The potential energy surface (PES) describing the interaction of the ethynyl (C2H) radical in its ground X̃2Σ+ electronic state with molecular hydrogen has been computed through restricted coupled cluster calculations including single, double, and (perturbative) triple excitations [RCCSD(T)], with the assumption of fixed molecular geometries. The computed points were fit to an analytical form suitable for time-independent quantum scattering calculations of rotationally inelastic cross sections and rate constants. A representative set of energy dependent state-to-state cross sections is presented and discussed. The PES and cross sections for collisions of H2(j = 0) are compared with a previous study [F. Najar et al., Chem. Phys. Lett. 614, 251 (2014)] of collisions of C2H with H2 treated as a spherical collision partner. Good agreement is found between the two sets of calculations when the H2 molecule in the present calculation is spherically averaged.

The ethynyl radical, C2H(X̃2Σ+), is one of the smallest unsaturated hydrocarbon radicals and is an important interstellar molecule. It also plays a role in the combustion of acetylene and appears as an intermediate in the combustion of fuel-rich mixtures of larger hydrocarbons. Laufer and Fahr1 have reviewed what is known about the reaction kinetics of the ethynyl radical.

This radical was detected in the interstellar medium through the observation of radiative transitions2 before its laboratory rotational spectrum was observed.3–5 Subsequently, C2H has been detected in a great variety of astrophysical environments, including cold dark clouds,6 photon dominated regions,7 protoplanetary disks,8 pre-stellar cores,9 and high-mass star-forming regions.10 Ethynyl is found to be a very abundant molecule in the interstellar medium and is by far the most abundant unsaturated hydrocarbon. The abundance of ethynyl is sufficiently high that the 13CCH and C13CH isotopologues have been observed in the interstellar medium.11 Even though the interstellar H to D ratio is only ∼10−5, the ratio of interstellar C2D to C2H is enhanced by a large factor and can be diagnostic for the chemical evolution of the cloud.

Laboratory studies of the rotational spectrum C2H radical3–5 have confirmed that the ground electronic state is a 2Σ+ state. Because of the nonzero nuclear spin of H (I=12), as well as of D (I = 1), the rotational lines of both C2H and C2D have a hyperfine structure, which is usually resolved in the astronomical spectra.

There has been a considerable spectroscopic study of its low-lying Ã2Π electronic state12–14 and excited vibronic levels of the ground state.15–22 The Ã2Π state is only ∼3600 cm−1 above the ground X̃2Σ+ electronic state. Because of the mixing of the ground state with the A′ component of the Π state as the molecule bends from its linear equilibrium structure,23 there is considerable vibronic mixing, even around the origin of the ÃX̃ transition. The vibronic levels of the ground state near the origin of this transition have considerable mixed character24 and have led to difficulties in the assignment of the bands.19 It should be noted that two Ã2ΠX̃2Σ+ bands near 4000 cm−1 have been detected in the outer region of the circumstellar envelope of CW Leo.25 

Determination of the abundance of C2H in the interstellar medium is of particular interest for understanding the chemistry of carbon containing species. Under typical conditions, the rotational levels of C2H are not in thermodynamic equilibrium because of the low density of the interstellar medium.9 Hence, the modeling of the molecular emission intensities requires application of a radiative transfer model (see, for example, Ref. 26). Such modeling requires the availability of radiative transition rates, as well as rate coefficients for rotational transitions induced by collisions with the dominant interstellar species, in particular, the H2 molecule.

In the absence of rate constants for rotationally inelastic collisions with H2, rate constants involving collisions with helium are often used as surrogates in radiative transfer calculations.27 Spiefiedel et al.28 computed a potential energy surface describing the interaction of C2H with helium and employed this surface to compute rate constants as a function of temperature for rotational transitions in C2H induced by collisions with helium. Since the hyperfine structure is spectrally resolved in most C2H astronomical observations, these workers employed the recoupling method29 to calculate cross sections for hyperfine-resolved C2H collisional transitions.

These C2H–He rate constants provided estimates of the rate constants for C2H–H2 inelastic collisions, with a scaling factor to account for the difference in the masses of He and H2. These scaled C2H–H2 rate constants have been employed in the application of radiative transfer models describing C2H in various astrophysical environments.30–34 However, based on comparisons of rate constants for collisions of He and H2 with other molecular species, this scaling leads only to qualitatively correct rate constants for H2 collisions.27,35,36

Najar et al.37 carried out a study of the collisional excitation of C2H with H2 in which they treated the hydrogen molecule as a spherical collision partner. The potential energy surface for the interaction of C2H with spherical H2, i. e., H2(j = 0), was computed by taking the average energy for three orientations of H2 with respect to C2H. Cross sections and rate constants could then be computed in a pseudo-atom molecule scattering calculation; this was much simpler than treating the full molecule–molecule interaction. This potential energy surface was recently employed with the recoupling method29 to compute hyperfine-resolved rate constants for collisions of C2H and C2D (with a shifting of the center of mass) rotational/fine-structure levels with para-H2(j = 0).38 

In the present study, the interaction of C2H with H2 is treated explicitly as a molecule–molecule interaction, and cross sections and rate constants for collisions with both para-H2(j = 0) and ortho-H2(j = 1) are computed. As in the previous studies of inelastic collisions of C2H,28,37 the molecule is treated as rigid. It should be noted that C2H can react with hydrogen to yield acetylene through the reaction C2H + H2 → C2H2 + H. This reaction has been investigated both experimentally and theoretically.39–43 The reaction proceeds through an early transition state barrier that has a linear HCC–H2 geometry. A range of barrier heights (2.3–3.1 kcal mol−1, including zero-point energy corrections) has been computed.39–42 Despite the low barrier, the reaction is inhibited by the early barrier, and the reaction probability was computed in a quantum mechanical calculation42 with an 8-dimensional PES41 to be <0.01 for collision energies up ∼1000 cm−1. The rate constant for this reaction is small at room temperature and below (k298 = 5.8 × 10−13 cm3 molecule−1 s−1).1 The reactive part of the C2H–H2 potential energy surface will not be accessed in the present calculation with the assumption of rigid structures for the collision partners.

Section II describes the calculation of the C2H–H2 potential energy surface; Sec. II A presents the ab initio calculation of the surface, while Sec. II B describes our fit of the calculated ab initio points to a form suitable for scattering calculations, and Sec. II C displays various features of the fitted potential energy surface. The scattering calculations are briefly described in Sec. III. The computed cross sections and rate constants are presented in Sec. IV. The paper concludes with a discussion in Sec. V.

