Photodissociation of CS from Excited Rovibrational Levels

, , , , , , and

Published 2018 April 27 © 2018. The American Astronomical Society. All rights reserved.
, , Citation R. J. Pattillo et al 2018 ApJ 858 10 DOI 10.3847/1538-4357/aab5b9

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/858/1/10

Abstract

Accurate photodissociation cross sections have been computed for transitions from the X 1Σ+ ground electronic state of CS to six low-lying excited electronic states. New ab initio potential curves and transition dipole moment functions have been obtained for these computations using the multi-reference configuration interaction approach with the Davidson correction (MRCI+Q) and aug-cc-pV6Z basis sets. State-resolved cross sections have been computed for transitions from nearly the full range of rovibrational levels of the X 1Σ+ state and for photon wavelengths ranging from 500 Å to threshold. Destruction of CS via predissociation in highly excited electronic states originating from the rovibrational ground state is found to be unimportant. Photodissociation cross sections are presented for temperatures in the range between 1000 and 10,000 K, where a Boltzmann distribution of initial rovibrational levels is assumed. Applications of the current computations to various astrophysical environments are briefly discussed focusing on photodissociation rates due to the standard interstellar and blackbody radiation fields.

Export citation and abstract BibTeX RIS

1. Introduction

CS is a molecule of great astrophysical interest. It is also one of the most abundant sulfur-bearing compounds in interstellar clouds and is found in a variety of astrophysical objects including star-forming regions (Walker et al. 1986), protostellar envelopes (Herpin et al. 2012), dense interstellar clouds (Hasegawa et al. 1984; Hayashi et al. 1985; Destree et al. 2009), carbon-rich stars (Bregman et al. 1978; Ridgway et al. 1997; Tenenbaum et al. 2010), oxygen-rich stars (Ziurys et al. 2007; Tenenbaum et al. 2010), planetary nebulae (Edwards & Ziurys 2014), and comets (Smith et al. 1980; Jackson et al. 1982; Canaves et al. 2007).

Photodissociation is an important mechanism for the destruction of molecules in environments with an intense radiation field, so accurate photodissociation rates are necessary to estimate the abundance of CS. Heays et al. (2017) presented photodissociation cross sections and photorates for CS using previous estimates (van Dishoeck 1988) applying measured wavelengths for transitions to the B 1Σ+ (or 3 1Σ+) from the ground state (Stark et al. 1987) and vertical excitation energies of higher states (Bruna et al. 1975). However, comprehensive photodissociation cross sections are needed to compute photorates in many environments. In response, we have calculated photodissociation cross sections for the CS molecule for several electronic transitions from a wide range of initial rovibrational levels. Photodissociation cross sections for transitions from the X 1Σ+ electronic ground state to the A 1Π, A' 1Σ+(2 1Σ+), 2 1Π, 3 1Π, B 1Σ+(3 1Σ+), and 4 1Π electronic states are studied here. Calculations have been performed for transitions from 14,908 initial bound rovibrational levels v'', J'' of the X state. We also explore predissociation out of the 3 1Π, B 1Σ+, and 4 1Π excited electronic states.

The present cross section calculations are performed using quantum-mechanical techniques. Applications of the cross sections to environments appropriate for local thermodynamic equilibrium (LTE) conditions are included, where a Boltzmann distribution of initial rovibrational levels is assumed. Photodissociation rates are computed for the standard interstellar radiation field (ISRF) and for blackbody radiation fields at a wide range of temperatures.

The layout of this paper is as follows. An overview of the theory of molecular photodissociation and the adopted molecular data is presented in Section 2. In Section 3, the computed state-resolved cross sections, LTE cross sections, and photodissociation rates are discussed. Finally, in Section 4, conclusions are drawn from our work. Atomic units are used throughout unless otherwise specified.

2. Theory and Calculations

2.1. Potential Curves and Transition Dipole Moments

In a similar manner to our recent molecular structure work on the SiO molecule (Forrey et al. 2016; Cairnie et al. 2017), which is iso-electronic to CS, the potential energy curves and transition dipole matrix (TDM) elements for several of the low-lying electronic states are calculated. We use a state-averaged-multi-configuration-self-consistent-field (SA-MCSCF) approach, followed by multi-reference configuration interaction (MRCI) calculations together with the Davidson correction (MRCI+Q; Helgaker et al. 2000). The SA-MCSCF method is used as the reference wave function for the MRCI calculations.

