 ⋅⋅⋅ ⋅⋅⋅ >0.95 >0.85    
6pπ1Π+u 1 0.7(0.5± 0.2) 0.3        
  2 0.6  0.3(0.4± 0.1)        
6pπ1Πu 1 0.05(<0.1) 0.24(0.3± 0.1)        
  2 0.10(<0.2) ⩽ 0.05(<0.15)        
7pπ1Π+u 1 0.5 >0.98        
  2 0.7 >0.98        
7pπ1Πu 1 0.0 >0.98        
  2 ⋅⋅⋅ >0.95        
  3 ⋅⋅⋅ >0.95        

Notes. aThe estimated error in the yields is 8% (i.e., ± 0.08), except for the vj = 4 and 5 levels of the D'1Π+u state, which is 12%. bWhen autoionization is energetically impossible, values in parentheses represent the predissociation yields of Glass-Maujean et al. (1987). When autoionizatin is possible, they denote the sum of the predissociation yields of Glass-Maujean et al. (1987) and autoionization yields of Dehmer & Chupka (1976). cEmission yields of the vj = 0 − 3 levels of the D'1Πu state are unity within experimental error. dFrom Glass-Maujean et al. (2008b).

Download table as:  ASCIITypeset image

Nonradiative yields for other higher Rydberg states were obtained after the yields of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and D'1Πu were established. Perturbations among the Rydberg series results in significant differences between the measured and calculated transition probabilities of many Q-branch excitations (Glass-Maujean et al. 2007c, 2008b). The calculated transition probabilities for the perturbed levels used in the model are adjusted to the measured values with the assumption that emission-branching ratios are insignificantly affected by the perturbation. Additionally, the transition probabilities of the P and R branches of the higher Rydberg states are generated from those of the Q branches with Hönl–London factors, even though perturbation will invariably introduce significant deviation. In absence of the nonadiabatic calculation, these two assumptions are necessary. The emission yields, obtained by assuming that the utilized transition probabilities are accurate, may thus deviate somewhat from the real values. Nevertheless, the fact that the most of the nonradiative yields in Tables 1 and 2 agree with the measured predissociation yields of Glass-Maujean et al. (1987) and the autoionization yields of Dehmer & Chupka (1976) indicates that the present approach is valid. More importantly, the use of the present emission yields and adjusted transition probabilities leads to the correct model intensities.

Predissociation yields of many rovibrational levels of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, D'1Πu, and many other higher npσ1Σ+u and npπ1Πu states have also been determined by Glass-Maujean et al. (1987) from a simultaneous experimental measurement of H2 absorption and H Lyman-α excitation spectra, and from measurements of Fano profiles of absorption transitions. For comparison purposes, their predissociation yields are listed in parentheses in Tables 1 and 2. It can be noted that Glass-Maujean et al. (1987) were only able to measure the lower limits of the predissociation yields of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ state. In this sense, the present study has obtained more definitive predissociation yields for the vj = 1 to 4 levels of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ state. However, the two sets of values, in general, agree within error limits, though a slight difference can be noted for the Jj = 1 and vj = 1 level of the D'1Π+u state (0.88± 0.08 versus 0.65± 0.15). The predissociation yield for the Jj = 2 of the 5pσ1Σ+u(1) is also significantly different from that given by Glass-Maujean et al. (1987). The difference, however, arises from problems in both the adiabatic and nonadiabatic calculations. Based on the predissociation yield of Glass-Maujean et al. (1987) and the calculated P(3) and R(1) branch transition probabilities, the model shows that the calculated emission intensities of the R(1) transition are too weak in a number of locations (e.g., 820.44 Å, 849.39 Å, 878.65 Å, 908.09 Å, and 937.57 Å) while those of the P(3) transitions roughly agree with observation. Based on Hönl–London factors, the P(3) transitions should be approximately 50% stronger than the R(1) transitions. The nonadiabatic calculation also predicts stronger P(3) branch transitions to vi = 0 − 5 levels of the X1Σ+g state. The observed intensities of the R(1) transitions in many of these bands, however, are stronger than the corresponding P(3) lines. It is unclear which perturbing state is responsible for the deviation in P- and R-branch relative intensities. The nearest known npσ1Σ+u state is the 7pσ1Σ+u(0) level, which is about 249 cm−1 higher in energy than the Jj = 2 level of the 5pσ1Σ+u(1) state. The R/P branches relative intensities of the 7pσ1Σ+u(0)−X1Σ+g transitions are also significantly stronger than those implied by Hönl–London factors (see below). Simple homogeneous coupling between the 5pσ1Σ+u(1) and 7pσ1Σ+u(0) states cannot increases the R-branch intensities of both states at the expense of the P branch. The deviation of the R/P branch relative intensities requires one or more interacting npπ1Πu levels such as 5pπ1Πu and 7pπ1Πu states. In any case, the intensities of the P(3) and R(1) lines of the 5pσ1Σ+u(1)−X1Σ+g bands can be approximately reproduced by partitioning their transition probabilities in a 2:3 ratio. The predissociation yield for Jj = 2 of the 5pσ1Σ+u(1) state is correspondingly lowered to ∼30%. The measured predissociation yield, 0.6± 0.1, is obviously more accurate than the derived predissociation yield. The large difference between the two results shows that neither absolute band transition probabilities nor P/R branch relative values are calculated reliably.

