This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.

Articles

CRITICAL REVIEW OF N, N+, N+2, N++, And N++2 MAIN PRODUCTION PROCESSES AND REACTIONS OF RELEVANCE TO TITAN'S ATMOSPHERE

, , , , , , , , , , , , , , , , , and

Published 2013 January 29 © 2013. The American Astronomical Society. All rights reserved.
, , Citation Odile Dutuit et al 2013 ApJS 204 20 DOI 10.1088/0067-0049/204/2/20

0067-0049/204/2/20

ABSTRACT

This paper is a detailed critical review of the production processes and reactions of N, N+, N+2, N++, and N++2 of relevance to Titan's atmosphere. The review includes neutral, ion–molecule, and recombination reactions. The review covers all possible active nitrogen species under Titan's atmospheric conditions, specifically N2 (A3Σ+u), N (4S), N (2D), N (2P), N+2, N+ (3P), N+ (1D), N++2, and N++ species, and includes a critical survey of the reactions of N, N+, N+2, N++, and N++2 with N2, H2, D2, CH4, C2H2, C2H4, C2H6, C3H8 and the deuterated hydrocarbon analogs, as well as the recombination reactions of N+2, N+, N++2, and N++. Production processes, lifetimes, and quenching by collisions with N2 of all reactant species are reviewed. The N (4S) state is reactive with radicals and its reactions with CH2, CH3, C2H3, and C2H5 are reviewed. Metastable states N2 (A3Σ+u), N (2D), and N (2P) are either reactive or quenched by collisions with the target molecules reviewed. The reactions of N+ (1D) have similar rate constants as N+ (3P), but the product branching ratios differ significantly. Temperature effects and the role of the kinetic energy content of reactants are investigated. In all cases, experimental uncertainties of laboratory data are reported or estimated. Recommended values with uncertainties, or estimated values when no data are available, are given for rate constants and product branching ratios at 300 K and at the atmospheric temperature range of Titan (150–200 K for neutral reactions and 150 K for ion reactions).

Export citation and abstract BibTeX RIS

1. INTRODUCTION

1.1. Chemistry in Titan's Atmosphere

The atmosphere of Titan is mainly composed of molecular nitrogen, with 2%–5% methane (Niemann et al. 2010; Yelle et al. 2008) and small amounts of hydrocarbons such as acetylene, ethylene, ethane, and propane, as well as traces of nitrogen and oxygen-bearing species (Cui et al. 2009; Koskinen et al. 2011; Vinatier et al. 2010; Vuitton et al. 2007). The heaviest neutral molecule that has unambiguously been detected to date in Titan's atmosphere is benzene (Vinatier et al. 2010; Vuitton et al. 2008). In addition, Titan's atmosphere also contains a large number of aerosol particles, suggested to be composed of organic material, which form a thick haze layer (Tomasko et al. 2008) as well as a detached thin layer in the stratosphere (Lavvas et al. 2009). The Cassini space probe, which in 2005 performed the first ever in situ measurements of Titan's atmosphere, revealed an extraordinarily complex ionospheric chemical composition. The Ion and Neutral Mass Spectrometer (INMS) detected about 50 positive ions in the 1–99 m/z mass range (Cravens et al. 2006) and the Cassini Plasma Spectrometer–Ion Beam Spectrometer (CAPS-IBS) provided evidence for even heavier positive ions with m/z up to ∼350 (Crary et al. 2009). Another striking result came from another sensor of the CAPS instrument, namely the Electron Spectrometer, which reported the presence of negatively charged ions with m/z up to ∼4000 (Coates et al. 2007). These heavy ions could be intermediate species leading to the formation of the aerosols which have been observed at altitudes as high as 900 km by the Ultraviolet Imaging Spectrograph (Koskinen et al. 2011). In Titan's atmosphere, the temperature varies from 93 K at the surface to about 250 K at the top of the ionosphere at ∼1600 km (Fulchignoni et al. 2005). In the atmosphere at altitudes between 250 km and 900 km where chemistry is dominated by neutral reactions, the HASI instrument on board Huygens registered wave-like variations in the temperature profile in the 150–200 K range, with a mean value of ∼170 K (Fulchignoni et al. 2005). In the ionosphere, at altitudes between 900 and 1600 km, where chemistry is dominated by ion–molecule reactions, the ion (Crary et al. 2009) and neutral (De la Haye et al. 2007) temperatures are not exactly equal and vary in the ranges 100–200 K and 145–160 K, respectively. Thus, one can conclude that the characteristic average temperature for ion–molecule reactions is about 150 K. In Titan's atmosphere, the electron temperature ranges from 150 to 10,000 K (Edberg et al. 2010), lying between 150 K and 1000 K range in the ionosphere (Richard et al. 2011), where ion–electron recombination plays a major role.