Potential energy curves (PECs) and TDMs as a function of internuclear distance R are calculated starting from a bond separation of R = 1.8 Bohr extending out to R = 12 Bohr. At bond distances beyond this value we use a multipole expansion, detailed below, to represent the long-range part of the potentials. The basis sets used in our work are the augmented correlation consistent polarized sextuplet (aug-cc-pV6Z (AV6Z)) Gaussian basis sets. The use of such large basis sets is well known to recover 98% of the electron correlation effects in molecular structure calculations (Helgaker et al. 2000). All the PEC and TDM calculations for the CS molecule were performed with the quantum chemistry program package MOLPRO 2015.1 (Werner et al. 2015), running on parallel architectures.

For molecules with degenerate symmetry, an Abelian subgroup is required to be used in MOLPRO. For a diatomic molecule like CS with ${{\rm{C}}}_{\infty v}$ symmetry, it will be substituted by C2v symmetry with the order of irreducible representations being (A1, B1, B2, A2). When symmetry is reduced from ${{\rm{C}}}_{\infty v}$ to C2v, the correlating relationships are $\sigma \to {a}_{1}$, $\pi \to $ (b1, b2), and $\delta \to $ (a1, a2). In order to take account of short-range interactions, we employed the non-relativistic state-averaged complete active-space-self-consistent-field (SA-CASSCF)/MRCI method available within the MOLPRO (Werner et al. 2012, 2015) quantum chemistry suite of codes.

For the CS molecule, eight molecular orbitals (MOs) are put into the active space, including four a1, two b1, and two b2 symmetry MOs which correspond to the 3s3p shell of sulfur and 2s2p shell of carbon. The rest of the electrons in the CS molecule are put into closed-shell orbitals, including four a1, one b1 and one b2 symmetry MOs. The molecular orbitals for the MRCI procedure were obtained using the SA-MCSF method, for which we carried out the averaging processes on the lowest three 1Σ+ (1A1), three 1Π (1B1), three 3Σ+ (3A1), three 3Π (3B1), two 1Δ (1A2) and two 3Δ (3A2) molecular states. The fourteen MOs (8a1, 3b1, 3b2, 0a2), i.e., (8, 3, 3, 0), were then used to perform all the PEC and TDM calculations for the electronic states of interest in the MRCI+Q approximation. Table 1 compares theoretical results for the permanent dipole moment μX of the X 1Σ+ ground state at various levels of approximation with experiment to demonstrate the accuracy of the MRCI+Q approximation applied here. As can be seen from Table 1 our MRCI+Q results for the permanent dipole moment of CS, for the ground state, at the equilibrium geometry, are within 4% of the experimental value (Winnewisser & Cook 1968). Table 2 compares the equilibrium distance Re (Å) and the dissociation energy De (eV) at various level of approximation for the X 1Σ+, A' 1Σ+, and A 1Π states of CS. We note that for the X 1Σ+ ground state, the early experimental work of Crawford & Shurcliff (1934) determined values Re = 1.2851 Å and De (eV) = 7.752 eV, in less favorable agreement with our present ab initio work or that of other high level molecular structure calculations. As shown in Table 2 the use of polarized-core-valence basis sets by Li et al. (2013) provides spectroscopically accurate results for Re (Å) and De (eV), respectively, being within 0.03 pm and 0.032 eV compared with the available experiment. We find that our TDMs differ slightly in magnitude but agree in trend with those presented by Li et al. (2013) on the range they are computed.

Table 1.  The Permanent Dipole Moment μX for the X 1Σ+ Ground State of the CS Molecule at 2.9 a0/1.5346 Å, a Value Near Equilibrium, Compared with Experiment, SCF, CAS-CI, MRCI+Q, and MCSCF Theoretical Calculations

CS Ground State Method Basis Set μX (Debye) Δ (%)
X 1Σ+ EXPTa 1.958 ± 0.005
MRCI+Qb aug-cc-pV6Z 2.042 +4.3%
MCSCFc aug-cc-pV6Z 2.179 +11%
CAS-CId double-zeta + polarization (DZP) 2.147 +9.7%
SCFe double-zeta + polarization (DZP) 1.783 −8.9%
CIf double-zeta (DZ) 2.350 +20%
HFg double-zeta (DZ) 1.650 −16%

Notes.

aExperiment (Winnewisser & Cook 1968). bMulti-reference configuration interaction with the Davidson correction (MRCI+Q; present work). cMulti-configuration-self-consistent-field (MCSCF; present work). dComplete-active-space configuration interaction (CAS-CI), with the SWEDEN codes (present work). eSelf-consistent field (SCF; Varambhia et al. 2010). fConfiguration interaction (CI; Robbe & Schamps 1976). gHartree–Fock (HF; Robbe & Schamps 1976).