Despite the strong vibronic mixing between the X̃2Σ+ and Ã2Π electronic states in C2H when the molecule bends, Najar et al.37 concluded that they could use coupled cluster theory to compute the interaction between linear C2H(X̃2Σ+) and H2. They carried out complete active space self-consistent field (CASSCF) calculations of the ground 2A′ state and found that the weight of the leading configuration was greater than 0.94 for the range of geometries investigated. We have thus employed coupled cluster theory to compute the full C2H(X̃2Σ+)–H2 potential energy surface. To take account of the zero point motion in the molecules,44 we have taken the H2 internuclear separation as the average, rHH = 1.449a0, over the probability distribution for the v = 0 vibrational level. For C2H, we take rCC = 2.299a0 and rCH = 1.968a0 from the substitution structure derived experimental rotational constants.45 

It is possible that the assumption of strict linearity of the C2H molecule in rotationally inelastic collisions is questionable. Denis-Alpizar et al.46 have considered the possible effect of neglecting the bending motion in treating rotationally inelastic collisions of HCN with helium. They find that this assumption barely affects the rotational excitation of the molecule, even for collision energies above the threshold for bending excitation. We verified through calculations of the C2H energy vs. bond angle that the molecule is linear in its equilibrium geometry. Since the equilibrium geometry, rotational constants, and vibrational modes of HCN and C2H are similar, we assume that the bending motion of C2H can be neglected in treating rotationally inelastic collisions. This assumption was also made by Najar et al.37 in their treatment of collisions of C2H with spherical H2.

Figure 1 illustrates the coordinates describing the geometry of the C2H–H2 complex. The interaction depends upon four coordinates: the angle θ1 between the C2H molecular axis and the Jacobi vector R, the angle θ2 between the H2 molecular axis and R, the dihedral angle ϕ, and the separation R between the centers of mass of C2H and H2.

FIG. 1.

Definition of the coordinate system describing the geometry of the C2H–H2 complex. The C2H moiety lies in the xz plane, with θ1 = 0° corresponding to the carbon end pointing toward the H2 moiety.

FIG. 1.

Definition of the coordinate system describing the geometry of the C2H–H2 complex. The C2H moiety lies in the xz plane, with θ1 = 0° corresponding to the carbon end pointing toward the H2 moiety.

Close modal

Restricted coupled cluster calculations with inclusion of single, double, and (perturbatively) triple excitations [RCCSD(T)47,48 were employed to compute points on the C2H–H2 potential energy surface. The MOLPRO 2012 suite of computer codes49 was employed for these calculations. A counterpoise correction50 was applied to the computed interaction energies for each configuration Ω = (θ1, θ2, ϕ),

(1)

where the energies of the C2H and H2 moieties are computed with the basis set for the full complex.

In preliminary calculations, the convergence of the basis set employed was examined. Radial cuts were computed for several orientations around the global minimum of the potential energy surface using the aug-cc-pVTZ, aug-cc-pVQZ, and aug-cc-pV5Z basis sets.51 It can be seen in Fig. 2 that for all geometries, the interaction is computed to be more attractive when computed with a successively larger basis set. The differences in the interaction energy computed with the aug-cc-pVQZ and aug-cc-pV5Z basis sets are less than 1 cm−1 for R > ∼7.5a0. As found by Najar et al.,37 the most attractive orientation has θ1 = 180°, θ2 = 90°, ϕ = 0°, corresponding to a T-shaped structure with the hydrogen end of C2H pointing toward the center of the H2 moiety. For this radial cut, the minimum of the interaction energy equals −132.3 and −133.4 cm−1 at R = 7.83 and 7.82a0 when computed with the aug-cc-pVQZ and aug-cc-pV5Z basis sets, respectively. This well depth is greater than the computed well depth (−61.0 cm−1) for the interaction of C2H with spherical H2,37 as is expected.

FIG. 2.

C2H–H2 interaction energy as a function of the intermolecular separation for several orientations, defined by (θ1, θ2, ϕ), computed with the aug-cc-pVTZ (blue curves), aug-cc-pVQZ (green curves), and aug-cc-pV5Z (red curves) basis sets.

FIG. 2.

C2H–H2 interaction energy as a function of the intermolecular separation for several orientations, defined by (θ1, θ2, ϕ), computed with the aug-cc-pVTZ (blue curves), aug-cc-pVQZ (green curves), and aug-cc-pV5Z (red curves) basis sets.

Close modal

The full C2H–H2 potential energy surface was computed with the aug-cc-pVQZ basis set. Some initial calculations were carried out with the addition of mid-bond functions.52,53 However, for many geometries, the calculations did not converge, as the dummy atom would lie too close to a real atom. In the end, the calculations were carried out without the addition of mid-bond functions. The calculations were carried out on a 4-dimensional grid (R,θ1,θ2,ϕ), at 28 values of the intermolecular separation R ranging from 4.25 to 20a0. The angular range covered was 0° < θ1 < 180°, 0° < θ2 < 90°, and 0° < ϕ < 180°. The initial angular grid included 858 orientations, with 11 values of θ1 defined by cos θ1 = −1 to 1 in steps of 0.2, 6 values of θ2 defined by cos θ2 = −1 to 0 in steps of 0.2, and 13 values of ϕ from 0° to 180° in steps of 15°. This choice of the angular grid provides a uniform sampling of the differential solid angle sin 1 sin 2.54 

Because of the significant anisotropy of the C2H–H2 potential energy surface and difficulties in fitting the surface, in particular for small values of R, discussed in Sec. II B, it was found necessary to compute the interaction energy at more orientations than specified above. Further calculations were carried out for 6 additional values of θ1 over the same range of θ2 and ϕ values, as well as other orientations. Interaction energies were computed for a total of 1591 and 1643 orientations for R ≥ 5.57a0 and R ≤ 5a0, respectively. The total number of geometries for which the interaction energy was computed equals 43 165.