The first results of the Cassini–Huygens mission triggered a tremendous modeling effort to interpret the probe's data. This modeling effort used the most complex chemical models employed to date to represent a planetary atmosphere (Cravens et al. 2009; Krasnopolsky 2009; Lavvas et al. 2008a, 2008b; Robertson et al. 2009; Vuitton et al. 2009; Westlake et al. 2012). N2 neutral molecules in their ground electronic state are non-reactive and therefore, the first steps of molecular growth are mainly due to chemically active nitrogen species. These active nitrogen species come from the dissociation and (dissociative) ionization of N2 by ultraviolet photons and supra-thermal electrons in the upper atmosphere (Lavvas et al. 2011). An extensive recent review on the chemistry of Titan's atmosphere can be found in the book on Titan to be published by Cambridge University Press (Vuitton et al. 2012).

1.2. Aim and Outline of the Paper

The aim of the present paper is to present a critical review of the reactions of all the possible chemically active nitrogen species playing an important role in Titan's atmospheric chemistry; specifically the reactions of N2 (A3Σ+u), N (4S), N (2D), N (2P), N+2, N+ (3P), N+ (1D), N++2, and N++ with the main constituents of Titan's atmosphere, N2, H2, CH4, as well as several minor hydrocarbons, i.e., C2H2, C2H4, C2H6, and C3H8. In the literature, there are several review papers that study these chemical reactions. These papers are currently widely used by planetary scientists to build their chemical atmospheric models. For neutral reactions, Herron (1999) published a complete critical evaluation of the reactions of N (2D), N (2P), and N2 (A3Σ+u) with many different molecules. For reactions between singly charged ions and neutrals, Anicich & McEwan (Anicich 2003, 1993; McEwan & Anicich 2007; McEwan et al. 1998) published several review papers considering all ion–molecule reactions, at room temperature, which can play a role in Titan's atmospheric chemistry. Electron–ion recombination reactions (Adams et al. 2006; Florescu-Mitchell & Mitchell 2006; Johnsen 2011; Thomas 2008; Vigren et al. 2012) were considered in several recent review papers. For doubly charged ion reactions, Thissen et al. (2011) published a recent review which gives a more general overview of this class of reactions and their possible role in planetary atmospheres. All these earlier compilations are incomplete and more recent work is now available. The current paper provides the necessary update to these earlier reports and also includes the urgently required critical evaluation of the available data.

Due to the lack of relevant experimental studies, predominantly reaction rates for ground state species at room temperature have been used in atmospheric models, or at best Arrhenius-based extrapolations of these rates to temperatures of interest (Hébrard et al. 2009). In this review, we specifically investigate the effects of temperature, as well as internal and kinetic energy of the reactant species, on the observed reactivity. We also emphazise effects which can play a significant role in influencing the reactivity to be expected under Titan's atmospheric conditions and make remarks on factors which do not influence the reactivity, within experimental uncertainties. Let us note that it has been shown experimentally and theoretically in the last two decades that many neutral–neutral reactions have no barrier on the minimum energy path leading from reactants to products. Therefore, especially in the low temperature range of Titan's atmosphere (150–200 K), the temperature dependence of their rate constants can follow only in some cases the simple Arrhenius law:

where A is the pre-exponential factor, Ea is the activation energy, and kb is the Boltzmann constant. See the paper of Smith (2008) for a discussion about the factors that control the rates of "non-Arrhenius" reactions.