In addition to the 5pσ1Σ+u(1) level, the R(Ji−1)/P(Ji+1) relative intensities of the 6pσ1Σ+u(0), 6pσ1Σ+u(1), and 7pσ1Σ+u(0)−X1Σ+g transitions are also abnormally strong. The emission originating from the 5pσ1Σ+u(1), 6pσ1Σ+u(0), and 6pσ1Σ+u(1) levels can be reconciled with the essential features of observed spectrum by repartitioning the P- and R-branch transition probabilities without a change in the calculated band transition probabilities. For the 7pσ1Σ+u(0) level, however, the band transition probabilities need to increase by a factor of ∼2.5, in addition to the adjustment of the P- and R-branch Hönl–London factors. Even with the adjustment in band transition probabilities, intensities of a few 7pσ1Σ+u(0)−X1Σ+g transitions are not well reproduced. In any case, while the predissociation yields obtained with these adjustments are not expected to be reliable, the inferred values for the 6pσ1Σ+u(0), 6pσ1Σ+u(1), and 7pσ1Σ+u(0) levels are consistent with those determined by Glass-Maujean et al. (1987). All other higher vj levels of the 6pσ1Σ+u and 7pσ1Σ+u states have negligible emission yields.

5.3. Excitation and Emission Cross Sections

The calculated transition probabilities of npσ1Σ+u and npπ1Πu band systems, along with Equations (3)−(6) and collision strength coefficients of Liu et al. (1998), permit a good estimation of the cross sections of the npσ1Σ+u and npπ1Πu, especially those of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and D'1Πu states. The solid lines of Figures 4 and 5 show the estimated excitation cross section of the $B^{\prime \prime }\bar{B}\,{^{1}\Sigma _{u}^{+}}- X\,{^{1}\Sigma _{g}^{+}}$ and D'1ΠuX1Σ+g band system as a function of excitation energy. The calculation of corresponding emission cross sections requires appropriate emission yields for the rovibrational levels. For the vj = 1 − 4 levels of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ state and vj = 1 − 5 levels of the D'1Πu state, the nonradiative yields listed in Tables 1 and 2 have been applied. For higher vibrational levels, the predissociation yields listed by Glass-Maujean et al. (1987) are utilized. For some rovibrational levels, lower limits of the predissociation yields were given by Glass-Maujean et al. (1987). In these cases, lower limits are used. For autoionization, the autoionization yields reported by Dehmer & Chupka (1976) have been applied. At room temperature, emission yields for rotational levels up to the Jj = 4 are required. Alternatively, the emission yields can be estimated from the calculated spontaneous transition probabilities and photoabsorption line width measurements of Glass-Maujean et al. (2007a, 2007b, 2007c, 2008a). While the experimental uncertainties of the width, typically ± 0.3 cm−1, are somewhat too large, they can provide a rough estimate of the yields for the high vj levels when other data are unavailable. Because emissions from these levels are generally very weak, the error in line width does not lead to significant change in the band emission cross sections, at least, for the low n Rydberg states. The emission cross sections of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and D'1Πu states are indicated by dotted lines in Figures 4 and 5.