Download table as:  ASCIITypeset image

Table 2.  Equilibrium Bond Distance Re (Å) and Dissociation Energies De (eV) for the X 1Σ+, A' 1Σ+, and A 1Π Molecular States of CS from the Present MRCI+Q Calculations Compared to Other Theoretical and Experimental Results

Molecular State Method Basis Set Re De/(eV)
X 1Σ+ MRCI + Qa aug-cc-pV6Z (AV6Z) 1.5314 7.6113
  MRCI + Qb aug-cc-pwCV5Z (ACV5Z) 1.5346 7.3851
  MRCI + Qc aug-cc-pV6Z (AV6Z) 1.5346
  MRCI + Q + cv + dkd aug-cc-pV6Z (AV6Z) 1.5377
  CCSD(T) + cv + dk + 56e aug-cc-pV6Z (AV6Z) 1.5387
  MRCIf aug-cc-pC5VZ (C) + aug-cc-pV5Z (S) 1.5334 7.3436
  M-S-APEFg aug-cc-pC5VZ (C) + aug-cc-pV5Z (S) 1.5403 7.3436
  HF/DF-B3LYPh aug-cc-pVTZ 1.5360 7.0644
  EXPTi 1.5349
  EXPTj 1.5350
  EXPTk 1.5349 7.3530 ± 0.025
  MORSE/RKRl 1.5349 7.4391
 
A' 1Σ+ (2 1Σ+) MRCI + Qa aug-cc-pV6Z (AV6Z) 1.9443 0.4558
  MRCI + Qb aug-cc-pwCV5Z (ACV5Z) 1.9399 0.4253
  MRCI + Qc aug-cc-pV6Z (AV6Z) 1.9399
  EXPTi 1.9440
  EXPTj 1.9440
 
A 1Π MRCI + Qa aug-cc-pV6Z (AV6Z) 1.5622 2.7333
  MRCI + Qb aug-cc-pwCV5Z (ACV5Z) 1.5676 2.6637
  MRCI + Qc aug-cc-pV6Z (AV6Z) 1.5676
  MRCI + Q + cv + dkd aug-cc-pV6Z (AV6Z) 1.5690
  EXPTi 1.5739
  EXPTj 1.5660

Notes. The data are given in units conventional to quantum chemistry with 1 Å = 10−10 m and 0.529177 Å ≈ 1 a0. The conversion factor 1.239842 × 10−4 eV = 1 cm−1 is also used.

aMRCI+Q, Multi-reference configuration interaction (MRCI) with Davidson correction (Q; present work). bMRCI+Q, ACV5Z (Li et al. 2013). cMRCI+Q, AV6Z (Shi et al. 2011). dMRCI+Q+cv+dk, core-valence (cv) and relativistic effects (dk; Shi et al. 2011). eCCSD(T)+cv+dk+56, Coupled cluster (CCSD(T)), core-valence, relativistic effects/basis set limit (Shi et al. 2011). fMRCI (Shi et al. 2010). gM-S-APEF, Murrell–Sobbell (M–S) fit with analytic potential energy function (APEF; Shi et al. 2010). hHF/DF-B3LYP, Hybrid density functional method (Midda & Das 2003). iExperiment (Huber & Herzbeg 1979). jExperiment (Bergeman & Crossart 1981). kExperiment (Coppens & Drowart 1995). lMorse with Rydberg–Klein–Rees (RKR) potential (Nadhem et al. 2015).

Download table as:  ASCIITypeset image

Beyond a bond separation of R = 12 Bohr, a multipole expansion is smoothly fitted to the PECs and TDMs up to R = 100 Bohr. For the PECs this has the form

Equation (1)

where C5 and C6 are coefficients for each electronic state shown in Table 3. For R < Rmin, down to a bond length of 1.5 a0, a short-range interaction potential of the form $V(R)=A\exp (-{BR})+C$ was fitted to the ab initio potential curves.