Temperature has little effect on the values of the rate constants of ion–molecule reactions, as most reactions are already fast. Such reactions usually proceed at close to the collision rate, which is given by the Langevin expression for exothermic reactions assuming reaction after reactant capture:

where e is the elementary charge, α is the polarizability of the neutral molecule, and μ is the reduced mass of the ion neutral pair. The Langevin rate is independent of temperature and is the maximum value for the rate constant when the target molecule has no permanent dipole, as is the case for all reactions considered in this paper. When the rate constant at 300 K is smaller than the Langevin rate, it may increase when the temperature decreases. This increase at lower temperatures can reasonably be neglected in a first approximation for most exothermic reactions (Wakelam et al. 2010). However, the effect of temperature on the product branching ratios of ion–molecule reactions is much more significant. Surprisingly, Carrasco et al. (2008a) showed, in a study of uncertainty propagation in a Titan's ionospheric model, that large uncertainties on branching ratios affect the calculated density of only a few major ions, but they strongly affect many minor ions. Considering the lack of experimental data on such temperature effects, uncertainty management should be an intrinsic component of atmospheric chemistry modeling.

For some applications involving modeling the upper atmosphere of a planet (study of the atmospheric escape for example), it is important to know how reaction cross sections vary with collision energy. Such considerations are necessary because reactant ions can possess significant kinetic energies. For exoergic ion–molecule reactions, the reaction cross sections σ have been found to decrease with collision energy, roughly as

where A is a constant and ECM is the energy in the center of mass frame. This expression has been used in this paper to make estimates of the rate constant at 150 K, when no experimental data were available.

Each section of this paper focuses on a particular nitrogen chemical species: Section 2 on N2 (A3u+), Section 3 on N (4S) ground state as well as N (2D) and N (2P) metastable atoms, Section 4 on N+2 ions, Section 5 on N+ (3P) ground state and N+ (1D) metastable atomic ions, Section 6 on N++2 molecular doubly charged ions, and Section 7 on N++ atomic doubly charged ions. For each chemical species, we first describe its production from N2 by both ultraviolet photons (50–3000 Å) and the impact of supra-thermal electrons (15–1000 eV). We then move on to considering the lifetime of the chemical species and any relevant quenching reactions, before considering its chemical reactions including ion–electron recombination reactions. We also provide an estimate of the percentage of N and N+ species that reside in atomic metastable states, given the production of these species by photodissociation, N2 dissociation by electron impact, dissociative ionization of N2, and electron–ion recombination reactions. Best estimates of rate constants and product branching ratios, with uncertainties, are given for all reactions at 300 K. We also give, when possible, the temperature dependence of the rate constant for neutral reactions and an estimate of the rate constants of ion–molecule reactions at 150 K, which is the representative of Titan's ionospheric temperature.

This review work has been initiated as the result of an interdisciplinary effort within the European Network Europlanet RI (Research Infrastructure), focusing on employing the expertise of physical chemists to improve the first steps of the chemistry in the models of Titan's atmosphere developed by planetary scientists. The authors are experts in the different topics reviewed in this paper (see Annex). Let us note that the present critical review can also be useful for other planetary atmospheres or even other applications outside planetary sciences. Our work on this review is linked to the effort to build the KIDA (KInetic Database for Astrochemistry) database of neutral and ion reactions, which includes an evaluation of the data from the literature for applications in astrophysics and planetary sciences (Wakelam et al. 2012).

1.3. Active Nitrogen

Dissociation and ionization thresholds of N2, as well as the lifetime of the excited states are gathered in Table 1. Figure 1 shows schematic potential energy curves of N2, N+2, and N++2, with the main dissociation channels, which are relevant for the present paper.

Figure 1.

Figure 1. Schematic potential energy curves of N2, N+2, and N++2. Adapted from references (Franceschi et al. 2007; Gilmore 1965; Hiyama & Iwata 1993; Lewis et al. 2005; Nicolas et al. 2003a).