Figure 4.

Figure 4. Excitation (solid) and emission (dotted) cross sections of the $B^{\prime \prime }\bar{B}\,{^{1}\Sigma _{u}^{+}}- X\,{^{1}\Sigma _{g}^{+}}$ band system as a function of excitation energy at 300 K.

Standard image High-resolution image
Figure 5.

Figure 5. Excitation (solid) and emission (dotted) cross sections the D'1ΠuX1Σ+g band system as a function of excitation energy at T = 300 K.

Standard image High-resolution image

Table 3 compares excitation and emission cross sections of singlet-ungerade states obtained in several studies at 100 eV and 300 K. The present cross sections for the B1Σ+u and C1Πu states differ from those reported by Liu et al. (1998) in two ways. First, excitation to H(1s)+H(2ℓ) continuum levels, neglected in Liu et al. (1998), is taken into account in the present work using oscillator strengths derived from photodissociation cross sections calculated by Glass-Maujean (1986). Second, while Liu et al. (1998) used the B1Σ+uX1Σ+g and C1ΠuX1Σ+g transition probabilities of Abgrall & Roueff (1989), the present work used transition probabilities from Abgrall et al. (1993a, 1993b, 1994). The emission cross sections of the B1Σ+u, C1Πu, and B'1Σ+u states do not include emission originating from the H(1s)+H(2ℓ) continuum, but include continuum emission originating from the discrete levels of the B1Σ+u, C1Πu, and B'1Σ+u states into the X1Σ+g continuum, H(1s)+H(1s). At 100 eV and 300 K, Abgrall et al. (1997) have shown that transitions to the X1Σ+g continuum contributes about 27.5% and 1.5% to the B1Σ+u and C1Πu emission cross sections, respectively. The transition probabilities of some high vj levels of the D1Πu state have been adjusted to the experimental values of Glass-Maujean et al. (2007c). The present cross sections of the D1Πu state differs slightly from those obtained by Jonin et al. (2000). In case of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and D'1Πu states, Table 3 shows that the present cross sections are significantly larger than those estimated by Jonin et al. (2000). Possible reasons for the large difference will be discussed in Section 6.

Table 3. Electronic-Band Cross Sections and Emission Yields of H2 Singlet-ungerade Statesa

State Present σex Previous σex Present σem Previous σem Present Em. Yield Previous Em. Yield
B1Σ+u 264b 262c 263 262c 99%b 100%
C1Πu 244b 241c 240b 241c 98%b 100%
B'1Σ+u 40b 38d,e 21 21d 53% 56%
D1Π+u 25 24d 11 11d 43% 46%
D1Πu 21 18d 21 18d 100% 100%
$B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ 11 >4d 2.2 1.6d 20% <40%
D'1Π+u 9.3 7.1d 1.6 1.0d 18% 14%
D'1Πu 7.3 ⩾5.3d 5.7 5.3d 78% ⩽ 100%
D''1Πu 3.2 >0.6 0.9 0.6 28% ...
5pσ1Σ+u ... ... 1.1 ... ... ...
6pσ1Σ+u ... ... 0.6 ... ... ...
6pπ1Πu ... ... 0.9 ... ... ...
7pσ1Σ+u ... ... 0.6 ... ... ...

Notes. aE = 100 eV and T = 300 K. Unit is 10−19 cm2. σex and σem denote excitation and emission cross sections, respectively. Certain numbers may not add up due to roundings. See Section 5.3 for estimated errors in cross sections. bExcitation cross sections include the excitation into the H(1s)+H(2ℓ) continuum, which is estimated from the calculation of Glass-Maujean (1986). Emission cross sections exclude emission from the H(1s)+H(2ℓ) continuum levels, but include continuum emission from the excited discrete levels into the continuum levels of the X1Σ+g state. Transitions to the X1Σ+g continuum contribute 27.5% and 1.5%, respectively, to total emission cross sections of B1Σ+uX1Σ+g and C1ΠuX1Σ+g (Abgrall et al. 1997). cFrom Liu et al. (1998). dFrom Jonin et al. (2000). eInclude excitations into the continuum levels of the B'1Σ+u state.

