1996MNRAS.281..137P Mon. Not. R. Astron. Soc. 281,137-144 (1996) Novel pathways to CN - within interstellar clouds and circumstellar envelopes: implications for IS and CS chemistry Simon Petrie* School of Chemistry, University College, University of New South Wales, ADFA, Canberra, ACT 2600, Australia Accepted 1996 January 18. Received 1996 January 18; in original form 1995 November 6 ABSTRACT Prospects for detection of the diatomic negative ion CN- within dense interstellar clouds and/or circumstellar envelopes are reassessed in the light of new pathways to small negative ions in such environments. Formation of CN- by charge transfer from PAH- to CN, and by dissociative attachment of free electrons to MgCN and MgNC, is discussed. These pathways are expected to lead to much higher CNabundances than can be formed by those reactions which have been included in previous models of interstellar negative-ion chemistry. Investigation of electron attachment to Mg(CN) is assisted byab initio calculations, at the G2level of theory, upon the structure and total energy of singlet and triplet states of MgCN- and MgNC - . The model of CN- chemistry which we present here shows that interstellar abundances of PAH- are probably too low to produce detectable CN-, but the efficient occurrence of dissociative attachment to MgNC and MgCN within the circumstellar envelope IRC + 10216 may produce CN- at an abundance of ",2 x 10- 10 n(Hz), which should permit the detection of CN- via its rotational spectrum. Key words: molecular processes - circumstellar matter - ISM: abundances - ISM: clouds - ISM: molecules. 1 INTRODUCTION Formation of small negative ions within dense interstellar (IS) clouds has been discussed previously by several authors (Dalgarno & McCray 1973; Sarre 1980; Herbst 1981). It is generally considered, however, that the role of small negative ions within gas-phase IS cloud chemistry is minimal, in contrast to the major contribution made by positive-ion chemistry. Negative-ion formation within IS clouds involves attachment of an electron to a neutral species: in its simplest form, this is a radiative process X+e--+X-+hv. (1) Reaction (1) is of negligible efficiency for small species X (e.g., 1-3 atoms) having few rotational and vibrational degrees of freedom available for energy dispersal, owing to the very short lifetime against autodetachment (X-)* --+X + e *E-mail: [email protected] (2) of the activated complex (X-)*. The lifetime of this complex increases steeply with increasing molecular complexity: Herbst (1981) contends that occurrence of radiative attachment (RA) to high-electron-affinity species as small as C 3N and C4H may be significant, and RA to polycyclic aromatic hydrocarbon (PAH) molecules PAH+e--+PAH- +hv (3) may well be efficient for several PAHs under both dense and diffuse IS cloud conditions (Omont 1986; Lepp & Dalgarno 1988). Radiative attachment to the fullerenes C60 and C70 has also been discussed by Millar (1992). Recent laboratory studies have, however, indicated that attachment of low-energy (T::;; 300 K) electrons to anthracene, C 14H lO (Canosa et al. 1994) and to C60 (Smith, Spanel & Mark 1993; laffke et al. 1994) is impeded by activation energy barriers, implying that RA to these species under IS cloud conditions will be very inefficient. Similar studies on C70 (Spanel & Smith 1994) suggest the absence of a barrier to one mode of electron attachment. Clearly, the topic of RA to molecules at the very low temperatures characteristic of IS clouds is one in which much remains to be explored. ©1996 RAS © Royal Astronomical Society • Provided by the NASA Astrophysics Data System 1996MNRAS.281..137P 138 S. Petrie The cyanogen radical CN is abundant within IS clouds for example, n(CN)=3 x 10- 8 n(Hz) within TMC-l (Irvine, Goldsmith & Hjalmarson 1987) - and larger species containing the CN group comprise an important subset of the species detected within dense IS clouds and circumstellar (CS) envelopes. Cyanogen also has a very high electron affinity, EA(CN)=3.82±0.02 eV (Berkowitz, Chupka & Walter 1969), probably the highest of any atomic or diatomic species. Furthermore, CN- is a polar species, ,u(CN-)~0.6 Debye (Botschwina 1985; Peterson & Woods 1989), which should be detectable by virtue of its rotational transitions: this contrasts with the situation for an atomic negative ion, whose identification requires detection of electronic absorption or emission features. The cyanide ion is therefore an ideal test species to investigate the factors affecting the formation of small negative ions within IS clouds and CS envelopes. We note that, while no negative ions have yet been detected within any IS or CS environments, the Giotto probe has detected several negative ions, having molecular masses in the ranges 7-19, 22-65 and 85-110 amu, as components of the inner coma of comet P/Halley (Chaizy et al. 1991): CN- has been implicated as a likely constituent of the 22-65 amu 'group' in this source. In the present work, we provide a reassessment of pathways to CN - formation under typical IS and CS conditions, and discuss the prospects for detection of this ion within objects such as TMC-l and IRC + 10216. predicts very low abundances for CN- as well as for the other small negative ions which are included. At 'early' time (3.16 x 105 yr), which is generally held to mark the peak abundance of many molecular species, n(CN-) = 1.28 x 10- 15 n(Hz) is obtained. This is far too low an abundance to permit the detection of CN - with current techniques - especially since, with ,u ~ 0.6 Debye (Botschwina 1985; Peterson & Woods 1989), CN- is not strongly polar. 