Standard image High-resolution image

Table 1. Dissociation and Ionization Thresholds of N2, as well as Lifetimes of Excited States

Species Threshold Energy Remarks
(eV)
N2 (X1Σ+g) 0  
N2 (A3Σ+u) 6.22 N2 (A) lifetime = 2.37 s (a)
N (4S) + N (4S) 9.76 Not observed
N (2D) + N (4S) 12.14 N (2D) lifetime = 13.6 hr and 36.7 hr * (b)
N (2P) + N (4S) 13.33 N (2P) lifetime = 11.1 s and 10.5 s ** (b)
N (2D) + N (2D) 14.52  
N+2 (X2Σ+g) 15.58  
N+2 (A2Πu) 16.93 N+2 (A) lifetime = 13.9 – 7.3 ms *** (c)
N+2 (B2Σ+u) 18.75 N+2 (B) lifetime = 67 ns (d)
N+ (3P) + N (4S) 24.29  
N+ (1D) + N (4S) 26.19 N+ (1D) lifetime = 258 s (e)
N+ (3P) + N (2D) 26.68  
N+ (3P) + N (2P) 27.87 Not observed
N+ (1S) + N (4S) 28.35 Not observed
N++2 (X1Σ+g) 42.88 N++2 lifetime = 3 s (f)
N+ (3P) + N+ (3P) 44.5  
N++ (2P) + N (4S) 53.9 Appearance energy is 55.2 eV (g)

Notes. *Lifetimes for the N (2D3/2) and N (2D5/2) sub-states, respectively. **Lifetimes for the N (2P1/2) and N (2P3/2) sub-states, respectively. ***The lifetime decreases with increasing vibrational level. References. (a) Piper 1993; (b) Ralchenko et al. 2011; (c) Peterson & Moseley 1973; (d) Wuerker et al. (1988); (e) Wiese & Fuhr 2007; (f) Mathur et al. 1995; (g) Franceschi et al. 2007.

Download table as:  ASCIITypeset image

N2 is characterized by a strong triple bond with associated dissociation energies of 9.76 eV, 8.71 eV, and 1.63 eV, for neutral N2, N+2, and N++2 ions, respectively. As a consequence, chemical reactions involving an N–N bond cleavage are unlikely for N2 and N+2 reactions. Such reactions are only possible for reactions of doubly charged N++2 ions. Such molecular dications are much less stable and most probably play a minor role in Titan's atmosphere due to their low density (Lilensten et al. 2005). Consequently chemical reactions in Titan's atmosphere are mainly induced by N and N+ fragment ions. However, N+2 reactants can generate very reactive species via dissociative charge transfer or by capture of a hydrogen atom from hydrocarbons forming N2H+ ions. Reactions of all nitrogen active species with N2 are mainly quenching reactions, whereas reactions with hydrocarbon molecules initiate the complex chemistry of Titan's atmosphere. Among the electronic excited species, only the long-lived N2 (A3π+u), N (2D), N (2P), and N+ (1D) excited species have a lifetime much longer than the time between collisions (∼1 s at 1000 km). However, except for N (2D) metastable atoms, the reactivity of these species have not been included in photochemical models as yet. The very reactive N++2 and N++ doubly charged ions have also not yet been introduced in models of Titan's chemistry, due to a lack of information, both on the double ionization cross sections for dication production and on the chemical reaction rate constants.

1.4. Uncertainties, Data Representation, and Key Reactions

1.4.1. Uncertainties

A series of studies on the photochemical modeling of Titan's atmospheric chemistry has demonstrated that model predictions were affected by large uncertainties, due to both structural uncertainty (incompleteness of the chemical scheme) and parametric uncertainty (limited precision of physico-chemical parameters) (Carrasco et al. 2007, 2008a, 2008b; Dobrijevic et al. 2008; Hébrard et al. 2006, 2007, 2009, 2012; Plessis et al. 2012). On the structural side, a lot of data are presently missing concerning the reaction rates (1) of heavier species (neutrals and ions) and the associated products (2) of excited states of primary species, (3) of isomers, and (4) of negative ions. On the parametric side, besides the unavoidable experimental uncertainty of measured rate constants, a major source of uncertainty is due to the necessity to extrapolate rate parameters to temperatures representative of Titan's atmosphere. The extrapolation uncertainty for Arrhenius-type laws can amount to several orders of magnitude (Hébrard et al. 2009). Moreover, less than 10% of all rate constants in the models of Titan's neutral chemistry have been measured at the required temperatures (Hébrard et al. 2009). This fraction is even lower for products branching ratios, where in fact only a handful of determinations have been made. In order to alleviate this problem, it is necessary to install a constructive dialog between modelers and experimentalists/theorists, for which the methodology of Key Reactions Improvement (KRI) is an optimal tool.