It is convenient to expand the molecule-molecule interaction energy in a basis of bispherical harmonics for each value of R to facilitate the solution of the close-coupling scattering equations. The expansion derived by Green55 for the interaction of two linear molecules both in Σ electronic states was employed here. The potential is expressed as

(2)

In Eq. (2), the basis functions are defined as

(3)

where (….|..) is a Clebsch-Gordan coefficient and the Ylm are spherical harmonics. Equation (2) can also be written as55 

(4)

where Plm(θ) is an unnormalized Legendre function, defined in terms of the spherical harmonic Ylm(θ, ϕ) as Ylm(θ,ϕ)=Plm(θ)eimϕ.

For each value of the intermolecular separation R, the computed values of the interaction energy computed on the grid of orientations were fit in a linear least squares procedure to Eqs. (2) and (3). Fits were carried out with different numbers of expansion coefficients, determined by the maximum values of l1 and l2 in the sum in Eq. (2). The allowed values of the l are given by the vector addition of l1 and l2, with the restriction that l1 + l2 + l must be even.56 

For certain orientations at small R, in particular with θ1 ∼ 180°, the potential is very repulsive. At these geometries, the H end of the C2H moiety, which is considerably displaced from the C2H center of mass, comes close to the H2 molecule. To facilitate the fits, we followed the procedure of Wormer et al.56 and damped the interaction energy with a hyperbolic tangent function up to a maximum value of Vmax. In the present study, Vmax was chosen to equal 20 000 cm−1.

Least squares fits for values of R in the radial grid of the ab initio calculations were carried out with different values of maximum values of l1 and l2. A acceptable fit was obtained for l1max=12 and l2max=6. The choice of these parameters leads to a total of 174 radial expansion coefficients vl1l2l(R) in Eq. (2). Fits with larger values of l1max (=13 and 14) yielded RMS deviations at small R essentially the same as for l1max = 12.

As an indication of the quality of the fit, we compare in Fig. 3 the RMS deviation of the fit to the absolute value of the average of the potential as a function of R. The kink in the absolute value of the average of the potential near R = 5.8a0 is due to a change in the sign of the average value of the potential. We see in Fig. 3 that the RMS deviation of the fit is very small at large R and increases as R is decreased. Except for the repulsive region for R ≤ 5a0, the RMS deviation of the fit is well under 1% of the average of the potential.

FIG. 3.

Logarithmic plot of the RMS deviation of the fit and absolute value of the average of the C2H–H2 potential as a function of the intermolecular separation R.

FIG. 3.

Logarithmic plot of the RMS deviation of the fit and absolute value of the average of the C2H–H2 potential as a function of the intermolecular separation R.

Close modal

For large R, a multipole expansion was employed to describe the long-range electrostatic interaction in terms of the multipole moments of C2H and H2, as was carried out for HCl–H2 and OH–H2.56,57 We considered the dipole, quadrupole, and octopole moments of C2H and the quadrupole moment of H2. We computed the multiple moments of C2H through MRCI calculations with an aug-cc-pV5Z basis set and obtained the values q10 = −0.2897 a.u., q20 = 3.3693 a.u., and q30 = −5.5812 a.u. For the quadrupole moment of H2, we employed the value (q20 = 0.5235 a.u.) computed by Ma et al.57 

These multipole moments were used to compute the following expansion terms at long-range: CH-dipole/H2 quadrupole (V123), quadrupole-quadrupole (V224), and octopole-quadrupole (V325). For these expansion coefficients, we switched from the values obtained from the least squares fits to the ab initio points at smaller R with a hyperbolic tangent function. The isotropic term (v000) was switched to a C6R−6 radial dependence at long range, with the C6 coefficient determined by the value of V000 at R = 12a0. The remaining expansion coefficients were damped to zero at large R with a hyperbolic tangent function.

The global minimum of the potential energy surface corresponds to a T-shape structure, with the hydrogen end of C2H pointing toward the center of H2. The radial cut for this orientation (θ1 = 180°, θ2 = 90°, ϕ = 0°) is displayed in Fig. 2(d). The dissociation energy De of the C2H complex is computed to equal 133.4 cm−1 at an equilibrium intermolecular separation Re = 7.82a0 (computed with the aug-cc-pV5Z basis).

Figure 4 presents contour plots as a function of the polar angles θ1 and θ2 of the C2H–H2 potential energy surface for various planar geometries (ϕ = 0°), with the intermolecular separation R ranging from the repulsive to attractive regions. We see that the most repulsive orientations for all the values of R displayed have θ1 near 180°, while the potential at small R for θ1 near 0° is more repulsive then for the intermediate values of θ1. These geometries correspond to having the hydrogen end of C2H pointing toward the H2 moiety. This significant repulsion arises because the H atom in C2H is a considerable distance from the center of mass of C2H. Indeed, the potential is repulsive even at R = 7.5a0 for geometries near θ1 = 180° and θ2 = 0° (see Fig. 4). By contrast, the potential for CH–H2 at R = 7.5a0 is attractive for all orientations; in the case, the hydrogen atom in displaced from the CH center of mass by a much smaller distance.58 

FIG. 4.

Contour plots of the C2H–H2 potential energy surface as a function of the polar angles θ1 and θ2 with the dihedral angle ϕ = 0° (corresponding to planar geometries) for several values of the intermolecular separation R.

FIG. 4.

Contour plots of the C2H–H2 potential energy surface as a function of the polar angles θ1 and θ2 with the dihedral angle ϕ = 0° (corresponding to planar geometries) for several values of the intermolecular separation R.

Close modal

To explore further the features of the C2H–H2 potential energy surface, we present in Fig. 5 contour plots for two nonplanar geometries (ϕ = 90°) at intermolecular separations R = 5.5 and 7.5a0. These contour plots can be compared with the plots in Figs. 4(c) and 4(d), which present plots planar geometry (ϕ = 0°) at the same values of R. For R = 5.5a0, the potential for some orientations is slightly less repulsive for ϕ = 90° than for ϕ = 0°. For R = 7.5a0, the potential is still repulsive for orientations around θ1 = 180°, θ2 = 0°, as is also the case for planar geometries. Comparing Figs. 4(d) and 5(b), the attractive part of the potential is slightly less strong for ϕ = 90°.

FIG. 5.

Contour plots of the C2H–H2 potential energy surface as a function of the polar angles θ1 and θ2 with the dihedral angle ϕ = 90° for several values of the intermolecular separation R.

FIG. 5.

Contour plots of the C2H–H2 potential energy surface as a function of the polar angles θ1 and θ2 with the dihedral angle ϕ = 90° for several values of the intermolecular separation R.