2.2 Radiative attachment to CN The inclusion of reaction (10) as a source for CN- is not expected to increase n (CN -) significantly. This reaction has not been studied experimentally, but is likely to proceed at a rate similar to that of RA to an atomic species. In the UMIST model (Millar et al. 1991), RA rate coefficients of k ~ 3 x 10 -15 em3 molecule -I s -I are estimated for C, 0 and S. If a value of klO~ 1 x 10- 14 em3 molecule-I S-I is used for reaction (10), and the UMIST early-time abundances are used for other species in the above reactions, then the relative rates of CN- formation by the UMIST reactions and by reaction (10) are (r4+r5+r6+r7):rlO~2. A significantly higher rate coefficient for reaction (10) would, of course, favour the occurrence of this reaction, but the value of 1 x 10- 14 cm3 molecule-I S-I which we have suggested appears realistic. A more detailed mechanism for RA, CN+e-+(CN-)* (12) 2 DISCUSSION (CN-)* -+CN + e (13) 2.1 Existing models of CN - chemistry (CN-)* -+CN- +hv, (14) Of the complex kinetic models of dense cloud chemical evolution, only the UMIST ratefile of Millar et al. (1991) has included a treatment of negative-ion formation and destruction. This model, which includes only the simple negative ions H-, C-, 0-, S-, OH- and CN-, incorporates CN- formation via the charge-transfer reactions illustrates the competition between autodetachment (13) and radiative stabilization (14) of the activated complex (CN-)*. For a polyatomic species, a high density of vibrational states at the energy of the activated complex allows dispersal of the excess energy of (CN -) *, greatly extending its lifetime against autodetachment and increasing the probability for radiative stabilization. The CN - ion possesses only one vibrational model, with a calculated frequency v=2052±6 cm- I (Botschwina 1985). When anharmonicity is taken into consideration, this yields a density of states N ~6.0 x 10- 4 (cm-I)-I at the energy of the activated complex. Use of this density-of-states value, in the expression proposed by Herbst (1981) as an approximate measure of the RA rate coefficient, yields klo~ 1.4 x 10- 17 cm3 molecule-I S-I at T=10K. This value, of a similar order of magnitude to those employed by Millar et al. (1991) for RA to C, 0 and S atoms, serves to illustrate that vibrational dispersal of the energy of (CN-)* is unable to promote efficient RA to the CN radical. The above discussion indicates that RA is incapable of generating detectable CN- within IS environments. 0- +CN-+CN- +0 (4) OH- +CN-+CN- +OH (5) and the proton-transfer reactions 0- + HCN -+CN- + OH (6) OH- + HCN -+CN- + HzO. (7) Of these processes, (7) predominates as a CN - source by virtue of the higher abundances calculated for the reactants involved. Removal of CN - in the UMIST model is achieved by the rapid associative detachment reactions CN- +H-+HCN +e (8) CN- + CH 3 -+CH3CN + e. (9) This simple model, which neglects CN- formation by radiative attachment CN +e-+CN- +hv, (10) The inclusion of reactions (4) and (5) as sources of CN - in the UMIST model (Millar et al. 1991) is of interest, even though these reactions produce only very small CN- abundances. Charge transfer (11) (15) as well as its removal by mutual neutralization CN- +M+-+M+CN, 2.3 Charge transfer to CN © 1996 RAS, MNRAS 281,137-144 © Royal Astronomical Society • Provided by the NASA Astrophysics Data System 1996MNRAS.281..137P CN- within interstellar clouds is intrinsically more efficient than radiative stabilization (10): a common feature of charge-transfer reactions involving negative ions and neutral molecules is that they occur at high rates, k - kc (where kc is the collision rate coefficient) and without apparent activation energy barriers when exothermic (Kebade & Chowdhury 1987). The very low yield of CN- from reactions (4) and (5) in the UMIST model is a consequence of the very low abundances of the reactants 0- and OH- [n(0-)=1.7 x 10- 17 n(H2); n(OH-)=6.8 x 10- 14 n(H2)] in this model (Millar et al. 1991). A high electron affinity [EA(CN) =3.82 eV] ensures that reaction (15) is exothermic for many negative ions Y-. If suitably abundant reactant ions Y- can be identified, therefore, reaction (15) may be significant at typical dense IS cloud temperatures. The prime candidates for reactants capable of initiating reaction (15) are PAH- ions, and perhaps fullerene ions such as C6ii. We have noted in Section 1 that recent studies indicate the presence of