1.4.2. Key Reactions Improvement Strategy

A key reaction is a reaction for which, at a given stage of a model's development, the reaction parameters need to be known with better accuracy in order to optimally improve the precision of model predictions (Dobrijevic et al. 2010). A list of key reactions can be generated by sensitivity analysis, which requires the implementation of sound uncertainty management in the modeling procedure (Carrasco et al. 2008b; Hébrard et al. 2009). Given these considerations, the outlook might appear hopeless (several hundreds to thousands of reactions, depending on the model), but all the previous studies show that the Pareto principle, or "80/20 rule," is at work here (Juran & Godfrey 1999). It states that roughly 80% of the effects come from 20% of the causes, which means in the present setup that only a small set of reactions is probably responsible for the major part of prediction uncertainty. One can thus expect to improve significantly the precision of the model's predictions through a few, well-targeted, experimental studies on the most influential key reactions. Incidentally, the work by Hébrard et al. (2009) confirms that the updating, in a model, of reactions not identified as key has no or little impact on the precision of model's predictions.

KRI is an iterative procedure, and new key reactions will be identified once the previous ones have been updated in the model. The process can be stopped when the model predictions are deemed precise enough to achieve a specific goal (such as comparison with observations). However, even if one were to loop indefinitely, and because of experimental precision limits, one should not expect to increase the predictions precision beyond a certain threshold: it has been shown for a Titan photochemical model that when all reaction parameters are allotted a very optimistic 10% relative uncertainty, the mole fractions of the heavier species are still predicted with sizeable relative uncertainty (up to 50%; Peng et al. 2010).

Global sensitivity analysis based on Monte Carlo uncertainty propagation has been used for the identification of key reactions in Titan's atmosphere (Carrasco et al. 2008b; Hébrard et al. 2009). This method requires that all uncertain reaction parameters are represented by a probability density function which describes as faithfully as possible the available information (Dobrijevic et al. 2010). Under- or overestimation of the uncertainty of the reactions parameters is counter-productive, because they can significantly bias the sensitivity analysis and produce "false" key reactions. In order to minimize the investment in long and costly studies of low-temperature kinetics, the uncertainty budget has to be as accurate as possible. This adds a significant load to the modeling task, but it is a necessity for the KRI procedure to be efficient.

1.4.3. Data Representation

A few methodological rules should ideally be respected by all actors in the KRI loop (experimentalists, data analysts, database managers, modelers, ...). (1) Experimentalists should provide detailed uncertainty budget, discriminating the systematic and random contributions. (2) The full variance/covariance matrix of the parameters resulting from the fit of a rate law to the experimental data should be published along with the best values of these parameters. This is necessary to derive extrapolation uncertainty (Hébrard et al. 2009); (3) sources of uncertainty coming from independently measured properties should not be mixed. There is a strong case against the use of partial reaction rates when they derive from the product of a global reaction rate and a product branching ratio. Branching ratios require specific probabilistic representations in order to account for their sum-to-one constraint. This intrinsic constraint cannot be preserved at the level of partial reactions (Carrasco & Pernot 2007; Plessis et al. 2010).

As a consequence, in the following, the separation between rate constants and product branching ratios is explicitly preserved. Best estimates of rate constants and product branching ratios, with their uncertainties, are given for all reactions under Titan's atmospheric conditions. Reported uncertainties are the standard relative uncertainties ux on the reference values x0 (rate constant or branching ratio). To exclude negative values from the confidence intervals resulting from large relative uncertainties (above 30%), one uses a multiplicative representation: a 67% (1σ) confidence interval (CI) can be estimated as [x0/(1 + ux), x0 × (1 + ux)] and a (nσ) CI as [x0/(1 + ux)n, x0 × (1 + ux)n].

