A&A 464, 1147-1154 (2007)
DOI: 10.1051/0004-6361:20066112
M. Wernli - L. Wiesenfeld - A. Faure - P. Valiron
Laboratoire d'Astrophysique de l'Observatoire de Grenoble, UMR 5571 CNRS/UJF, Université Joseph-Fourier, Boîte postale 53, 38041 Grenoble Cedex 09, France
Received 26 July 2006 / Accepted 20 November 2006
Abstract
Aims. Rates for rotational excitation of
by collisions with He atoms and H2 molecules are computed for kinetic temperatures in the range 5-20 K and 5-100 K, respectively.
Methods. These rates are obtained from extensive quantum and quasi-classical calculations using new accurate potential energy surfaces (PES). The
-He PES is in excellent agreement with the recent literature. The
-H2 angular dependence is approximated using 5 independent H2 orientations. An accurate angular expansion of both PES suitable for low energy scattering is achieved despite the severe steric hindrance effects by the
rod.
Results. The rod-like symmetry of the PES strongly favours even
transfers and efficiently drives large
transfers. Despite the large dipole moment of
,
rates involving ortho-H2 are very similar to those involving para-H2, because of the predominance of the geometry effects. Except for the even
propensity rule, quasi classical calculations are in excellent agreement with close coupling quantum calculations. As a first application, we present a simple steady-state population model that shows population inversions for the lowest HC3N levels at H2 densities in the range 104-106 cm-3.
Conclusions. The
molecule is large enough to present an original collisional behaviour where steric hindrance effects hide the details of the interaction. This finding, combined with the fair accuracy of quasi classical rate calculations, is promising in view of collisional studies of larger molecules.
Key words: molecular data - molecular processes
Cyanopolyyne molecules, with general formula HC2n+1N, ,
have been detected in a great variety of astronomical
environments and belong to the most abundant species in cold and
dense interstellar clouds (Bell et al. 1997). One of these,
HC11N, is currently the largest unambiguously detected
interstellar molecule (Bell & Matthews 1985). The simplest one,
(cyanoacetylene), is the most abundant of the family. In addition
to interstellar clouds,
has been observed in circumstellar
envelopes (Pardo et al. 2004), in the Saturn satellite Titan
(Kunde et al. 1981), in comets (Bockelée-Morvan et al. 2000) and in extragalactic
sources (Mauersberger et al. 1990). Furthermore,
has been detected
both in the ground level and in excited vibrational levels, thanks
to the presence of low-lying bending modes
(e.g. Wyrowski et al. 2003). Because of its low rotational constant
and large dipole moment, cyanoacetylene lines are observable over
a wide range of excitation energies and
is therefore
considered as a very good probe of physical conditions in many
environments.
Radiative transfer models for the interpretation of observed
spectra require the knowledge of collisional excitation rates
participating in line formation. To the best of our knowledge, the
only available collisional rates are those of Green & Chapman (1978) for
the rotational excitation of HC3N by He below 100 K. In cold
and dense clouds, however, the most abundant colliding partner is H2. In such environments, para-
is only populated in the
J=0 level and may be treated as a spherical body.
Green & Chapman (1978) and Bhattacharyya & Dickinson (1982) postulated that the
collisional cross-sections with para-
(J=0) are similar to
those with He (assuming thus an identical interaction and
insensitivity of the scattering to the reduced mass). As a result,
rates for excitation by para-
were estimated by scaling the
rates for excitation by He; rates involving ortho-
were not
considered.
In the present study, we have computed new rate coefficients for
rotational excitation of
by He, para-
(J=0) and
ortho-
(J=1), in the temperature range 5-20 K for He and
5-100 K for H2. A comparison between the different partners
is presented and the collisional selection rules are investigated
in detail. The next section describes details of the PES calculations. The cross-section and rate calculations are presented in Sect. 3. A discussion and a first application of these rates is given in Sect. 4. Conclusions are drawn in Sect. 5. The following units are used throughout except where otherwise stated: bond lengths and
distances in Bohr; angles in degrees; energies in cm-1; cross-sections in
.
Two accurate interatomic potential energy surfaces (PES) have
recently been calculated in our group, for the interaction of
with He and H2. Both surfaces involved the same geometrical
setup and similar ab initio accuracy. An outline of those
PES is given below and a detailed presentation will be published
in a forthcoming article.