activation energy barriers to electron attachment to anthracene and C60 : Canosa et al. (1994) have suggested that the rate coefficients used in previous models of interstellar PAH chemistry (e.g. Omont 1986; Lepp & Dalgarno 1988) for RA to PAHs may be too high by several orders of magnitude. However, the apparent 'lack of an activation barrier to s-wave electron capture by ~O, manifested by a negative temperature dependence for this process (Spanel & Smith 1994), illustrates the dangers of overgeneralizing from a small collection of measurements. It appears that RA to some PAHs and fullerenes will be very inefficient at IS cloud temperatures: it remains, however, also very likely that some PAHs will readily attach electrons under IS conditions. Further studies are needed on this matter. Here we assume that a total PAH- abundance of (1-5) x 10- 9 n (H2) is reasonable, this value being somewhat lower than the most optimistic estimates based on efficient RA to all PAHs (Omont 1986; Lepp & Dalgarno 1988). The composition of neutral, as well as negatively charged, PAHs within IS environments is still the subject of much conjecture: while PAHs as a class of compounds are widely held to be the carriers of the diffuse IS bands (Leger & d'Hendecourt 1985; van der Zwet & Allamandola 1985; Cossart-Magos & Leach 1990), no particular PAH has yet been conclusively identified as an IS molecule. Similarly, fullerenes such as Coo have also been proposed as probable large IS molecules (see, e.g., Hare & Kroto 1992; Webster 1992; Petrie, Javahery & Bohme 1993), and, while the formation of fullerene negative ions within dense clouds has been discussed by Millar (1992), evidence for the presence of fullerenes in IS environments is rather tenuous (Webster 1993; Foing & Ehrenfreund 1994). Nevertheless, an impression of the reactivity of interstellar PAHs and fullerenes can be gained from the general features of the chemistry of these compounds under laboratory conditions. With regard to PAH- reactivity, the most important parameter is the EA of the corresponding PAH molecule. We list representative values of EA (PAH) in Table 1. The exothermicity of charge transfer 139 Table 1. Electron affinities of polycyclic aromatic hydrocarbons (PAHs). PAHa Naphthalene Formula EA: Expt b Theor c CIOHS 0.15 d 0.07 0.26 CU#12 0.28 d Phenanthrene C14H IO 0.31d 0.27 Chrysene CIsHI2 0.42d 0.51 0.61 Triphenylene C20H 12 0.49 d Benzo[c]phenanthrene CIsHI2 O.54 d 0.38 Anthracene Cl~1O 0.57 e 0.65 Pyrene CIsHlO 0.58 d 0.66 Dibenz[a.h]anthracene C22H 14 O.68 d 0.65 Dibenz[a,j]anthracene Benzo[c]pyrene C22H I4 0.69 d 0.58 Benz[a]anthracene CIsHI2 0.70 d 0.64 Benz[a]pyrene CzoHI2 O.7g e 0.93 Perylene C2oH12 0.97 e Tetracene C1SH 12 1.04 e Pentacene C22H 14 1.35 e Buckminsterfullerene C60 1.06 >2.271 -2.7K Notes. ·PAH molecules, in order of increasing (experimental) EA. bEx_ perimental electron affinity in eV. cTheoretical electron affinity in eV, from molecular orbital calculations (Dewar, Hashmall & Venier 1968). dFrom Becker & Chen (1966). eFrom Crocker, Wang & Kebarle (1993). fFrom Sunderlin et al. (1991). "From Yang et al. (1987). for all of the PAHs listed; charge transfer from C6ii is also exothermic, but considerably less so. These values suggest that all charge-transfer reactions of odd-electron PAH - ions to CN are probably exothermic, with the potential to be efficient under IS cloud conditions. Consideration should also be given to charge transfer to CN from even-electron PAH- ions, since such species may also account for a significant fraction of PAH- ions in IS clouds. However, very few thermochemical data exist on such species. The tabulation of Lias et al. (1988) indicates only that charge transfer from the fluorenide ion (17) is exothermic by -2.0 eV. The product channel (16) is not the only possible outcome of PAH-/CN reactions. Other channels which may compete with charge transfer from odd-electron PAH - ions are radiative association PAH- +CN-+PAH.CN- +hv (18) and hydrogen-atom transfer CnH';;- + CN -+CnH';;-_1 + HCN, (19) (16) while associative detachment is a probable competing process for both odd- and even-electron PAH- ions: is given by -MlI6=EA(CN)-EA(PAH). The values in Table 1 show that charge transfer is exothermic by ~ 2.4 e V (20) © 1996 RAS, MNRAS 281, 137-144 © Royal Astronomical Society • Provided by the NASA Astrophysics Data System 1996MNRAS.281..137P 140 S. Petrie No such reactions involving CN have yet been studied experimentally. There are grounds for expecting that reactions (18) and (19) are unlikely to dominate over (16) when the latter is considerably exothermic, although competition from channel (20) may be important in limiting the efficiency of charge transfer (16). Association (channel 18}. Model calculations suggest that radiative association reactions involving many-atom species are usually efficient under IS conditions (Dunbar 1990; Herbst & Dunbar 1991). The principal requirements for efficient radiative association are that the collision complex potential energy well is sufficiently