Uncertainties on ion–molecule reaction branching ratios are almost never given by the authors. We have estimated them to be ±5% for yields higher than 0.1 and to be ±10% for those which are lower. However, the dispersion of the yields measured by different groups is sometimes outside this uncertainty range, due to other factors, for example, the occurrence of secondary reactions which can alter the product branching ratios.

For the implementation of uncertainty management in chemical models from the data provided in the present paper, we invite the reader to refer to Hébrard et al. (2006, 2007, 2009) for reaction rates of neutrals, to Carrasco et al. (Carrasco et al. 2007, 2008a; Carrasco & Pernot 2007) for ion reactions and to Plessis et al. (2010, 2012) for dissociative recombination reactions, and more generally for the treatment of branching ratios.

1.4.4. Key Reactions Related to the Present Review

Some reactive forms of nitrogen have been found to participate in a set of key reactions for Titan's atmospheric chemistry. For neutrals, photolysis of N2 appears to be an important source of uncertainty at a global level, i.e., having an influence on a large number of species in the upper atmosphere (Hébrard et al. 2009). Recently, Lavvas et al. (2011) have shown that the use of high-resolution N2 photolysis cross sections has a strong impact on the altitude-dependent photolysis rate. A detailed study on the production of HCN/HNC in Titan's upper atmosphere identified N (2D) as a key reactant, notably through its reaction with CH4 (Hébrard et al. 2012). For ions, a set of reactions necessary to reproduce the INMS ions mass spectrum at 1200 km during the T5 flyby was identified by Carrasco et al. (2008b). In this restricted set, N+ (3P) and N+2 play an important role through their reactions with H2, CH4, C2H2, and C2H4. However, it is certain that other reactions involving reactive forms of nitrogen still require review.

1.5. Definitions and Notations

In this paper, the experimental methods used to study the reactions are indicated by acronyms. We used the same acronyms as in the review of Herron (1999) and as in the compilation of Anicich (2003) for neutral reactions and ion–molecule reactions, respectively. All acronyms are listed in Table 2.

Table 2. Acronyms for the Experimental Methods Cited in the Text and in the Tables

CL Chemi-luminescence
CRESU Cinétique de Réactions en Ecoulement
  Supersonique Uniforme
DF Discharge flow
DT Drift tube
ES Emission spectroscopy
ESR Electron spin resonance
FA Flowing afterglow
FJFR Free jet flow reactor
FP Flash photolysis
GIB Guided ion beams
GIB-TOF Guided ion beams-time of flight
ICR Ion cyclotron resonance
LIF Laser-induced fluorescence
LP Laser photolysis
LPI Laser photo-ionization
Opt Optical methods
P Steady-state photolysis
PD Pulsed discharge
PIMS Photo-ionization mass spectrometry
PR Pulsed radiolysis
RA Resonance absorption
REMPI Resonance enhanced multiphoton
  ionization
RF Resonance fluorescence
SIFDT Selected ion flow drift tube
SIFT Selected ion flow tube
SIFT-LIF Selected ion flow tube–Laser-induced
  fluorescence
TPEPICO Threshold photo-electron photo-ion
  coincidences

Download table as:  ASCIITypeset image

We note that in the literature, the probability of reactions is often expressed as reaction cross sections σ, instead of rate constants k, in particular for reactant ions having some initial kinetic energy. The conversion between these two quantities can easily be made in first approximation by k = σ 〈v〉, where k is the rate constant in cm3 s−1, σ is the reaction cross section in cm2, and 〈v〉 is the average relative velocity in cm s−1. A more precise treatment can be found in the paper from Ervin & Armentrout (1985). Reactions can either be measured at a given temperature T (with a Boltzmann distribution of reactant velocities at T) or at given collision energies in the center-of-mass frame (ECM), i.e., when reactants have a velocity in the laboratory frame, defined with an uncertainty by experimental conditions.