In the present work, we focus on low-temperature collision rates,
well below the threshold for the excitation of the lower bending
mode
at 223 cm-1. The collision partners may thus
safely be approximated to be rigid, in order to keep the number of
degrees of freedom as small as possible. For small van der Waals
complexes, previous studies have suggested that properly averaged
molecular geometries provide a better description of experimental
data than equilibrium geometries (
geometries)
(Jeziorska et al. 2000; Jankowski & Szalewicz 2005). For the
-
system,
geometries averaged over ground-state vibrational wave-functions
(r0 geometries) were shown to provide an optimal approximation
of the effective interaction (Faure et al. 2005a; Wernli 2006).
Accordingly, we used the
bond separation
Bohr obtained by averaging over the ground-state vibrational
wave-function, similarly to previous calculations
(Hodges et al. 2004; Faure et al. 2005a; Wernli et al. 2006). For
,
as vibrational
wave-functions are not readily available from the literature, we
resorted to experimental geometries deduced from the rotational
spectrum of
and its isotopologues (Thorwirth et al. 2000; see also
Table 5.8 in Gordy & Cook 1984). The resulting bond separations are the
following:
;
;
;
,
and should be close to vibrationally averaged values.
For the
-He collision, only two coordinates are needed to fully
determine the overall geometry. Let
be the vector between
the center of mass of
and He. The two coordinates are the distance
and the angle
between the
rod and the
vector R. In our conventions,
corresponds to an approach towards the H end of the
rod. For the
collision with H2, two more angles have to be added,
and
,
that respectively orient the
molecule in the rod-R plane and out of the plane. The
-He PES
has thus two degrees of freedom, the
-
four degrees of freedom.
As we aim to solve close coupling equations for the scattering, we
need ultimately to expand the PES function V over a suitable angular
expansion for any intermolecular distance R. In the simpler case of
the
-He system, this expansion is in the form:
For the
-
system, the expansion becomes:
Because the Legendre polynomials form a complete set, such
expansions should always be possible. However, Chapman & Green (1977)
failed to converge the above expansion (1) due to the
steric hindrance of He by the impenetrable
rod, and
Green & Chapman (1978) abandoned quantum calculations, resorting to quasi
classical trajectories (QCT) studies. Similar difficulties arise
for the interaction with H2. As can be seen in Fig. 1 for small R values, the interaction is moderate or possibly weakly attractive for
and is extremely repulsive or undefined for
,
leading to singularities in the angular
expansion and severe Gibbs oscillations in the numerical fit of
the PES over Legendre expansions.
Accordingly, we resorted to a cautious sampling strategy for the PES, building a spline interpolation in a first step, and postponing the troublesome angular Legendre expansion to a second step. All details will be published elsewhere. We summarize this first step for He, then for H2.
![]() |
Figure 1:
The
![]() ![]() ![]() ![]() ![]() ![]() |
Open with DEXTER |
For the
-He PES, we selected an irregular grid in the
coordinates. The first order
derivatives of the angular spline were forced to zero for
in order to comply with the PES symmetries. For each distance, angles were added until a smooth convergence of the angular spline fit was achieved, resulting in
typical angular steps between 2 and 15
.
Then, distances
were added until a smooth bicubic spline fit was obtained,
amounting to 38 distances in the range 2.75-25 Bohr and a total
of 644 geometries. The resulting PES is perfectly suited to run
quasi classical trajectories.
We used a similar strategy to describe the interaction with H2,
while minimizing the number of calculations. We selected a few
orientation sets, bearing in mind
that the dependence of the final PES on the orientation of
is
weak. In terms of spherical harmonics, the PES depends only on
,
with
and
,
.
Terms in
Yl2m2 and
Yl2 -m2 are equal by symmetry. Previous studies (Faure et al. 2005b; Wernli et al. 2006) have shown that terms with l2> 2 are
small, and we consequently truncated the
Yl2m2 series to
.
Hence, only four basis functions remain for the
orientation of
:
Y00,Y20,Y21 and Y22.