deep, and the coupling between internal modes of energy dispersal is sufficiently strong, to promote a long-lived collision complex. These conditions may well not be met in the case of reactions of type (18), since PAHs possess (for their size) comparatively few low-energy bending or stretching vibrational modes available for energy dispersal. The product channel for reaction (18) presumably involves a collison complex featuring the interaction of CN- with the 1t cloud above one face ofthe PAH. Short-range repulsive forces between the negative charge on CN- and the 1t cloud of the PAH will limit the collision complex well depth, and are hence likely to disfavour adduct formation in the presence of a competing exothermic product channel (16). Furthermore, and regardless of the nature of the collision complex, the coupling between the C-C stretching vibrational modes (which are associated with the PAH's 1t electron cloud) and the C-H modes (which are not) is likely to be weak, further limiting the accessible degrees of freedom for complex stabilization. Hydrogen transfer (channel 19). Grounds can also be established for expecting that reaction (19) is inefficient, especially in comparison to the exothermic transfer of an electron (16). Hydrogen atom transfer necessitates an interaction between CN and a C-H bond of PAH -, but CN should be preferentially drawn (by ion-induced dipole interactions) towards the 1t cloud of PAH - , limiting the accessibility of product channel (19) for which an 'edge-on' attack of PAH - by CN appears most favourable. Experimental evidence for the inefficiency of hydrogen atom abstraction reactions of polyatomic, aromatic ions with radicals has been reported recently, in a selected-ion flow tube (SIFT) study involving the reactions of C6H5+ ,C6~+ and C6Hi with atomic hydrogen: for each of these ions, hydrogen atom abstraction is the sole exothermic bimolecular product channel C6H,! +H-+C6H'!_1 +Hz (21) but does not occur detectably: kZI < 1 X 10- 11 em3 molecule-I S-I at 300 K (Petrie, Javahery & Bohme 1992). Associative detachment (channel 20). The thermochemistry of reaction (20) is known only for the reaction of the naphthalenide ion CloH 7- , an even-electron PAH - ion. Associative detachment involving CN is exothermic in this instance by -5.4 eV (Lias et al. 1988): this is much more exothermic than the competing charge-transfer channel (16) is expected to be, suggesting that associative detachment should dominate in this instance. Since associative detachment of CN to other even-electron PAH- ions is likely to be similarly exothermic, we expect that reaction (20) should dominate over reaction (16) for all even-electron PAH - ions. The situation regarding associative detachment to odd-electron PAH - is less certain: while almost certainly exothermic, it involves formation of a neutral radical and hence is likely to be less exothermic than channel (20) involving even-electron PAH-. Conversely, product channel (16) is almost certainly more exothermic for most odd-electron PAH - than for even-electron PAH - , and so we anticipate that channel (16) is more likely to compete with channel (20) in the reactions of odd-electron PAH- with CN. Experimental investigation of this issue would be very useful. For odd-electron PAH- ions, one further factor likely to affect the efficiency of the possible product channels (16), (18) or (19) is that of electron spin conservation. If spin is conserved in these reactions, and if it is assumed that reactions (16), (18) and (19) are exothermic for generation of ground-state products but endothermic for all product channels involving excited electronic states, then a selection factor of less than unity must be imposed on these product channels upon the grounds of spin conservation. It should be noted, however, that positive ion/radical reactions are often not constrained by spin conservation (Ferguson 1983; Federer et al. 1986), and the same may well be the case in the reactions of negative ions. It should also be borne in mind that associative detachment (20) of CN will always conserve spin, and so should dominate for those spin states which correspond to excited states of the products of channels (16), (18) and (19). In the model which we present below, we shall consider that charge-transfer reactions (16) conserve spin. However, since the thermochemistry of even-electron PAH- ions is not well established [and since associative detachment (20) is likely to dominate in the reactions of these ions with CN], we shall limit our model to the inclusion only of odd-electron PAH - ions. 2.4 Proton transfer from H(eN) By analogy with the consideration of charge transfer from PAH - to CN, it is of interest also to consider the prospects for proton transfer from HCN or HNC to PAH-: CnH';;- + HCN-+CnHm+1+CN- (22) CnH,;;- + HNC-+CnHm+1+ CN-. (23) Proton transfer is generally rapid when exothermic (Bartmess & McIver 1979). The exothermicity of reaction (22) or (23) depends upon the relative gas-phase acidities of CnHm+1 and