Download table as:  ASCIITypeset image

The estimated errors of the electronic-state cross sections listed in Table 3 are as follows. The errors for the B1Σ+u, C1Πu, and D1Πu states, which primarily arise from the uncertainty in the excitation function, are no greater than 10%. For the B'1Σ+u and D'1Πu states, uncertainty is less than 12%–13%. The $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0)−X1Σ+g emission makes a substantial contribution to the total $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ state emission cross section. As the calculated nonadiabatic transition probabilities are unable to satisfactorily reproduce the observed relative intensities of the R(0) and P(2) of the B'1Σ+u(4) and $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0) levels, there is an additional uncertainty in the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ state cross section, approaching to 15% for the excitation cross section and ∼18% for the emission cross section. The error limit for the 5pσ1Σ+u, 6pσ1Σ+u, and 7pσ1Σ+u state emission cross sections can be as high as 25%–35% because of weak emission intensities and, more importantly, rovibronic coupling noted in Section 5.2.

6. DISCUSSION

The vj = 0 level of the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and D'1Π+u states lies below the H(1s)+H(22ℓ) dissociation limit and is, therefore, free from the predissociation of the B'1Σ+u continuum. Except for the R(1) transitions, the adiabatic transition probabilities for the D'1Π+u(0) level actually reproduces the observed relative intensities reasonably well. Very prominent discrepancies exist between the calculated and observed relative emission intensities for the B'1Σ+u(4) and $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0) states. A homogeneous coupling Hamiltonian of 5.7 cm−1 between the B'1Σ+u(4) state and $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0) state were inferred from the early absorption study of Namioka (1964b). Indeed, a simple perturbation treatment based on the Hamiltonian and adiabatic transition probabilities of the B'1Σ+u(4) and $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0) states enables one to obtain good agreement between the calculated and observed spectral intensities of the P(1) branch transitions. Because of the 1Σ+u − 1Π+u coupling, however, the simple perturbation coupling scheme fails for the transitions involving Jj > 0 levels.

The nonadiabatic calculation successfully reproduces the observed P(1) transition intensities from the B'1Σ+u(4) and $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0) levels and thus correctly accounts for the homogeneous interactions between the two vibronic levels. Except for the transitions from the Jj = 1 levels, the calculation reproduces the observed intensities of other Jj levels of the B'1Σ+u(4) and $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0) states well within the experimental errors. For the Jj = 1 levels, however, discrepancies in the R(0)- and P(2)-branch absolute intensities and the R(0)/P(2) relative intensities are beyond expected experimental error. Because the Jj = 1 levels of the B'1Σ+u(4) and $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0) are strongly coupled, respectively with ∼53% and ∼47% mixture (Wolniewicz et al. 2006), the error in the coupling matrix elements is amplified. Thus, the discrepancies in Jj = 1 spectral intensities is most likely caused by the nonperturbative nature of nonadiatic coupling, which is inadequately approximated by the present perturbative treatment. Expanding the basis set of the calculation will probably lead to better agreement with experiment. It is also interesting to note that the nonadiabatic model overestimates the spectral intensities of the R(1) and P(3) branches of the D1Πu(2)−X1Σ+g(0) band. Since the Q-branch intensities calculated from the D1ΠuX1Σ+g transition probabilities of Abgrall et al. (1994) agree with experimental values over the entire measurement range, the small discrepancies in the R(1) and P(3) branches' spectral intensity also indicate the error in the nonadiabatic treatment.