Close modal

The contour plots presented in Figs. 4 and 5 show that the potential for the interaction of C2H with H2 is strongly anisotropic. This is a result by the large displacement of the H atom in C2H from the center of mass of the molecule. We may anticipate that collisions of C2H with H2 will exhibit significant inelasticity. This has already been partly demonstrated in the scattering calculations carried out by Najar et al.,37 on collisions of C2H with spherical H2. We may anticipate larger cross sections for inelastic cross sections for collisions of ortho-H2(j = 1), for which the full anisotropy of the potential energy surface can be accessed.

Time-independent quantum scattering calculations were carried with the HIBRIDON suite of programs.59 The theory of scattering of two linear molecules both in 1Σ+ electronic states has been worked out by Green.55 We have extended this treatment to describe collisions of a molecule in a 2Σ+ electronic state with a molecule in a 1Σ+ state. We present in the  Appendix the expression for the matrix element of the potential between space-frame scattering channel basis functions for this system. Hereafter, we denote the angular momenta of C2H and H2 with subscripts 1 and 2.

Full close-coupling calculations were carried out to compute state-to-state integral cross sections describing inelastic collisions of C2H rotational/fine-structure levels with para-H2(j2 = 0) and ortho-H2(j2 = 1). Spectroscopic constants for C2H were taken from the work of Müller et al.5 In our calculations, the Hamiltonian for the isolated C2H(X̃2Σ+) molecule was described by the rotational constant B, centrifugal distortion constant D, and fine-structure constant γ. The rotational constant B for H2 was taken from Huber and Herzberg.60 The total angular momentum exclusive of electron spin is denoted for the C2H(X̃2Σ+) molecule as n1. The coupling of n1 with the electron spin s1=12 yields two closely spaced fine-structure levels labeled F1 and F2, with total angular momentum j1=n1+12 and j1=n112, respectively. In this work, we do not consider the hyperfine splitting in C2H.

Scattering calculations were carried out for 410 energies up to 1000 cm−1 for collisions with para-H2(j2 = 0) and 729 total energies up to 1200 cm−1 for collisions with ortho-H2(j2 = 1). The calculations were checked for convergence of the cross sections with respect to the C2H rotational basis, spacing in the radial grid, and total number of partial waves included. Depending upon the collision energy, C2H rotational levels up to n1 = 25 were included in the channel basis. The H2 rotational basis included j2 = 0 and 2 for para-H2 and j2 = 1 only for ortho-H2. Partial waves up total angular momentum J=93.5 were included in the calculations.

The computed state-to-state integral cross sections were employed to compute the corresponding rate constants between 15 and 300 K for collisions of C2H rotational/fine-structure levels with H2j2 = 0 and 1 rotational levels by averaging over the collision energy Ec,61 

(5)

where μ is the collision reduced mass and kB is the Boltzmann constant. The integral was evaluated out to collision energies of 1000 cm−1 for collisions of H2(j2 = 0) and 1100 cm−1 for collisions of H2(j2 = 1). The high-energy tail of the integrand in Eq. (5) was evaluated by exponentially extrapolating the cross sections out to 1500 cm−1; this upper limit is sufficient for temperatures up to 300 K. State-to-state cross sections and rate constants were computed for collisions of the 20 lowest C2H rotational/fine-structure levels (j1 ≤ 9.5) with both para-H2(j2 = 0) and ortho-H2(j2 = 1).

It should be noted that cross sections for rotationally inelastic transitions could be slightly overestimated at high collision energies because of the chemical reaction, as was found for H–D2 and H–HCl collisions.62,63 However, the effect of chemical reaction on C2H–H2 inelastic rate constants should not be very significant for temperatures at or below 300 K because of the small reaction probability at low collision energies.42 

Figure 6 presents energy dependent cross sections for the excitation of the C2H n1 = 0 F1 rotational/fine-structure level to n1 = 1–4 levels in collisions with para-H2(j2 = 0) and ortho-H2(j2 = 1). Figures 6(a) and 6(b) display cross sections for fine-structure conserving (Δj1 = Δn1) transitions, and Figs. 6(c) and 6(d) display cross sections for fine-structure changing (Δj1Δn1) transitions. We see that the latter are somewhat smaller than the former. As we show below, this trend becomes more pronounced for high initial rotational levels. This is consistent with the propensity rule derived by Alexander64 for collisions of a molecule in a 2Σ+ electronic state with a structureless target and reflects the fact that the direction of the electron spin cannot be changed in a molecular collision.

FIG. 6.

Integral cross sections for the excitation of the C2H(n1 = 0 F1) level in collisions with para-H2(j2 = 0) and ortho-H2(j2 = 1) as a function of collision energy up to 300 cm−1. Panels (a) and (b) display cross sections for fine-structure conserving transitions, while panels (c) and (d) display cross sections for fine-structure changing transitions, for collisions with H2(j2 = 0) and H2(j2 = 1), respectively. The solid curves in (a) and (c) were computed with a maximum value of j2 = 2 in the scattering basis, while the dashed curves in these panels were computed with a maximum value of j2 = 0. Also included in panels (b) and (d) are cross sections (open circles) for excitation of the n1 = 0 F1 level in collisions with H2(j2 = 2) at selected collision energies.

FIG. 6.

Integral cross sections for the excitation of the C2H(n1 = 0 F1) level in collisions with para-H2(j2 = 0) and ortho-H2(j2 = 1) as a function of collision energy up to 300 cm−1. Panels (a) and (b) display cross sections for fine-structure conserving transitions, while panels (c) and (d) display cross sections for fine-structure changing transitions, for collisions with H2(j2 = 0) and H2(j2 = 1), respectively. The solid curves in (a) and (c) were computed with a maximum value of j2 = 2 in the scattering basis, while the dashed curves in these panels were computed with a maximum value of j2 = 0. Also included in panels (b) and (d) are cross sections (open circles) for excitation of the n1 = 0 F1 level in collisions with H2(j2 = 2) at selected collision energies.