H(CN): -MlZZ=Mlacid(CnHm+I)Mlacid(HCN), and -Ml23=Mlacid(CnHm+I)MlaciiHNC), where Mlacid(HCN) = 1438 kJ mol-I (Lias et al. 1988) and Mlacid(HNC) -1377 kJ mol-I. Few thermochemical data exist on the gas-phase acidities of PAHs: measurements on derivatized PAHs (Lias et al. 1988) suggest that reaction (22) may be exothermic only for small PAH- ions, while reaction (23) may be exothermic for much larger PAH - ions. Given the lack of thermochemical data on such processes, we will not include reactions (22) or (23) in our model of CN- chemistry: however, consideration of such processes may be worthwhile if gas-phase acidities of representative PAHs are determined. © 1996 RAS, MNRAS 281, 137-144 © Royal Astronomical Society • Provided by the NASA Astrophysics Data System 1996MNRAS.281..137P CN- within interstellar clouds 2.S Dissociative attachment to ZCN 141 Dissociative attachment (DA) to CN-containing species micity of DA is not an absolute guarantee of its efficiency. For example, the DA reactions ZCN+e--+CN-+Z Clz + e --+ CI + CI- , (27) CH3Br + e --+ CH3+ Br- (28) (24) has not been considered in previous models of IS chemistry. Reaction (24) is exothermic if the bond strength D(Z-CN) is less than 3.82 eV, the electron affinity of CN. In Table 2, we list the bond strengths of known interstellar or circumstellar CN-containing species. The table shows that covalent bonds from CN to C, H or N are too strong for DA to such species to be exothermic, and so none of the dense-cloud species in Table 2 can act as sources of CN -. Of the CS metal cyanides, which have recently been detected within IRC + 10216 (Kawaguchi et al. 1993; Turner, Steimle & Meerts 1994; Ziurys et al. 1995), the Na-CN bond strength is also too high to permit DA: however, the comparatively low bond strength calculated (Gardner et al. 1993; Petrie 1996a) for both isomers of Mg(CN) indicates that DA is exothermic for MgCN and for MgNC: MgCN +e--+Mg+CN-, (25) MgNC + e --+ Mg + CN- . (26) Given the comparatively high abundance of MgNC within IRC + 10216 (Kawaguchi et al. 1993), these processes may be of considerable importance as potential precursors to CN-. It is necessary to estimate the efficiency of reactions (25) and (26). Very few studies of DA at subthermal energies have been reported (Dunning 1995). Room-temperature studies, involving low-energy DA to haloalkanes (Underwood-Lemons, Gergel & Moore 1995), to HI (Smith & Adams 1987), to XeF2 (Sides & Tieman 1976), and to various 'strong acids' such as HN0 3 and H 2S04 (Adams et al. 1986), suggest that exothermic DA often approaches collisional efficiency at low energies. However, the exother- Table 2. Bond strengths D (Z-CN) for interstellar and circumstellar CN-containing species. are exothermic by 1.1 and 0.3 e V respectively (Lias et al. 1988), yet are inhibited by activation energy barriers (Christodoulides, Schumacher & Schindler 1975; Datskos, Christophorou & Carter 1992; Underwood-Lemons et al. 1995) which would render them very inefficient at low temperatures. The efficiency of DA involving MgCN or MgNC can only be assessed through experimental investigation: in the absence of such a study, we shall assume here the most 'optimistic' case, in which reactions (25) and (26) are unimpeded by activation barriers. An additional consideration which has not previously been addressed in studies of low-energy DA is the influence of spin conservation. The reaction of doublet Mg(CN) (either isomer) with an electron will occur upon the triplet potential energy surface on 75 per cent of collisions, and upon the singlet surface the remaining 25 per cent. To investigate the thermochemistry of the [Mg(CN)-)* activated complexes, we have performed Gaussian-2 (G2) ab initio calculations according to the method of Curtiss et al. (1991) upon the singlet and triplet negative ions formed by electron attachment to MgCN and MgNC. The results of these calculations are summarized in Table 3. Comparison of the results for the lowest-lying singlet and triplet states of Mg(CN) - indicates that formation of triplet Mg( CN) - from Mg(CN) + e is endothermic for both isomers, and thus presumably involves approach upon a repulsive surface. While such a surface is not necessarily an impediment to DA, formation of triplet (Mg + CN-) from Mg(CN) + e is also substantially endothermic, so that DA upon the triplet surface is possible only upon spin flipping. In contrast, DA upon the singlet surface is exothermic and fully spin-allowed. As mentioned above, we assume here that exothermic DA to MgCN and MgNC (i.e., upon the singlet surface) does not involve any barriers, and hence is efficient at low temperatures. Reference 2.6 Loss processes for CN- HCN 5.36 eV Lias et al. 1988 b HNC 4.68 eV Lias et al. 1988 b In keeping with previous treatments of negative-ion IS chemistry (Dalgarno & McCray 1973; Herbst 1981; Millar et al. 1991), we consider the two principal loss processes for CN- to be the associative detachment reaction (8) and mutual neutralization (11). Experimental studies suggest that both of these classes