The good agreement between the present and previous predissociation yields listed in Tables 1 and 2 suggests the relative accuracy of the calculated transition probabilities. As mentioned, the predissociation yield of Glass-Maujean et al. (1987) was obtained by a simultaneous measurement of molecular hydrogen absorption and atomic hydrogen Lyman-α excitation spectra. Once the absolute scale is established, the ratio of Lyman-α emission intensity to absorption intensity directly produces the predissociation yields. So, in the absence of significant spectral overlap, predissociation yields of Glass-Maujean et al. (1987) were directly obtained from measurement and, therefore, are very accurate. Predissociation yields of the present work are determined by matching the relative intensity of synthetic spectrum with that of the observed one through the adjustment of emission-branching ratio. Specifically, the accuracy of the line intensities in synthetic spectra depends on the accuracy of emission-branching ratios and the consistency of oscillator strengths (i.e., transition probabilities) among different rovibronic excitations. The good agreement in the two sets of predissociation yield, therefore, shows the good consistency among the calculated transition probabilities of the B1Σ+u, C1Πu, B'1Σ+u, D1Πu, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, and D'1ΠuX1Σ+g band systems.

The excitation and emission cross sections of $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ and D'1Πu states at 100 eV excitation energy have been estimated by Jonin et al. (2000) based on modeling and extrapolating a number of experimental emission lines. For the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$ state at 100 eV excitation energy, Jonin et al. (2000) obtained 0.4 × 10−18 cm2 as the lower limit of the excitation cross section, 0.16 × 10−18 cm2 for the emission cross section, and 40% as the upper limit of the emission yield. Those numbers can be compared with the present values of 1.1 × 10−18 cm2 for the excitation cross section, 0.29 × 10−18 cm2 for the emission cross section, and 21% for the band emission yield (see Table 3). For the D'1Πu state at the same excitation energy, Jonin et al. (2000) derived 1.2 × 10−18 cm2, 0.63 × 10−18 cm2, and 52%, respectively, for the excitation and emission cross sections, and emission yield. These numbers can be compared with 1.7 × 10−18 cm2 and 0.70 × 10−18 cm2, and 42%, respectively, obtained in the present work. The probable causes for this significant difference are that the relative instrumental sensitivity between 800 Å and 850 Å were overestimated by Jonin et al. (2000) and that significant portions of the D'1ΠuX1Σ+g transition takes place in this spectral region. Furthermore, the extrapolation of the inaccurate instrumental sensitivity into the 760 to 800 Å region by Jonin et al. (2000) also contributed to the errors in the D'1Πu state.

The calculated transition probabilities of the npσ1Σ+u and npπ1Πu − X1Σ+g (n = 4 − 8) band systems, adiabatic or nonadiabatic, represent an important step toward accurate modeling of the electron-impact induced emission spectrum of H2 in the wavelength region below 900 Å. Liu et al. (1995), Abgrall et al. (1997), and Jonin et al. (2000) have shown that models utilizing the calculated transition probabilities for the B1Σ+uX1Σ+g, C1ΠuX1Σ+g, B'1Σ+uX1Σ+g, and D1ΠuX1Σ+g band systems by Abgrall et al. (1993a, 1993b, 1993c, 1994, 2000) can accurately reproduce experimental H2 spectral intensities between the 1040 Å and 1660 Å wavelength region. Measurement and analysis by Jonin et al. (2000) have also demonstrated that emissions from $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, D'1Πu, and higher npσ1Σ+u and npπ1Πu states are not negligible in certain spectral regions below 1040 Å. Between 900 Å and 1040 Å, emissions from the B1Σ+uX1Σ+g, C1ΠuX1Σ+g, B'1Σ+uX1Σ+g, and D1ΠuX1Σ+g band systems contribute over 95% of the total H2 emission intensity, with the $B^{\prime \prime }\bar{B}\,{^{1}\Sigma _{u}^{+}}- X\,{^{1}\Sigma _{g}^{+}}$ and D'1ΠuX1Σ+g band systems contributing another ∼3%. Jonin et al. (2000) have noted that the observed relative intensities for some (i.e., "strong") emissions from the vj = 0 and 1 levels of the B'1Σ+u state are significantly stronger than their calculated counterparts. The discrepancy was attributed to the cascade excitation of the low-vj levels of the B'1Σ+u state (Liu et al. 2002). The present work shows that the available transition probabilities can accurately account for over 99% of the molecular hydrogen spectral intensity in the 900 Å to 1040 Å wavelength region. Emissions from higher-Rydberg (n ⩾  5) states become significant on the blue side of 900 Å, with the lower B'1Σ+u, C1Πu, B'1Σ+u, D1Πu, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, and D'1Πu states contributing about 90%–92% to the total observed spectral intensities between 790 Å and 900 Å. In a few small wavelength regions, emissions from the 5pσ, 6pσ, 7pσ, 5pπ, 6π, and 7π are the dominant spectral features. The differences between the model and observation for a few transitions of the B'1Σ+u(4), D1Π+u(2), $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$(0), 6pσ1Σ+u(0), 6pσ1Σ+u(1), and 7pσ1Σ+u(0) levels are beyond experimental error, which is 8%–12% for moderately strong transitions. The overall emission intensity from these levels, however, is less than 2% of the total emission intensity between 790 Å and 900 Å. The current model is thus capable of reproducing ∼98% of e+H2 emission intensity at room temperature and 100 eV excitation energy within experimental error.