Table 3.  CS Electronic States

Molecular Separated Atom United Atom
State Atomic State Energy (eV)a Energy (eV)b C5c C6d State
X 1Σ+ C(2s22p2 3P) + S(3s23p4 3P) 0.0 0.0 27.34 58.02 3d24s2a1D
A 1Π ${\rm{C}}(2{s}^{2}2{p}^{2}{}^{3}P)+{\rm{S}}(3{s}^{2}3{p}^{4}{}^{3}P)$ 0.0 7.95(−4) 0.0 58.02 3d24s2a1D
A' 1Σ+ C(2s22p2 3P) + S(3s23p4 3P) 0.0 1.14(−3) 0.0 58.02 3d24s2a1G
2 1Π C(2s22p2 3P) + S(3s23p4 3P) 0.0 6.52(−4) −18.23 58.02 3d24s2a1G
B 1Σ+ C(2s22p2 1D) + S(3s23p4 1D) 2.3812287 2.38172 27.26 55.56 3d34s b1G
3 1Π C(2s22p2 1D) + S(3s23p4 1D) 2.3812287 2.38652 10.06 55.56 3d34s b1G
4 1Π C(2s22p2 1D) + S(3s23p4 1D) 2.3812287 2.39969 −11.81 55.56 3d34s a1H

Notes.

aExperimental data from NIST Atomic Spectra Database (Kramida et al. 2016). bCurrent theory extrapolated to the asymptotic limit with Equation (1). cEstimated following Chang (1967). See the text for details. dEstimated from the London formula. See the text for details.

Download table as:  ASCIITypeset image

A method to estimate the value of the quadrupole–quadrupole coefficient C5 for an electronic state of a diatomic molecule like CS is given by Chang (1967). In order to compute the long-range dispersion coefficient C6, the London formula

Equation (2)

is applied, where α is the dipole polarizability and ${ \mathcal I }$ is the ionization energy of each of the atoms in a given atomic state. The ionization energies are taken from the NIST Atomic Spectra Database (Kramida et al. 2016). For the sulfur atom, the dipole polarizabilities of αS = 18.8 and αS = 19.5, respectively, are used for the ground state 3s23p4 3P and the excited state 3s23p4 1D (Mukherjee & Ohno 1989). For the ground state 2s22p2 3P of atomic carbon, a dipole polarizability of αC = 10.39 is used (Miller & Kelly 1972). An estimated value of αC = 10.78 is used for the excited state 2s22p2 1D of carbon: this value was obtained by scaling the ground state polarizability to match the ratio of the 3P and 1D polarizabilities of sulfur.

The potentials for the excited states of the CS molecule were shifted so that the asymptotic energies as $R\to \infty $ agree with the separated atom energy differences found in the NIST Atomic Spectra Database (Kramida et al. 2016) shown in Table 3. Except for the 4 1Π state, shifts are less than ∼5 meV indicating the reliability of the MRCI+Q calculations within the uncertainty of the estimated dispersion coefficients. The potential curves for CS are shown in Figure 1.

Figure 1.

Figure 1. Potential energy curves for each considered CS molecular state.

Standard image High-resolution image

The TDMs for the CS molecule are similarly extended to long and short-range internuclear bond distances. For R > Rmax a functional fit of the form $D(R)=a\exp (-{bR})+c$ is applied, while in the short-range R < Rmin a quadratic fit of the form D(R) = a' R2 + b' R + c' is adopted. We deduce from the atomic states of C and S that the long-range $R\to \infty $ limit of each TDM is zero. Similarly, the united-atom limit (which is the Ti atom) as R → 0 of each TDM is zero as well (see Table 3). The TDMs are shown in Figure 2.

Figure 2.

Figure 2. Transition dipole moments for transitions from the ground state to each CS excited state.

Standard image High-resolution image

The wave functions of the bound rovibrational levels are computed by solving the radial Schrödinger equation for nuclear motion on the X 1Σ+ potential curve. The wave functions are obtained numerically using the standard Numerov method (Cooley 1961; Johnson 1977) with a step size of 0.001 Bohr. We find 85 vibrational levels with a total of 14,908 rovibrational levels. This covers nearly the full range of rovibrational levels in the X 1Σ+ state.