of reaction are highly efficient (Howard, Fehsenfeld & McFarland 1974; Smith, Church & Miller 1978), with typical values (at 300 K) of k8 - 1 X 10- 9 cm3 molecule-I S-I and kl1 -(4-10) x 10- 8 cm3 molecule-I S-I. Mutual neutralization reaction rate coefficients are expected, on theoretical grounds (Hickman 1979), to possess a temperature dependence of _T- O.5, so that at IS cloud temperatures of - 10 K, values of kl1 - (2-6) X 10- 7 em3molecule -I s -I are anticipated. If n (H) _ 5 x 10-5 n (Hz) and n(M+)-3 x 10- 8 n(Hz) within a 'typical' dense IS cloud, it can be inferred that reactions (8) and (11) will be broadly comparable in importance as loss processes for CN-. ZCN D(Z-CN) " - 6.0eV CHaCN Lias et al. 1988 b CH3CN 5.23 eV Lias et al. 1988 b CCCN 5.23 eV Francisco & Richardson 1994 c HCCCN 6.37 eV Francisco & Richardson 1994 c CH2CHCN 5.34 eV Lias et al. 1988 b H2NCN 5.34 eV Lias et al. 1988 b Na(CN) 4.50eV Petrie 1996b c MgCN 3.26 eV Gardner et al. 1993 c MgNC 3.33 eV Petrie 1996a c Notes. "Bond dissociation energy Dg in eV. bExperimental value. <Theoretical value, obtained from the G2 ab initio method; estimated uncertainty of ±O.1 eV. © 1996 RAS, MNRAS 281,137-144 © Royal Astronomical Society • Provided by the NASA Astrophysics Data System 1996MNRAS.281..137P 142 S. Petrie Table 3. Geometries and total energies determined for Mg(CN)-. Geometries: Species' j' r(Mg-C)b IMgNCIMgCN" 2.285 3MgNC3MgCN" 2.174 Eo (Hartrees) C IMgNC- Bb products kj C f(A) d n(B) • r(Mg-N)b 10 e CN CN- +hv 1(-14) 0.9 3(-8) 1.196 2.126 16 PAir CN PAH + CN'" 5(-10) 0.031 3(-8) 1.191 25 e MgCN Mg+CN'" 2.5(-8) 0.9 0; 7(-10)g,/l 1.194 26 e MgNC Mg+CN'" 2.5(-8) 0.9 0; 1.5(-8) g,i 8 CN'" H H(CN) +e 1(-9) 7 5(-5) 11 M+ CN'" M+CN 5(-7) 1 2.051 1.179 Erel d tlll'lf,o e -292.41879 -117.8 90.9 IMgCN" -292.41369 -104.4 104.3 Mg + eN" -292.373931 0.0 MgNC + e -292.35009 h 62.6 271.3 MgCN + e -292.34777 i 68.7 277.4 3MgCN" -292.33079 113.2 321.9 -292.33057 113.8 322.5 3MgNC- Ab r(C-N)b Energies: Species Table 4. Mechanism of CN - formation and destruction in IS clouds and CS envelopes. 208.7g Notes. "All the structures listed are linear (LMgXY = 180°). bBond length in A, obtained from the optimized geometry at the MP2(full)/6-31G* level of theory. CTotal energy (including ZPE) in hartree, obtained according to the G2 procedure of Curtiss et al. (1991). dG2 energy (in kJ mol-I) expressed relative to Mg + CN-. 'G2 enthalpy of formation (zero-K value) of the molecular species indicated. 'From Curtiss et al. (1991). gExperimental value: LlliJ.o(Mg)+MIJ(CW)=220.5 kJ mol- 1 (Lias et al. 1988). hFrom Petrie (1996a). 'From Gardner et al. (1993). 2.7 A general model of eN - chemistry The reactions discussed in Sections 2.1-2.6 can be incorporated in a model of cyanide-ion chemistry, which is detailed in Table 4. This model does not include those reactions (4)-(7) which feature in the UMIST model (Millar et al. 1991): these reactions are excluded here because they appear not to lead to significant abundances of CN- and because the abundances of the reactant ions 0and OH- are arguably very uncertain. In contrast, the radiative attachment reaction (10) has been included despite its believed unimportance as a pathway to CN- because the abundances of the reactants involved are well established in various models of cloud chemistry. It should be recalled, from Section 2.2, that reaction (10) is of comparable efficiency to the UMIST reaction set (4)-(7) as a pathway to CN- within IS clouds. Some comments on the abundances and rate coefficients chosen for Table 4 are required. We assume n(I-)=3 x 10- 8 n(H2) as a total abundance of negativecharge-carrying species 1-, and abundances of free electrons, PAH - and M + are expressed relative to this. The odd-electron PAH- abundance of 0.03 n(I-) reflects the recent indications that PAH- formation may be comparatively inefficient: free electrons are therefore regarded as the predominant negative charge carriers. The rate coeffi- Notes. "Reaction number, as identified in text. bReactants. CEstimated rate coefficient in units of cm3 molecule- 1 S-I. dEstimated relative abundance of reactant A, expressed as a fraction f of the total abundance n (I -) of negative-charge-carrying species. 'Estimated abundance of reactant B, expressed relative to n (H2) = 1. The notation a( -b) represents a x 1O- b• 'Estimated fractional abundance of odd-electron PAH - ions. gAbundance within dark clouds (e.g., TMC-1) and within circumstellar envelopes (e.g., IRC + 10216) respectively. hCircumstellar abundance taken from Ziurys et al. (1995). 