Under this assumption, the whole
-
surface can be
obtained knowing its value for four sets of
angles at each value of R. We selected five sets,
thus having an over-determined system allowing for the monitoring
of the accuracy of the l2 truncation. Consequently, we
determined five independent PES, each constructed similarly
to the
-He one as a bicubic spline fit over an irregular
grid in
coordinates. The angular mesh
is slightly denser than for the
-He PES for small R distances to account for more severe steric hindrance effects involving
.
In total, we computed 3420
geometries. The
-H2 interaction
can be readily reconstructed from these five PES by expressing its
analytical dependence over
(Wernli 2006).
For each value of the intermolecular geometry
or
,
the intermolecular potential energy is calculated at
the conventional CCSD(T) level of theory, including the usual
counterpoise correction of the Basis Set Superposition Error
(Jansen & Ross 1969; Boys & Bernardi 1970). We used augmented correlation-consistent
atomic sets of triple zeta quality (Dunning's aug-cc-pVTZ) to describe
the
rod. In order to avoid any possible steric hindrance problems
at the basis set level, we did not use bond functions and instead
chose larger Dunning's aug-cc-pV5Z and aug-cc-pVQZ basis set to
better describe the polarizable (He, H2) targets, respectively. All
calculations employed the direct parallel code DIRCCR12
(Noga et al. 2003-2006).
Comparison of the
-He PES with existing surfaces
(Akin-Ojo et al. 2003; Topic & Wolfgang 2005) showed an excellent agreement. The
-para-
(J=0) interaction (obtained by averaging the
-H2 PES over
and
)
is qualitatively similar to
the
-He PES with a deeper minimum (see values at the end of
present section). As illustrated in Fig. 1, these
PES are largely dominated by the rod-like shape of
,
implying a prolate ellipsoid symmetry of the equipotentials.
In a second step, we consider how to circumvent the difficulty of the angular expansion of the above PES, in order to obtain reliable expansions for He and H2 (Eqs. (1) and (2)).
Using the angular spline representation, we first expressed each
PES over a fine
mesh suitable for a subsequent high l1 expansion. As expected from the work of Chapman & Green (1977),
high l1 expansions (1) resulted in severe Gibbs
oscillations for R in the range 5-7 Bohr, making impossible the
description of the low energy features of the PES. For low energy
scattering applications, we regularized the PES by introducing a scaling function
.
We replaced
by
,
where
returns V when V is lower than a prescribed threshold, and then smoothly saturates to
a limiting value when V increases to very repulsive values.
Consequently, the regularized PES retains only the low energy
content of the original PES, unmodified up to the range of the
threshold energy; it should not be used for higher collisional
energies. However, in contrast to the original PES, it can be
easily expanded over Legendre functions to an excellent accuracy
and is thus suitable for quantum close coupling studies. We
selected a threshold value of 300 cm-1, and improved the
quality of the expansion by applying a weighted fitting strategy
(e.g. Hodges et al. 2004) to focus the fit on the details of the
attractive and weakly repulsive regions of the PES. Using
,
both the He and H2 PES fits were converged to within 1 cm-1 for
cm-1. These expansions still describe the range
300<V<1000 cm-1 to within an accuracy of
a few
.
The corresponding absolute minima are the following (in cm-1 and
Bohr): for
-He, V=-40.25 for R=6.32 and
;
for
-para-
(J=0), V=-111.24 for
R=6.41 and
;
and for
-
,
V=-192.49for R=9.59,
,
and
.
In the following,
,
denote the initial and final
angular momentum of the
molecule, and J2 denotes the
angular momentum of H2. We also denote the largest value of
as
.
The most reliable way to compute inelastic cross sections
is to perform quantum close
coupling calculations. In the case of molecules with a small
rotational constant, like
(
,
see
e.g. Thorwirth et al. 2000), quantum calculations soon become intractable,
because of the large number of open channels involved. While
observations at cm-mm wavelengths culminate with
(Audinos et al. 1994), sub-mm observations can probe
transitions as high as
,
at a frequency of
363.785 GHz and a rotational energy of
(Kuan et al. 2004; Caux 2006; Pardo et al. 2004). It is thus necessary to compute rates with transitions up to J1=50 (E=
386.8 cm-1), in order to properly converge radiative
transfer models. Also, we aim to compute rates up to a temperature of 100 K for H2. We used two methods. For
,
we performed quantum inelastic scattering
calculations, as presented in next Sect. 3.1. For
,
we used the QCT method, as presented in Sect. 3.2.