2.2. The Photodissociation Cross Section

Here, we present a brief overview of the state-resolved photodissociation cross section calculation; further details are given in previous work (Miyake et al. 2011). In units of cm2, the state-resolved cross section for a bound–free transition from initial rovibrational level v'' J'' is

Equation (3)

(Kirby & van Dishoeck 1988), where k'J' are the continuum states of the final electronic state. The Hönl–London factors, SJ'(J'') (Watson 2008), are expressed for a ${\rm{\Sigma }}\leftarrow {\rm{\Sigma }}$ electronic transition as

Equation (4)

and for a ${\rm{\Pi }}\leftarrow {\rm{\Sigma }}$ transition as

Equation (5)

The matrix element of the electric TDM for absorption from v'' J'' to the continuum k'J' is

Equation (6)

with the integration taken over R where D(R) is the appropriate TDM function. The bound rovibrational wave functions χv''J'' and continuum wave functions χk'J'(R) are computed using the standard Numerov method with a step size of 0.001 Bohr. They are normalized such that they behave asymptotically as

Equation (7)

where ηJ' is the single-channel phase shift of the upper electronic state. Finally, the degeneracy factor g is given by

Equation (8)

where Λ' and Λ'' are the angular momenta projected along the nuclear axis for the final and initial electronic states, respectively.

Predissociation is also possible through an intermediate transition to a bound level of an excited state. In units of cm2, the predissociation cross section is

Equation (9)

(Heays et al. 2017), where λ is the photon wavelength in Å and fuℓ is the oscillator strength of the transition from lower state  to upper state u. We approximate the ground-state fractional population x and the upper level tunneling probability ηd to both be 1 to give an upper limit to the predissociation cross section.

2.3. LTE Cross Sections

In LTE, a Boltzmann population distribution is assumed for the rovibrational levels in the electronic ground state. The total quantum-mechanical photodissociation cross section as a function of both temperature T and wavelength λ is

Equation (10)

where giv''J'' = 2J'' + 1 is the total statistical weight, Ev''J'' is the magnitude of the binding energy of the rovibrational level v'' J'', and kb is the Boltzmann constant. The denominator is the rovibrational partition function.

2.4. Photodissociation Rates

The photodissociation rate for a molecule in an ultraviolet radiation field is given by

Equation (11)

where σ(λ) is the photodissociation cross section and I(λ) is the photon radiation intensity summed over all incident angles. The photon radiation intensity emitted by a blackbody with temperature T is

Equation (12)

where h is the Planck constant and c is the speed of light.

We also compute the photodissociation rate in the unattenuated ISRF, as given by Draine (1978), but modified for λ > 2000 Å by Heays et al. (2017), using Equation (11). In an interstellar cloud the radiation field is attenuated by dust reducing the photodissociation rate as a function of depth into the cloud, or parameterized as the visual extinction AV. Assuming a plane-parallel, semi-infinite slab, with both sides of the cloud exposed isotropically to the ISRF, we applied the radiative transfer code of Roberge et al. (1991) to compute the photodissociation rate as a function of AV and fit the rate to the forms

Equation (13)

Equation (14)

where E2 is the second-order exponential integral. The grain model of Draine & Lee (1984) that was adopted corresponds to the galactic average of the total-to-selective extinction RV = 3.1.

3. Results and Discussion

3.1. State-resolved Cross Sections

State-resolved photodissociation cross sections have been computed for transitions from 14,908 initial rovibrational levels in the X 1Σ+ ground electronic state to the six considered excited electronic states. Cross sections are computed for photons with wavelengths starting at 500 Å up to at most 50,000 Å in 1 Å increments, typically stopping at the relevant threshold. A smaller wavelength step size is used near thresholds to resolve appropriate resonances. In Figure 3, a comparison of the state-resolved cross sections from the ground rovibrational level v'', J'' = 0, 0 for each transition is shown. The 2 1Π and A' 1Σ+ (2 1Σ+) transitions have the dominant cross sections from the ground rovibrational level, while the transition to the A 1Π state makes very little contribution. The behavior of the current cross sections are significantly different from those adopted in Heays et al. (2017).

Figure 3.

Figure 3. Comparison of CS state-resolved cross sections for transitions from the ground rovibrational level v'', J'' = 0, 0. The CS cross section estimate of van Dishoeck (1988), as adopted in Heays et al. (2017), is shown for comparison. Continuum cross sections (solid lines); predissociation (points) longward of ∼1250 Å.