'Circumstellar abundance taken from Kawaguchi et al. (1993). cients for reactions (16), (25) and (26) assume that only 1/4 of such collisions lead to product formation upon the singlet potential energy surface: CN is weakly polar, while MgCN and MgNC are calculated to have large dipole moments and so their collision rate coefficients with e are expected to be high (kc -1O- 7 cm3 molecule-I S-I) at IS or CS temperatures. Finally, we assume for the sake of simplicity that, excepting MgCN and MgNC, the abundances of reactants within CS envelopes are identical to those within IS clouds. A pseudo-steady-state expression for the CN- abundance is obtained from this model: n(CN-)= j(e)[klo1l(CN) +k25n(MgCN) +k26n(MgNC)] +k 1d(PAH-)n(CN) where all parameters are as defined in Table 4. Two sets of conditions can be distinguished. In the first, which we call the 'IS model', MgCN and MgNC are absent, and so CNarises only via reactions (10) and (16). This appears to mirror the situation in a dense cloud such as TMC-1, in which no magnesium- (or other metal) containing species have been detected: in this scenario, the steady-state abundance found is n(CN-)=2 x 10- 13 n(H2)' This abundance is approximately two orders of magnitude greater than that expected if CN- arises only via reaction (10) and/or via the UMIST reaction set (4)-(7). Clearly, the CN- formed via the IS model arises almost exclusively by charge transfer from PAH- ions. However, the calculated CN- abundance is still almost certainly too low to permit its radioastronomical detection within IS clouds. The second scenario, which we term the 'CS model', includes MgCN and MgNC at their observed abund© 1996 RAS, MNRAS 281,137-144 © Royal Astronomical Society • Provided by the NASA Astrophysics Data System 1996MNRAS.281..137P CN- within interstellar clouds ances within IRC + 10216. In this model, n(CN-)=1.7 x 10- 10 n(Hz) is achieved if it is assumed that DA occurs with spin conservation [Le., only 1/4 of Mg(CN) + e collisions lead to DA upon the singlet surface]. Even with this restriction, dissociative attachment to Mg(CN) is clearly the dominant pathway to CN-, and the calculated abundance should permit detection of CNwithin a source such as IRC + 10216. A higher abundance of CN- [up to ",7 X 10- 10 n(H2 )] would be expected if the DA reactions (25) and (26) occur without spin conservation, which is feasible given the tendency for violation of spin conservation in several classes of ion/molecule reactions (Ferguson 1983; Federer et al. 1986). Radioastronomical detection of CN - would require determination of the spectroscopic constants of this ion to high accuracy. High-level ab initio calculations on CN - have been performed by Botschwina (1985) and by Peterson & Woods (1989), but as yet no experimental study has supplemented the theoretical results. Such a study would be of great benefit in investigating the prospects for production of CN-, and of other small negative ions, within IS clouds and CS envelopes. 2.8 Implications for IS and CS chemistry The prospects for chemical evolution via the reactions of CN- are poor, since this species is expected to be depleted rapidly by reactions (8 and 11) which do not lead to a significant increase in molecular complexity. This does not mitigate the importance of CN- as a possible 'indicator' species: it should be noted that, even though negatively charged species (including free electrons) must be present within dense IS clouds and CS envelopes, no such species have yet been detected. The CN - ion constitutes a (potentially) readily identifiable negative-charge carrier, and its detection, perhaps within a source such as IRC + 10216, would serve a useful role in expanding our understanding of IS or CS chemical processes. A full understanding of the processes of CN- formation and removal would permit an estimation of free electron abundances (and, hence, of the degree of ionization) within various sources based on their observed CN- abundances. Some comments on the formation of other small negatively charged species, by reactions analogous to (16), (25) and (26), are in order. The halogens F, CI and Br all have large ( > 3.3 eV) electron affinities, and so charge transfer from PAH - is a viable route to F -, Cl- and Be, and is likely to predominate over electron attachment by analogy with the factors which have been discussed for CN- production. Other species such as H, C, 0, Si and S have lower electron affinities (0.7 <EA <2.1 eV) but are expected to be present in IS clouds at substantially higher abundances than the halogen atoms; charge transfer from some PAHions may also represent a route to H-, C-, etc. However, none of the atomic negative ions which might arise by these pathways are easily detectable within IS clouds, since they lack rotational spectra. Of diatomic species, OH (EA = 1.828 eV; Miller 1991) is the most abundant IS species having a significant electron affinity: however, as with CN, charge transfer from PAH- to OH is probably insufficient to produce OH- in detectable quantities. Identification of a larger species having a very weak Z-OH bond might allow the formation of OH- by dissociative 143 attachment, but no such species has yet been observed. We conclude, therefore, that CN - remains the most favourable diatomic negative ion for IS or CS detection. 