The present work makes it possible to improve the accuracy of calibration for the Cassini Ultraviolet Imaging Spectrograph (UVIS) instrument. Before this work, the relative sensitivity of the UVIS instrument in the spectral direction of the EUV region was established with the electron-impact induced emission spectrum of H2 (Shemansky & Liu 2000), based on the model and experimental spectra of Jonin et al. (2000). As the spectra of Jonin et al. (2000) is accurate only for the wavelength region above 900 Å, it was difficult to obtain an accurate calibration for the UVIS instrument in the shorter wavelength region. Indeed, the large point-spread function of UVIS actually limited the accurate calibration to the λ> 920 Å region (Shemansky & Liu 2000). The present work has not only extended the accurate calibration into the 800−920 Å region but has also improved the accuracy above 920 Å. The present results, along with those obtained in Abgrall et al. (1997) and Jonin et al. (2000), make it possible to obtain accurate relative sensitivity curves of laboratory spectrometers and Cassini UVIS instrument over the range 800 Å to 1630 Å.

Electron-impact excitation of H2 in the atmospheres of outer planets usually takes place at temperatures higher than 300 K (e.g., 500−1200 K), over a wide range of electron excitation energies. Accurately modeling the emission intensity in the EUV region obviously requires emission yields for the higher Jj levels. Nevertheless, assuming that the predissociation and autoionization yields for the high Jj levels of the B'1Σ+u and D'1Πu states are similar to those of the low Jj levels, the calculated transition probabilities of the B1Σ+u, C1Πu, B'1Σ+u, D1Πu, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, D'1Πu, and higher npσ1Σ+u and npπ1Πu − X1Σ+g band systems, along with those involved in the cascade excitation via the singlet-gerade states (Liu et al. 2002), provide basic physical parameters for an accurate interpretation of spacecraft observations of the electron-impact induced emission spectrum of H2 over the entire VUV region.

In summary, adiabatic transition probabilities of the npσ1Σ+u and npπ1Πu − X1Σ+g (n = 4 − 8), and nonadiabatic transition probabilities of the B'1Σ+u, $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, and D1Π+uX1Σ+g band systems have been obtained. The high-resolution electron impact-induced emission spectrum of H2 obtained by Jonin et al. (2000) and Liu et al. (2000) has been re-examined with presently and previously calculated transition probabilities. Adiabatic transition probabilities are found to be consistent with experimental observation when localized rovibronic coupling of the npσ1Σ+u and npπ1Πu states is insignificant. When localized coupling is significant, nonadiabatic calculations are partially successful in removing the discrepancies between model and observation. Emission yields obtained by comparison of calculated and experimental intensities also agree with the predissociation yield reported by Glass-Maujean et al. (1987) and autoionization yield by Dehmer & Chupka (1976). Refined excitation and emission cross sections for the $B^{\prime \prime }\bar{B}\,^{1}\Sigma _{u}^{+}$, D'1Πu, and several higher Rydberg states have been obtained.

The work at the Space Environment Technologies (SET) has been supported by the Cassini UVIS contract with the University of Colorado, and NASA-NNG06GH76G issued to SET through the Planetary Atmospheres Program. We thank Charles Malone for valuable suggestions.

Please wait… references are loading.
10.1088/0067-0049/180/1/38