Standard image High-resolution image

Predissociation is possible following bound–bound transitions to the B 1Σ+, 3 1Π, and 4 1Π states. Estimates of predissociation cross sections are computed for transitions to a wide range of bound rovibrational levels. Cross sections for transitions from v'', J'' = 0, 0 are shown in Figure 3 computed using Equation (9). We find that the line cross sections due to predissociation are much smaller than the direct cross sections for the 4 1Π state. However, while predissociation through the B 1Σ+ and 3 1Π states give cross sections comparable to those of their direct continuum cross sections, the continuum cross section for the 2 1Π dominates the predissociation lines by more than an order of magnitude over the relevant wavelength range. Predissociation does not appear to be important for the photodestruction of CS and is therefore not considered further.

The ${{\rm{A}}}^{{\prime} }{}^{1}{{\rm{\Sigma }}}^{+}\,\leftarrow $ X 1Σ+ transition generally has large state-resolved cross sections; so a sampling of cross sections are displayed in Figure 4. Cross sections are plotted for several rotational levels of the ground vibrational level v'' = 0, and for several vibrational levels at their respective lowest rotational level, J'' = 0. State-resolved cross sections for the other five electronic transitions have also been computed (not shown).

Figure 4.

Figure 4. A sample of CS state-resolved cross sections for the A' 1Σ+ $\leftarrow $ X 1Σ+ photodissociation transition. Transitions from initial rovibrational levels (a) where J'' = 0 and (b) where v'' = 0 are shown.

Standard image High-resolution image

3.2. LTE Cross Sections

LTE cross sections have been computed for each transition using the state-resolved cross sections from 1000 K to 10,000 K in 1000 K intervals. A comparison of LTE cross sections for each transition as a function of photon wavelength at 3000 K is displayed in Figure 5. The ${{\rm{A}}}^{{\prime} }{}^{1}{{\rm{\Sigma }}}^{+}\,\leftarrow $ X 1Σ+ transition is the dominant transition at longer wavelengths, while the $4{}^{1}{\rm{\Pi }}\,\leftarrow $ X 1Σ+ transition dominates for short wavelengths. Since the ${{\rm{A}}}^{{\prime} }{}^{1}{{\rm{\Sigma }}}^{+}\,\leftarrow $ X 1Σ+ transition is dominant for the majority of wavelengths, LTE cross sections for this transition at several temperatures are shown in Figure 6.

Figure 5.

Figure 5. CS LTE cross sections at 3000 K for each of the six considered photodissociation transitions.

Standard image High-resolution image
Figure 6.

Figure 6. LTE cross sections for various kinetic temperatures for the ${{\rm{A}}}^{{\prime} }{}^{1}{{\rm{\Sigma }}}^{+}\leftarrow {\rm{X}}{}^{1}{{\rm{\Sigma }}}^{+}$ transition of CS. The v'', J'' = 0, 0 state-resolved cross section is included as well for comparison.

Standard image High-resolution image

3.3. Photodissociation Rates

Photodissociation rates for transitions from v'', J'' = 0, 0 for all six electronic transitions have been computed for the unattenuated ISRF and for the attenuated ISRF into interstellar clouds with total visual extinction. These values are listed and compared with those of Heays et al. (2017) and the UMIST compilation (McElroy et al. 2013) in Table 4. We consider fiducial diffuse and dense clouds with total visual extinctions of AV = 1 and 20, respectively. Consistent with the cross section magnitudes, the ISRF photodissociation rates are dominated by the A${}^{{\prime} }{}^{1}{{\rm{\Sigma }}}^{+}\,\leftarrow $ X 1Σ+ and 2 1Π $\leftarrow $ X 1Σ+ transitions, which leave the two atoms in their ground states. However, about 10% of the photodissociation yield results in both C and S in their 1D metastable states through the 4 1Π $\leftarrow $ X 1Σ+ transition. Using reliable CS photodissociation cross sections, the current unattenuated ISRF rates are about a factor of ∼2.5–3 smaller than the estimates adopted by Heays et al. (2017) and McElroy et al. (2013).