3 CONCLUSIONS Radiative electron attachment has traditionally been regarded as the principal route to negative ions within IS and CS environments. We have suggested here that two chemical pathways - namely, charge transfer from PAHions to CN, and dissociative electron attachment to Mg(CN) - offer substantially more effective pathways to CN- than does radiative attachment. Simple model calculations indicate that, if dissociative attachment to Mg(CN) is efficient, this process can lead to detectable CN- abundances within CS envelopes such as IRC + 10216. The significance of CN - as a constituent of IS clouds or CS envelopes is more as an 'indicator' of physical or chemical conditions - for example, the degree of ionization - than as a pathway to other IS species. Neutralization of CN- is expected to yield principally CN, HCN and (possibly) HNC, all of which are formed by several other pathways also. In any event, the detection of CN- will rely on the determination of the rotational spectrum for this species. ACKNOWLEDGMENTS The author thanks Professor David Smith for supplying preprints, and Professor Eric Magnusson for helpful advice. REFERENCES Adams N. G., Smith D., Viggiano A A, Paulson J. F., Henchman M. J., 1986, J. Chern. Phys., 84, 6728 Bartmess J. E., Mciver R. T., 1979, in Bowers M. T., ed., Gas Phase Ion Chemistry, Vol. 2. Academic Press, New York, p. 87 Becker R. S., Chen E., 1966, J. Chern. Phys., 45, 2403 Berkowitz J., Chupka W. A, Walter T. A, 1969, J. Chern Phys., 50, 1497 Botschwina P., 1985, Chern Phys. Lett., 114,58 Canosa A, Parent D. c., Pasquerault D., Gornet J. c., Laube S., Rowe B. R., 1994, Chern. Phys. Lett., 228, 26 Chaizy P. et ai., 1991, Nat, 349, 393 Christodoulides A A, Schumacher R., Schindler R. N., 1975, J. Phys. Chern., 1904 Cossart-Magos c., Leach S., 1990, A&A, 233, 559 Crocker L., Wang T., Kebarle P., 1993, J. Am. Chern. Soc., 115, 7818 Curtiss L. A, Raghavachari K., Trucks G. W., Pople J. A, 1991, J. Chern. Phys., 94, 7221 Dalgarno A, McCray R. A, 1973, ApJ, 181, 95 Datskos P. G., Christophorou L. G., Carter J. G., 1992, J. Chern. Phys., 97, 9031 Dewar M. J. S., Hashrnall J. A, Venier C. G., 1968, J. Am. Chern. Soc., 90, 1953 Dunbar R. c., 1990, Int. J. Mass Spectrorn. Ion Processes, 100, 423 Dunning F. B., 1995, J. Phys. B: At. Mol. Opt. Phys., 28,1645 Federer W., Villinger H., Lindinger W., Ferguson E. E., 1986, Chern. Phys. Lett., 123, 12 Ferguson E. E., 1983, Chern. Phys. Lett., 99, 89 Foing B. H., Ehrenfreund P., 1994, Nat, 369, 296 Fram;isco J. S., Richardson S. L., 1994, J. Chern. Phys., 101,7707 © 1996 RAS, MNRAS 281,137-144 © Royal Astronomical Society • Provided by the NASA Astrophysics Data System 1996MNRAS.281..137P 144 S. Petrie Gardner P. J., Preston S. R., Siertsma R., Steele D., 1993, J. Comput. Chem., 14, 1523 Hare J. P., Kroto H. W., 1992, Ace. Chem. Res., 25, 106 Herbst E., 1981, Nat, 289, 656 Herbst E., Dunbar R. c., 1991, MNRAS, 253, 341 Hickman A P., 1979, J. Chem. Phys., 70, 4872 Howard C. J., Fehsenfeld F. c., McFarland M., 1974, J. Chem. Phys., 60, 5086 Irvine W. M., Goldsmith P. F., Hjalmarson A., 1987, in Hollenbach D. J., Thronson H. A, Jr., eds, Interstellar Processes. Reidel, Dordrecht, p. 561 Jaffke T., Illenberger E., Lezius M., Matejcik S., Smith D., Mark T. D., 1994, Chem Phys. Lett., 226, 213 Kawaguchi K., Kagi E., Hirano T., Takano S., Saito S., 1993, ApJ, 406, L39 Kebarle P., Chowdhury S., 1987, Chem. Rev., 87, 513 Leger A, d'Hendecourt L., 1985, A&A, 146, 81 Lepp S., Dalgarno A, 1988, ApJ, 324, 553 Lias S. G., Bartmess J. E., Liebman J. F., Holmes J. L., Levin R. D., Mallard W. G., 1988, J. Phys. Chem. Ref. Data, 17, Suppl. No. 1 Millar T. J., 1992, MNRAS, 259, 35P Millar T. J., Rawlings J. M. c., Bennett A, Brown P. D., Charnley S. B., 1991, A&AS, 87, 585 Miller T. M., 1991, in Lide D. R., ed., CRC Handbook of Chemistry & Physics, 72nd ed. CRC Press, Boca Raton, p. 10 Omont A, 1986, A&A, 164, 159 Peterson K. A, Woods R. c., 1989, J. Chem. Phys., 90,7239 Petrie S., 1996a, J. Chem. Soc., Far. Trans., in press Petrie S., 1996b, J. Phys. Chem., submitted Petrie S., Javahery G., Bohme D. K., 1992, J. Am. Chem. Soc., 114, 9205 Petrie S., Javahery G., Bohme D. K., 1993, A&A, 271, 662 Sarre P. J., 1980, J. Chim. Phys., 77, 769 Sides G. D., Tiernan T. 0., 1976, J. Chem. Phys., 65, 3392 Smith D., Adams N. G., 1987, J. Phys. B: At. Mol. Phys., 20, 4903 Smith D., Church M. J., Miller T. M., 1978, J. Chem. Phys., 68, 1224 Smith D., Spanel P., Mark T. D., 1993, Chem. Phys. Lett., 213, 202 Spanel P., Smith D., 1994, Chem. Phys. Lett., 229, 262 Sunderlin L. S., Paulino J. A, Chow J., Kahr B., Ben-Amotz D., Squires R. R., 1991, J. Am. Chem. Soc., 113, 5489 Turner B. E., Steimle T. C., Meerts L., 1994, ApJ, 426, L97 Underwood-Lemons T., Gergel T. J., Moore J. H., 1995, J. Chem. Phys., 102, 119 van der Zwet G. P., Allamandola L. J., 1985, A&A, 146,76 Webster AS., 1992, A&A, 257, 750 Webster A, 1993, MNRAS, 264, L1 Yang S. H., Pettiette C. L., Conceicao J., Chesnovsky 0., Smalley R. E., 1987, Chem. Phys. Lett., 139,233 Ziurys L. M., Apponi A J., Guelin M., Cernicharo J., 1995, ApJ, 445, L47 © 1996 RAS, MNRAS 281,137-144 © Royal Astronomical Society • Provided by the NASA Astrophysics Data System
© Copyright 2025 Paperzz