For He, of less astrophysical importance ([He]/[H]
0.1), only
quantum calculations were performed and were limited to the low
temperature regime (T = 5-20 K and J1<10).
All calculations were made using the rigid rotor approximation,
with rotational constants
cm-1 and
cm-1, using the MOLSCAT code
(Hutson & Green 1994). All quantum calculations for
-ortho-
were performed with
.
Calculations for
-para-
were performed with J2=0. We checked at
that the
inclusion of the closed J2=2 channel had negligible effects.
The energy grid was adjusted to reproduce all the details of the
resonances, as they are essential to calculate the rates with high
confidence (Dubernet & Grosjean 2002; Grosjean et al. 2003; Wernli et al. 2006). The energy grid
and the quantum methods used are detailed in
Table 1. Using this grid, we calculated the whole
resonance structure of all the transitions up to J1=15 for the
-para-
collisions. At least 10 closed channels were
included at each energy to fully converge the
rotational
basis. We used the hybrid log-derivative/Airy propagator
(Alexander & Manolopoulos 1987). We increased the parameter STEPS at
the lowest energies to constrain the step length of the integrator
below 0.1 to 0.2 Bohr, in order to properly follow the details of
the radial coefficients. Other propagation parameters were taken
as the MOLSCAT default values.
Table 1:
Details of the quantum MOLSCAT cross section
calculation parameters.
(10 for He). Methods: CC, Close coupling, CS, Coupled states approx., IOS, Infinite Order Sudden approx.
Two examples of deexcitation cross-sections are shown in Fig. 2. We see that for energies between threshold and about 20 cm-1 above threshold, the cross-section displays many shape resonances, justifying a posteriori our very fine energy grid. This behaviour is similar to most earlier calculations is many different systems, see e.g. Dubernet & Grosjean (2002); Wernli et al. (2006) for a discussion. From a semi-classical point of view, these shape resonances reflect the trapping of the wave-packet between the inner repulsive wall and the outer centrifugal barrier, see Wiesenfeld et al. (2003), Abrol et al. (2001). At energies higher than about 20 cm-1 above threshold, all cross-sections become smooth functions of the energy.
Figure 2 also shows that ortho-
inelastic
cross-sections very closely follow the para-
ones, including
the position of resonances. Examination of all cross-sections
reveals that the relative difference between
and
is less than
.
This justifies a posteriori the much smaller amount of computational effort devoted to ortho-
collisions as
well as the neglect of J2 = 2 closed para-
channels. A detailed discussion of this behaviour is given in Sect. 4.1.
![]() |
Figure 2:
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
Open with DEXTER |
The quantum collisional rates are calculated for
at astrophysically relevant temperatures, from 5 K to 100 K. We
average the cross-sections described in the preceding section over
the Maxwell distribution of velocities, up to a kinetic energy at
least 10 times
.
The quantum calculations at the higher end of
the energy range are approximated at the IOS level (see Table 1), which is justified at these energies by the
smallness of the rotational constant
.
Also we used
a coarse energy grid for the IOS calculations because the energy
dependence of the cross-sections becomes very smooth.
For values of J1 > 15, the close coupling approach enters a complexity barrier due to the rapid increase of the number of channels involved in calculations, while memory and CPU requirements scale as the square and the cube of this number, respectively. Resorting to quantum CS or IOS approximations is inaccurate, because the energy is close to threshold for high-J1 channels. In the meanwhile, the accuracy of classical approximations improves for higher collisional energies. For the energy range where J1 > 15 channels are open and for deexcitation processes involving those channels, we employ a Quasi-Classical Trajectory (QCT) method, which has been shown to be a valid approximation for higher collisional energies and high rates (Lepp et al. 1995; Chapman & Green 1977; Mandy & Progrebnya 2004; Faure et al. 2006).
For Monte-Carlo QCT methods, we must devise a way of defining an ensemble of initial conditions for classical trajectories, on the
one hand, and of analyzing the final state of each trajectory, on
the other hand. Contrary to the asymmetric rotor case (like
water, see Faure & Wiesenfeld 2004), the analysis of final conditions for a linear molecule is straightforward. Using the simplest quantization approximation, we bin the final classical angular
momentum J'1 of
to the nearest integer. While the quantum
formalism goes through a microcanonical calculation - calculating
for fixed energies, then averaging over
velocity distributions - it is possible for QCT calculations to
directly resort to a canonical formalism, i.e. to select the
initial velocities of the Monte-Carlo ensemble according to the
relevant Maxwell-Boltzmann distribution and find the rates as:
Some results are shown in Tables 2 and 3 and are illustrated in Figs. 3 and 4.