Table 4.  Interstellar CS Photodissociation Rate Fitsa

Source ISRF Dense Cloud     Diffuse Cloud   Products
  k(s−1) a1(s−1) a2 a1(s−1) a2 a3 C + S
    [a4(s−1)] [a5]        
A 1Π $\leftarrow $ X 1.50(−21) 5.43(−22) 2.085 8.29(−22) 3.73 4.00 3P + 3P
2 1Π $\leftarrow $ X 1.35(−10) 5.11(−11) 2.50 7.29(−11) 4.16 4.37  
A' 1Σ+ $\leftarrow $ X 1.94(−10) 7.59(−11) 2.19 1.08(−10) 3.12 3.86  
3 1Π $\leftarrow $ X 5.28(−15) 1.86(−15) 3.16 2.75(−15) 5.55 5.55 1D + 1D
B 1Σ+ $\leftarrow $ X 9.56(−13) 3.53(−13) 2.84 5.05(−13) 4.89 5.00  
4 1Π $\leftarrow $ X 4.05(−11) 1.47(−11) 3.02 2.12(−11) 5.25 5.30  
Total 3.70(−10) 1.48(−10) 2.32 2.163(−10) 3.98 4.21
  [2.13(−10)] [1.69]
Heaysb 9.49(−10) 5.41(−10) 2.49
  [9.49(−10)] [1.95]
UMISTc 9.70(−10) 9.70(−10) 2.00

Notes.

aFits to Equations (13) and (14). bHeays et al. (2017). cMcElroy et al. (2013).

Download table as:  ASCIITypeset image

We computed photodissociation rates for a blackbody radiation field; Figure 7 shows a plot of the photodissociation rates when the molecule is initially in the v'', J'' = 0, 0 rovibrational level for each final electronic state versus the blackbody temperature for a wide range of temperatures. Blackbody photodissociation rates were also obtained by Heays et al. (2017), but they were normalized to reproduce the ISRF energy density from 912 to 2000 Å, as opposed to the normalization inherent in Equation (12) adopted here. Appropriate scale factors, e.g., geometric dilution, should be applied for the relevant astrophysical environment. At the highest temperatures, the current photorates should be taken as a lower limit as photoionization and photodissociation through high-lying Rydberg states will begin to become important.

Figure 7.

Figure 7. CS photorates in blackbody radiation fields for transitions from v'', J'' = 0, 0 to each excited electronic state as a function of radiation temperature.

Standard image High-resolution image

Finally, we consider a situation where a gas containing CS is in LTE at a certain temperature and is immersed in a radiation field generated by a blackbody at the same temperature (i.e., equal gas kinetic and radiation temperatures). The photodissociation rates of CS in such a situation are computed using the LTE cross sections; a plot of these rates against the blackbody/gas temperature is shown in Figure 8.

Figure 8.

Figure 8. LTE blackbody photorates for each CS transition as a function of temperature when the kinetic and radiation temperatures are equal.

Standard image High-resolution image

4. Conclusions

Accurate cross sections for the photodissociation of the CS molecule have been computed for transitions to several excited electronic states using new ab initio potentials and transition dipole moment functions. The state-resolved cross sections have been computed for nearly all rotational transitions from vibrational levels v'' = 0 through v'' = 84 of the X 1Σ+ ground electronic state of CS. Predissociation is found to be significantly smaller than direct photodissociation for CS. Additionally, LTE cross sections have been computed for temperatures ranging from 1000 to 10,000 K. The computed cross sections are applicable to the photodissociation of CS in a variety of UV-irradiated interstellar environments including diffuse and translucent clouds, circumstellar disks, and protoplanetary disks. Photodissociation rates in the interstellar medium and in regions with a blackbody radiation field have been computed as well. To facilitate the calculation of local photorates for particular astrophysical environments, all photodissociation cross section data can be obtained from the UGA Molecular Opacity Project website.6

The work of R.J.P. and P.C.S. was supported by NASA grant NNX15AI61G. B.M.M. acknowledges support by the U.S. National Science Foundation through a grant to ITAMP at the Harvard-Smithsonian Center for Astrophysics under the visitor's program and Queen's University Belfast for a visiting research fellowship (VRF). The molecular structure calculations were performed at the National Energy Research Scientific Computing Center (NERSC) in Berkeley, CA, USA, and at the High Performance Computing Center Stuttgart (HLRS) of the University of Stuttgart, Stuttgart, Germany, where grants of time are gratefully acknowledged. R.C.F. acknowledges support from NSF grant No. PHY-1503615. ITAMP is supported in part by NSF grant No. PHY-1607396. We thank the referee for useful comments that improved the manuscript.

Footnotes

Please wait… references are loading.
10.3847/1538-4357/aab5b9