Close modal

For collisions of para-H2, we compare in Figs. 6(a) and 6(c) cross sections calculated with a minimal H2 rotational basis (maximum value of j2 = 0) with a basis with a maximum value of j2 = 2. There are slight differences in the cross sections computed with the two rotational bases, which become larger at higher collision energies. The cross sections computed with a maximum value j2 = 0 are equivalent to the scattering calculations performed by Najar et al.37 for collisions with spherical H2, albeit with a different computed potential energy surface. Comparison of the cross sections displayed in Figs. 6(a) and 6(c) with those presented by Najar et al. in their Fig. 4 shows that there is quite good agreement in the two sets of computed cross sections. In fact, the contour plot for the interaction of C2H with spherical H2, presented in their Fig. 2, and a contour plot computed by spherically averaging the present potential energy surface [which is equivalent to taking only the l2 = 0 terms in the angular expansion in Eq. (2)] are in very good agreement.

The energy dependent cross sections below 60 cm−1 collision energy, presented in Fig. 6 for collisions of para-H2(j2 = 0), show a dense resonance structure consistent with the formation and decay of quasibound states of the collision complex, as was also seen by Najar et al.37 We also see in Figs. 6(b) and 6(d) a structured, but less resolved, energy dependence for cross sections involving collisions of ortho-H2(j2 = 1). These cross sections have less structure because there are many more, and hence overlapping, resonances in the interaction of the molecule with H2(j2 = 1) than with H2(j2 = 0), as was observed for OH–H2 and NH3–H2 collisions.65,66

We also observe in Fig. 6 that the cross sections for collisions with ortho-H2 are larger by a factor of approximately 2 than the cross sections for the same C2H transitions in collisions with para-H2. A similar enhancement of the cross sections for collisions of ortho-H2 was also found in OH–H2 and NH3–H2 collisions.65,66 In the case of the HCN–H2 collision pair,67 the enhancement was roughly a factor of 3, suggesting that the anisotropy with respect to the H2 orientation is greater in this system.

In order to explore the dependence of the cross sections upon the initial H2 level, we have also computed cross sections for collisions of the C2H(n1 = 0 F1) level with H2(j2 = 2); these cross sections at selected collision energies are plotted in Figs. 6(b) and 6(d). We see that the cross sections for collisions of H2(j2 = 1 and 2) are nearly identical. This behavior is quite similar to what was found for OH–H2 and HCN–H2 collisions.66,67 In the interaction of C2H with H2, the full anisotropy of the potential energy surface is experienced in interactions of C2H with H2(j2 > 0) rotational levels. In addition, the long-range interaction is stronger for this nuclear spin modification since the j2 > 1 levels have a nonzero quadrupole moment while the j2 = 0 level does not.

To explore the dependence of the state-to-state cross sections upon the initial C2H level, we present in Fig. 7 cross sections for transitions out of the C2H n1 = 2 F1 level. A dense resonance structure is observed at low collision energies, as is the case for the initial n2 = 0 F1 level. We see a slight difference in the computed cross sections for collisions with H2(j2=0) when the H2 rotational basis is reduced to a maximum value j2 = 0.

FIG. 7.

Integral cross sections for the excitation of the C2H(n1 = 2 F1) level in collisions with para-H2(j2 = 0) and ortho-H2(j2 = 1) as a function of collision energy up to 300 cm−1. Panels (a) and (b) display cross sections for fine-structure conserving transitions, while panels (c) and (d) display cross sections for fine-structure changing transitions, for collisions with H2(j2 = 0) and H2(j2 = 1), respectively. The solid curves in (a) and (c) were computed with a maximum value of j2 = 2 in the scattering basis, while the dashed curves in these panels were computed with a maximum value of j2 = 0.

FIG. 7.

Integral cross sections for the excitation of the C2H(n1 = 2 F1) level in collisions with para-H2(j2 = 0) and ortho-H2(j2 = 1) as a function of collision energy up to 300 cm−1. Panels (a) and (b) display cross sections for fine-structure conserving transitions, while panels (c) and (d) display cross sections for fine-structure changing transitions, for collisions with H2(j2 = 0) and H2(j2 = 1), respectively. The solid curves in (a) and (c) were computed with a maximum value of j2 = 2 in the scattering basis, while the dashed curves in these panels were computed with a maximum value of j2 = 0.

Close modal

We also see that the cross sections for fine-structure changing transitions are much smaller than for fine-structure conserving transitions. Hence, the propensity rule for the dominance of fine-structure conserving transitions has become stronger than for the initial n2 = 0 F1 level. This is expected from the formal analysis by Alexander64 of collisions of 2Σ+ molecule with a structureless target.

We again see that the cross sections for transitions involving collisions with ortho-H2(j2 = 1) are significantly larger than the cross sections for the corresponding transitions involving collisions with para-H2(j2 = 0). The magnitude for the fine-structure conserving transitions for a given change in rotational angular momentum n1 is smaller for the initial level n1 = 2 than for n1 = 0; this reflects the larger energy gap between the initial and final levels for the former.

Finally, we see that for collisions of both para-H2(j2 = 0) and ortho-H2(j2 = 1), the fine-structure conserving transitions with the largest cross sections involve Δn1 = +2. This contrasts with the HCN–H2 collision pair, for which different propensity rules are observed for para-H2(j2 = 0) and ortho-H2(j2 = 1).67 The potential energy surface for the interaction of C2H with spherical H2 (see Fig. 3 in Ref. 37) displays an approximate forward-backward (θ1 < 90° vs. θ1 > 90°) symmetry, indicative of larger angular expansion coefficients for even values of l1, and implies larger cross sections for even Δn1 collisional transitions. This propensity for even Δn1 transitions carries over to the full C2H–H2 potential energy surface, as we see in Figs. 6 and 7 that the largest cross sections for C2H transitions involve Δn1 = +2 for collisions with H2(j = 1).

Rate constants for transitions between the 20 lowest rotational/fine-structure levels of C2H induced by collisions with H2(j = 0) and H2(j = 1) were computed for temperatures between 10 and 300 K, using the energy dependent integral cross sections, such as those displayed in Figs. 6 and 7. A table of de-excitation rate constants is included in the supplementary material. Computed rate constants for excitation transitions out of the C2H n1 = 0 F1 initial level induced by collisions with H2(j2 = 0) and H2(j2 = 1) are presented in Fig. 8. Also plotted in panels (a) and (b) of Fig. 8 are rate constants computed by Najar et al.37 for C2H–para-H2 collisions. There is quite reasonable agreement between these rate constants and the ones computed in the present work. The slight differences are due to small differences in the PESs employed and the difference in the H2 rotational bases (j2max=2 and 0 in the present work and in the work of Najar et al., respectively).