As an alternative to QCT calculations, we tested J-extrapolation
techniques, using the form of DePristo et al. (1979) generally used by
astrophysicists (see for example Schöier et al. (2005), Sect. 6). We
found that even if it reproduces the interference pattern, the
extrapolation systematically underestimates the rates, for .
Hence, QCT rates are more precise on average.
Table 2:
-para-
(J=0) collisions. Quantum deexcitation
rates in
,
for J1'=0, for successive
initial J1 and for various temperatures. Powers of ten are
denoted in parentheses.
Table 3:
-para-
collisions. Quantum or QCT (
)
deexcitation rates in
,
for
J1-J'1=1,2,3,4, for various temperatures and
representative values of J1. Powers of ten are denoted in
parentheses.
For H2, all deexcitation rates
kJ1J1'(T),
,
are fitted with the following formula (Wernli et al. 2006):
![]() |
Figure 3:
Collisional excitation rates for
![]() ![]() ![]() ![]() ![]() |
Open with DEXTER |
![]() |
Figure 4:
![]() ![]() |
Open with DEXTER |
A comparison of the
cross sections for
with ortho-
and
para-
is given in Fig. 2. The difference
between the two spin species of
is tiny, in any case smaller
than other PES and cross-section uncertainties.
This is an unexpected result, as sizeable differences between
para-
and ortho-
inelastic cross-sections exist for other
molecules. These differences were expected to increase for a molecule possessing a large dipolar moment, in view of the results
obtained for the C2 molecule (Phillips 1994), the CO molecule (Wernli et al. 2006), the OH radical (Offer et al. 1994), the NH3 molecule (Offer & Flower 1989; Flower & Offer 1994) and the
molecule (Phillips et al. 1996; Dubernet & Grosjean 2002; Dubernet et al. 2006; Grosjean et al. 2003), due
to the interaction between the dipole of the molecule and the
quadrupole of
(J2 > 0).
This apparently null result deserves an explanation. We focus on
Eq. (9) of Green (1975). This equation describes the different
matrix elements that couple the various channels in the close-coupling
equations. Some triangle rules apply which restrict the number of
terms in the sum of Eq. (9); the relevant angular coupling
algebra is represented there as a sum of terms of the type
![]() |
Figure 5:
Comparison of the coupling terms (![]() ![]() ![]() ![]() ![]() |
Open with DEXTER |
In Fig. 3, we compare the various rates that
we obtain here with the ones previously published by
Green & Chapman (1978). These authors used a coarse electron-gas
approximation for the PES, and computed rates by a QCT classical
approach. Despite these approximations, we see that the rates
obtained by Green & Chapman (1978) are qualitatively comparable with the
quantum rates obtained here, on average. However, as Table 3 and Figs. 4 and 3 show clearly, only quantum calculations
manifest the strong
propensity rule. This rule
originates in the shape of the PES, being nearly a prolate
ellipsoid, dominated by the rod shape of
and not
dominated by the large dipole of HC3N molecule (3.724 Debye).
Because of the very good approximate symmetry
,
the l1 even terms
(Eq. (6) and Green 1975) are the most
important ones, directing the inelastic transition toward even
values. This propensity has also been explained
semi-classically by McCurdy & Miller (1977) in terms of an interference
effect related to the even anisotropy of the PES. These authors
show in particular that the reverse propensity can also occur if
the odd anisotropy of the PES is sufficiently large. This reverse
effect is indeed observed in Fig. 4 for transitions
with
.
A similar propensity rule has been
experimentally observed for CO-He collisions (Carty et al. 2004).
Besides this strong
propensity rule, one can see
from Table 3 and Figs. 3,
4 that the rod-like interaction drives large
transfers. For instance, for T > 20 K, rates for
are generally larger than rates for
,
and rates for
are only one order of magnitude below
those for
.