FIG. 8.

Rate constants for the excitation of the C2H(n1 = 0 F1) level in collisions with para-H2(j2 = 0) and ortho-H2(j2 = 1) as a function of temperature between 10 and 300 K. Panels (a) and (b) display cross sections for fine-structure conserving transitions, while panels (c) and (d) display cross sections for fine-structure changing transitions, for collisions with H2(j2 = 0) and H2(j2 = 1), respectively. The dashed lines in panels (a) and (c) present rate constants (plotted up to 80 K) computed in the work of Najar et al (Ref. 37, respectively).

FIG. 8.

Rate constants for the excitation of the C2H(n1 = 0 F1) level in collisions with para-H2(j2 = 0) and ortho-H2(j2 = 1) as a function of temperature between 10 and 300 K. Panels (a) and (b) display cross sections for fine-structure conserving transitions, while panels (c) and (d) display cross sections for fine-structure changing transitions, for collisions with H2(j2 = 0) and H2(j2 = 1), respectively. The dashed lines in panels (a) and (c) present rate constants (plotted up to 80 K) computed in the work of Najar et al (Ref. 37, respectively).

Close modal

As with the energy dependent cross sections, the largest rate constants involve Δn1 = +2 transitions in C2H. We also see that the rate constants for collisions with H2(j2 = 1) are significantly larger than the corresponding rate constants involving collisions with H2(j2 = 0). The differing dependence of the cross sections upon the collision energy for collisions of H2(j2 = 0) and H2(j2 = 1) (see Fig. 6) is reflected in differences in the temperature dependence of the corresponding rate constants. As found by Najar et al.,37 the rate constants for the H2(j2 = 0) collision partner rise gradually with temperature from their small values at low temperature. By contrast, the rate constants at low temperature change rapidly with temperature and then become relatively constant at higher temperature. The same propensity rules favoring transitions for which Δj1 = Δn1 are also found, as with the energy dependent cross sections.

In this work, we have reported the calculation of the potential energy surface describing the interaction between C2H(X̃2Σ+) and H2. These calculations were carried out using coupled cluster theory [RCCSD(T)]. We find that the interaction is strongly anisotropic, principally because of the repulsion between the H atom in the C2H and H2 molecule. The former is located at a considerable distance from the C2H center of mass. Since the geometries of the collision partners are fixed, the computed potential energy surface does not describe the reactive portion of the surface.

This potential energy surface was employed in time independent quantum scattering calculations to compute cross sections for transitions between C2H rotational/fine-structure levels induced by collisions with both para-H2(j2 = 0) and ortho-H2(j2 = 1) as a function of collision energy. The cross sections computed for collisions with para-H2 are in reasonable agreement with the work of Najar et al.,37 who treated H2(j2 = 0) as a spherical collision partner, by averaging the C2H–H2 interaction over three orientations of H2.

In the present study, we have taken into account the full anisotropy of the C2H potential energy surface and found that the cross sections and rate constants for collisions with ortho-H2 are significantly larger than for para-H2. This implies that the rate constants for collision of C2H with the hydrogen molecule will depend strongly on the ratio of para-H2 and ortho-H2. The C2H–H2 collision pair is yet another example for which the rotational level of H2 strongly affects the anisotropy of the interaction. This has recently been dramatically demonstrated in the presence or absence of resonances in low-energy collisions.68 

Our primary motivation for this work is to provide C2H–H2 rate constants for use in radiative transfer calculations to interpret astronomical observations of C2H spectroscopic transitions. The rotational spectrum of C2H has a resolvable hyperfine structure,3–5 which can usually be resolved in the astronomical observations.30–34 Hence, it would be desirable to compute rate constants for transitions between hyperfine-resolved rotational/fine-structure levels of C2H in collisions with both para-H2(j2 = 0) and ortho-H2(j2 = 1). Dumouchel et al.38 have computed rate constants for hyperfine-resolved transitions in C2H with para-H2(j = 0), using the PES of Najar e al.37 They also computed hyperfine-resolved transitions in C2D–para-H2(j = 0) collisions. In this case, the PES was obtained from the C2H–H2 PES by shifting the C2H/D center of mass. In future work, we plan to employ the recoupling method29 to compute rate constants for transitions between hyperfine-resolved rotational/fine-structure transitions in C2H induced by collisions with both H2(j2 = 0) and H2(j2 = 1), for use in radiative transfer simulations for the interpretation of astronomical observations of C2H.

See supplementary material for the file PES.tar, which contains a Fortran program for computing the C2H–H2 interaction energy for entered values of R, θ1, θ2, ϕ, and a table of rate constants for de-excitation transitions between the 20 lowest rotational/fine-structure levels of C2H induced by collisions with H2(j = 0) and H2(j = 1) for 7 temperatures between 10 and 300 K.

The author is grateful for support from the U.S. National Science Foundation (Grant No. CHE-1565872). François Lique kindly provided the rate constants computed in the work of Najar et al. (Ref. 37). The quantum chemistry calculations were performed at the Maryland Advanced Research Computing Cluster, which was funded by the state of Maryland and is jointly managed by Johns Hopkins University and the University of Maryland College Park.

This appendix presents the expression for the matrix element of the potential between space-frame scattering channel basis functions for the collision of a linear molecule in a 2Σ+ electronic state with a linear molecule in a 1Σ+ electronic state. The angular momentum of the two molecules is denoted j1 and j2, respectively. The angular momentum j12 is the vector sum of j1 and j2. The total angular momentum and orbital angular momentum of the complex are denoted J and L, respectively. Hence, the channel basis functions are defined by the quantum numbers j1Fij2j12LJM, where Fi, i = 1 or 2, labels the fine-structure level of the 2Σ+ molecule.

The matrix element of the potential, which is described by the angular expansion in Eqs. (2) and (3), equals

(A1)

where [x] = 2x + 1, (:::), {:::}, and are 3-j, 6-j, and 9-j symbols, respectively.69 The unprimed and primed quantum numbers in Eq. (A1) apply to the initial state i and final state f, respectively. The symmetry index64 ϵ in Eq. (A1) equals +1 and −1 for the F1 and F2 fine-structure levels, respectively, of the 2Σ+ molecule.

1.
A. H.
Laufer
and
A.
Fahr
,
Chem. Rev.
104
,
2813
(
2004
).
2.
K. D.
Tucker
,
M. L.
Kutner
, and
P.
Thaddeus
,
Astrophys. J.
193
,
L115
(
1974
).
3.
K. V. L. N.
Sastry
,
P.
Helminger
,
A.
Charo
,
E.
Herbst
, and
F. C.
DeLucia
,
Astrophys. J.
251
,
L119
(
1981
).
4.
C. A.
Gottlieb
,
E. W.
Gottlieb
, and
P.
Thaddeus
,
Astrophys. J.
264
,
740
(
1983
).
5.
H. S. P.
Müller
,
T.
Klaus
, and
G.
Winnewisser
,
Astron. Astrophys.
357
,
L65
(
2000
).
6.
A.
Wootten
,
E. P.
Bozyan
,
D. P.
Garrett
,
R. B.
Loren
, and
R. L.
Snell
,
Astrophys. J.
239
,
844
(
1980
).
7.
D.
Teyssier
,
D.
Fosse
,
M.
Gerin
,
J.
Pety
,
A.
Abergel
, and
E.
Roueff
,
Astron. Astrophys.
417
,
135
(
2004
).
8.
A.
Dutrey
,
S.
Guilloteau
, and
M.
Guélin
,
Astron. Astrophys.
317
,
L55
(
1997
).
9.
M.
Padovani
,
C. M.
Walmsley
,
M.
Tafalla
,
D.
Galli
, and
H. S. P.
Müller
,
Astron. Astrophys.
505
,
1199
(
2009
).
10.
H.
Beuther
,
D.
Semenov
,
T.
Henning
, and
H.
Linz
,
Astrophys. J. Lett.
675
,
L33
(
2008
).
11.
A. H.
Saleck
,
R.
Simon
,
G.
Winnewisser
, and
J. G. A.
Wouterloot
,
Can. J. Phys.
72
,
747
(
1994
).
12.
W. B.
Yan
,
C. B.
Dane
,
D.
Zeitz
,
J. L.
Hall
, and
R. F.
Curl
,
J. Mol. Spectrosc.
123
,
486
(
1987
).
13.
W. B.
Yan
,
J. L.
Hall
,
J. W.
Stephens
,
M. L.
Richlow
, and
R. F.
Curl
,
J. Chem. Phys.
86
,
1657
(
1987
).
14.
C.
Pfelzer
,
M.
Havenith
,
M.
Peric
,
P.
Murtz
, and
W.
Urban
,
J. Mol. Spectrosc.
176
,
28
(
1996
).
15.
H.
Kanamori
,
K.
Seki
, and
E.
Hirota
,
J. Chem. Phys.
87
,
73
(
1987
).
16.
H.
Kanamori
and
E.
Hirota
,
J. Chem. Phys.
89
,
3962
(
1988
).
17.
H.
Kanamori
,
T.
Amano
, and
E.
Hirota
,
J. Mol. Spectrosc.
131
,
58
(
1988
).
18.
J. W.
Stephens
,
W. B.
Yan
,
M. L.
Richnow
,
H.
Solka
, and
R. F.
Curl
,
J. Mol. Struct.
190
,
41
(
1988
).
19.
E. N.
Sharp-Williams
,
M. A.
Roberts
, and
D. J.
Nesbitt
,
J. Chem. Phys.
134
,
064314
(
2011
).
20.
Y.-C.
Hsu
,
J. J.-M.
Lin
,
D.
Papousek
, and
J. J.
Tsai
,
J. Chem. Phys.
98
,
6690
(
1993
).
21.
W. B.
Yan
and
T.
Amano
,
J. Chem. Phys.
99
,
4312
(
1993
).
22.
W.-Y.
Chiang
and
Y.-C.
Hsu
,
J. Chem. Phys.
111
,
1454
(
1999
).
23.
M.
Peric
,
R. J.
Buenker
, and
S. D.
Peyerimhoff
,
Mol. Phys.
71
,
673
(
1990
).
24.
R.
Tarroni
and
S.
Carter
,
J. Chem. Phys.
119
,
12878
(
2003
).
25.
J. J.
Keady
and
K. H.
Hinkle
,
Astrophys. J.
331
,
546
(
1988
).
26.
F. F. S.
van der Tak
,
J. H.
Black
,
F. L.
Schöier
,
D. J.
Jansen
, and
E. F.
van Dishoeck
,
Astron. Astrophys.
468
,
627
(
2007
).
27.
F.
Lique
,
R.
Toboła
,
J.
Kłos
,
N.
Feautrier
,
A.
Spielfiedel
,
L. F. M.
Vincent
,
G.
Chałasiński
, and
M. H.
Alexander
,
Astron. Astrophys.
475
,
567
(
2008
).
28.
A.
Spielfiedel
,
N.
Feautrier
,
F.
Najar
,
D.
Ben Abdallah
,
F.
Dayou
,
M. L.
Senent
, and
F.
Lique
,
Mon. Not. R. Astron. Soc.
421
,
1981
(
2012
).
29.
M. H.
Alexander
and
P. J.
Dagdigian
,
J. Chem. Phys.
83
,
2191
(
1985
).
30.
J. H.
Kastner
,
C.
Qi
,
U.
Gorti
,
P.
Hily-Blant
,
K.
Oberg
,
T.
Forveille
,
S.
Andrews
, and
D.
Wilner
,
Astrophys. J.
806
,
75
(
2015
).
31.
S.
Cuadrado
,
J. R. J. R.
Goicoechea
,
P.
Pilleri
,
J.
Cernicharo
,
A.
Fuente
, and
C.
Joblin
,
Astron. Astrophys.
575
,
A82
(
2015
).
32.
Z.
Nagy
,
V.
Ossenkopf
,
F. F. S.
van der Tak
,
A.
Faure
,
Z.
Makai
, and
E. A.
Bergin
,
Astron. Astrophys.
578
,
A124
(
2015
).
33.
E. A.
Bergin
,
F.
Du
,
L. I.
Cleeves
,
G. A.
Blake
,
K.
Schwarz
,
R.
Visser
, and
K.
Zhang
,
Astrophys. J.
831
,
101
(
2016
).
34.
Z.
Pan
,
D.
Li
,
Q.
Chang
,
L.
Qian
,
E. A.
Bergin
, and
J.
Wang
,
Astrophys. J.
836
,
194
(
2017
).
35.
M.
Wernli
,
L.
Wiesenfeld
,
A.
Faure
, and
P.
Valiron
,
Astron. Astrophys.
464
,
1147
(
2007
).
36.
Y.
Kalugina
,
F.
Lique