This behaviour is likely to emphasize
the role of collisional effects versus radiative ones. This
effect, of purely geometric origin, has been predicted previously
(Bosanac 1980) and is of even greater importance for longer
rods like
,
,
,
see
Snell et al. (1981), Bhattacharyya & Dickinson (1982).
We also observe that the ratio
is in
average close to
,
thus confirming the
similarity of He and para-
as projectiles, as generally
assumed. But it is also far from being a constant, as already
observed for H2O (Phillips et al. 1996) or CO (Wernli et al. 2006). Our
data shows that the
scaling rule results in errors of
up to 50%.
Because of the strong
propensity rule,
population inversion could be strengthened if LTE conditions are
not met, even neglecting hyperfine effects
(Hunt et al. 1999). In order to see the density conditions giving rise to population inversion, we
solved the steady-state equations for the population of the
levels of
,
including collisions with
(densities ranging from 102 to
), a black-body photon bath at 2.7 K, in the optically thin approximation, (Goldsmith 1972):
![]() |
Figure 6:
Population per
J1,mJ1-sub-level, following
Eq. (7), for varying ![]() ![]() |
Open with DEXTER |
![]() |
Figure 7:
![]() ![]() ![]() ![]() ![]() ![]() |
Open with DEXTER |
Similar effects should appear for the whole cyanopolyyne (
)
family, where cross-sections should scale
approximately with the rod length (Bhattacharyya & Dickinson 1982). It is
expected that the propensity rule
should
remain valid. Also, the critical density should decrease for the
higher members of the cyanopolyyne family, as the Einstein
Aij coefficients, hence facilitating the LTE conditions.
We have computed two ab initio surfaces for the
-He
and
-
systems. The latter was built using a carefully
selected set of
orientations, limiting the computational
effort to approximately five times the
-He one. Both
surfaces were successfully expanded on a rotational basis suitable
for quantum calculations using a smooth regularization of the
potentials. This approach circumvented the severe convergence
problems already noticed by Chapman & Green (1977) for such large
molecules. The final accuracy of both PES is a few cm-1 for
potential energies below 1000 cm-1.
Rates for rotational excitation of
by collisions with He atoms
and
molecules were computed for kinetic temperatures in the
range 5 to 20 K and 5 to 100 K, respectively, combining quantum
close coupling and quasi-classical calculations. The rod-like
symmetry of the PES strongly favours even
transfers
and efficiently drives large
transfers. Quasi
classical calculations are in excellent agreement with close
coupling quantum calculations but do not account for the even
interferences. For He, results compare fairly well
with Green & Chapman (1978) QCT rates, indicating a weak dependance on
the details of the PES. For para-H2, rates are compatible on
average with the generally assumed
scaling rule, with a spread of about 50%. Despite the large dipole moment of
,
rates involving ortho-H2 are very similar to
those involving para-H2, due to the predominance of the rod
interactions.
A simple steady-state population model shows population inversions
for the lowest
levels at
densities in the range
104-106 cm-3. This inversion pattern shows the
importance of large angular momentum transfer, and is enhanced by
the even
quantum propensity rule.
The
molecule is large enough to present an original
collisional behaviour, where steric hindrance effects hide the
details of the interaction, and where quasi classical rate
calculations achieve a fair accuracy even at low temperatures.
With these findings, approximate studies for large and heavy
molecules should become feasible including possibly the modelling
of large
transfer collisions and ro-vibrational
excitation of low energy bending or floppy modes.
Acknowledgements
This research was supported by the CNRS national program "Physique et Chimie du Milieu Interstellaire'' and the "Centre National d'Études Spatiales''. L.W. was partly supported by a CNRS/NSF contract. M.W. was supported by the Ministère de l'Enseignement Supérieur et de la Recherche. CCSD(T) calculations were performed on the IDRIS and CINES French national computing centers (projects no. 051141 and x2005 04 20820). MOLSCAT and QCT calculations were performed on local workstations and on the "Service Commun de Calcul Intensif de l'Observatoire de Grenoble'' (SCCI) with the valuable help of F. Roch.
Table A.1:
Fitting coefficients of HC3N-para-H2(J2 = 0) rates,
following formula (5). J1 and J'1 are the initial and
final rotation states, respectively. The rates thus obtained are
in
.
Rates are quantum for
,
classical otherwise.