, and
J.
Kłos
,
Mon. Not. R. Astron. Soc.
422
,
812
(
2012
).
37.
F.
Najar
,
D.
Ben Abdallah
,
A.
Spiefiedel
,
F.
Dayou
,
F.
Lique
, and
N.
Feuatrier
,
Chem. Phys. Lett.
614
,
251
(
2014
).
38.
F.
Dumouchel
,
F.
Lique
,
A.
Spielfiedel
, and
N.
Feautrier
,
Mon. Not. R. Astron. Soc.
471
,
1849
(
2017
).
39.
X.
Zhang
,
Y.-H.
Ding
,
Y.-S.
Li
,
X.-R.
Huang
, and
C.-C.
Sun
,
J. Phys. Chem. A
104
,
8375
(
2000
).
40.
J.
Peeters
,
B.
Ceursters
,
H. M. T.
Nguyen
, and
M. T.
Nguyen
,
J. Chem. Phys.
118
,
3700
(
2002
).
41.
L.-P.
Ju
,
X.
Xie
,
X.
Zhang
, and
K.-L.
Han
,
Chem. Phys. Lett.
409
,
249
(
2005
).
42.
D.
Wang
and
W. M.
Huo
,
J. Chem. Phys.
127
,
154304
(
2007
).
43.
A.
Matsugi
,
K.
Suma
, and
A.
Miyoshi
,
Phys. Chem. Chem. Phys.
13
,
4022
(
2011
).
44.
M.
Jeziorska
,
P.
Jankowski
,
K.
Szalewicz
, and
B.
Jeziorski
,
J. Chem. Phys.
113
,
2957
(
2000
).
45.
M.
Bogey
,
C.
Demuynck
, and
J. L.
Destombes
,
Mol. Phys.
66
,
955
(
1989
).
46.
O.
Denis-Alpizar
,
T.
Stoecklin
,
P.
Halvick
, and
M.-L.
Dubernet
,
J. Chem. Phys.
139
,
034304
(
2013
).
47.
C.
Hampel
,
K. A.
Petersen
, and
H.-J.
Werner
,
Chem. Phys. Lett.
190
,
1
(
1992
).
48.
M.
Deegan
and
P. J.
Knowles
,
Chem. Phys. Lett.
227
,
321
(
1994
).
49.
H.-J.
Werner
,
P. J.
Knowles
,
G.
Knizia
,
F. R.
Manby
,
M.
Schütz
 et al., molpro, version 2012.1, a package of ab initio programs, 2012, see http://www.molpro.net.
50.
S. F.
Boys
and
F.
Benardi
,
Mol. Phys.
100
,
65
(
2002
).
51.
T. H.
Dunning
, Jr.
,
J. Chem. Phys.
90
,
1007
(
1989
).
52.
F.-M.
Tao
and
Y.-K.
Pan
,
J. Chem. Phys.
97
,
4989
(
1992
).
53.
H.
Koch
,
B.
Fernandez
, and
O.
Christiansen
,
J. Chem. Phys.
108
,
2784
(
1998
).
54.
C.
Rist
and
A.
Faure
,
J. Math. Chem.
50
,
588
(
2012
).
55.
S.
Green
,
J. Chem. Phys.
62
,
2271
(
1975
).
56.
P. E. S.
Wormer
,
J. A.
Kłos
,
G. C.
Groenenboom
, and
A.
van der Avoird
,
J. Chem. Phys.
122
,
244325
(
2005
).
57.
Q.
Ma
,
J.
Kłos
,
M. H.
Alexander
,
A.
van der Avoird
, and
P. J.
Dagdigian
,
J. Chem. Phys.
141
,
174309
(
2014
).
58.
P. J.
Dagdigian
,
J. Chem. Phys.
145
,
114301
(
2016
).
59.
HIBRIDON is a package of programs for the time-independent quantum treatment of inelastic collisions and photodissociation written by
M. H.
Alexander
,
D. E.
Manolopoulos
,
H.-J.
Werner
,
B.
Follmeg
,
P. J.
Dagdigian
, and others. More information and/or a copy of the code can be obtained from the website http://www2.chem.umd.edu/groups/alexander/hibridon.
60.
K. P.
Huber
and
G.
Herzberg
,
Molecular Spectra and Molecular Structure IV. Constants of Diatomic Molecules
(
Van Nostrand Reinhold
,
New York
,
1979
).
61.
E. F.
Greene
and
A.
Kuppermann
,
J. Chem. Educ.
45
,
361
(
1968
).
62.
F.
Lique
and
A.
Faure
,
J. Chem. Phys.
136
,
031101
(
2012
).
63.
F.
Lique
,
J. Chem. Phys.
142
,
241102
(
2015
).
64.
M. H.
Alexander
,
J. Chem. Phys.
76
,
3637
(
1982
).
65.
H. C.
Schewe
,
Q.
Ma
,
N.
Vanhaecke
,
X.
Wang
,
J.
Kłos
,
M. H.
Alexander
,
S. Y. T.
van de Meerakker
,
G.
Meijer
,
A.
van der Avoird
, and
P. J.
Dagdigian
,
J. Chem. Phys.
142
,
204310
(
2015
).
66.
Q.
Ma
,
A.
van der Avoird
,
J.
Loreau
,
M. H.
Alexander
,
S. Y. T.
van de Meerakker
, and
P. J.
Dagdigian
,
J. Chem. Phys.
143
,
044312
(
2015
).
67.
M.
Hernández Vera
,
Y.
Kalugina
,
O.
Denis-Alpizar
,
T.
Stoecklin
, and
F.
Lique
,
J. Chem. Phys.
140
,
224302
(
2014
).
68.
A.
Klein
,
Y.
Shagam
,
W.
Skomorowski
,
M.
Zuchowski
,
P. S.
Pawlak
,
L. M. C.
Janssen
,
N.
Moiseyev
,
S. Y. T.
van de Meerakker
,
A.
van der Avoird
,
C. P.
Koch
, and
E.
Narevicius
,
Nat. Phys.
13
,
35
(
2017
).
69.
R. N.
Zare
,
Angular Momentum
(
Wiley
,
New York
,
1988
).

Supplementary Material