2. METASTABLE N2 (A3Σ+u) MOLECULES

2.1. Production of N2 (A 3Σ+u) Molecules

This triplet metastable state cannot be efficiently produced by photoexcitation of N2, because the optical transition from the N2 (X1Σ+g) ground state is dipole forbidden. However, it can be produced by electron excitation, either directly from the N2 (X1Σ+g) ground state or by cascades from higher molecular excited states, in particular by the N2 (C3ΠuB3Πg) emission, followed by the N2 (B3ΠgA3Σ+u) emission. The N2 (A) metastable state is indeed observed in terrestrial aurorae with the N2 (A3Σ+uX1Σ+g) Vegard–Kaplan forbidden emission in the infrared, and several atmospheric models have been developed to fit the observations (see, for example, Morrill & Benesch 1996 and Broadfoot et al. 1997). Very recently Jain et al. (Jain & Bhardwaj 2011) modeled the N2 Vegard–Kaplan emission observed in the Mars dayglow by the SPICAM instrument on board Mars Express (Leblanc et al. 2006, 2007). This emission had been predicted a long time ago by Fox & Dalgarno (1979), but only observed recently. Bhardwaj & Jain (2012) now predict its presence in the Venus dayglow and it could also be the case for Titan.

2.2. Lifetime and Quenching

Among the N2 molecular excited states which can be formed, only the N2(A) state has a sufficiently long lifetime, equal to 2.37 s (Piper 1993), to be able to contribute to Titan's atmospheric chemistry through chemical reactions. Its quenching by collisions with N2 is very inefficient, the rate constant being in the range of 10−18 cm3 s−1 (see the discussion in the review of Herron 1999 and our discussion in Section 2.3.1).

2.3. N2 (A3Σ+u) Chemical Reactions

Collisions of the N2 (A3Σ+u) molecular excited state with molecular targets do not produce the N2 bond breakage, but they result either in quenching of this excited state or in the excitation or dissociation of the target molecule via dissociative energy transfer (see the reviews of Golde 1988 and Herron 1999). In the following tables, we report the rate constant measured for the ground vibrational level of N2 (A3Σ+u). Our recommended rate constant values are also for the ground vibrational level. The reactions with N2, H2, and CH4 have very small rate constants for the vibrational ground state of N2 and are most probably resulting in the quenching of N2 into its ground electronic state. When N2 (A) is in an excited vibrational state, the rate constant is somewhat higher and corresponds to vibrational quenching. On the contrary, the reactions with unsaturated hydrocarbons, C2H2 and C2H4, have large rate constants and lead to dissociation of the target molecule with the release of H atoms. The rate constant dependence with vibrational energy of N2 (A) is rather weak for these two reactions. The reactions of N2 (A3Σ+u) with the C2H6 and C3H8 saturated hydrocarbons are intermediate cases between reactions with N2, H2, and CH4 on one side, and C2H2 and C2H4 on the other side.

2.3.1. Reaction N2 (A 3Σ+u) + N2

Kinetics:

Recommended value at 300 K: k ⩽ 3 × 10−18 cm3 s−1

The different rate constant data for the N2 (A3Σ+u) + N2 reaction can be found in Table 3, which summarizes all published rate constant data for N2 (A3Σ+u) and N (2D, 2P) reactions. The second and fourth columns indicate the temperature and the experimental method used, respectively.

Table 3. Rate Constant Measurements for the Neutral N2 (A 3Σ+u) and N (2D, 2P) Reactions

Reaction T k Method Reference
(K) (cm3 s−1)
N2 (A 3Σ+u) + N2 298 ⩽10 × 10−18 FP-CL (Callear & Wood 1971)
N2 (A 3Σ+u) + N2 298 ⩽3.7 × 10−16 PR-RA (Dreyer & Perner 1973)
N2 (A 3Σ+u) + N2 298 4.5 × 10−17 DF (Vidaud et al. 1976)
N2 (A 3Σ+u) + N2 298 2.6 × 10−18 PD-ES (Levron & Phelps 1978)
N2 (A 3Σ+u) + N2 298 3.7 × 10−17 PD (Suzuki et al. 1993a)
N2 (A 3Σ+u) + N2 298 1.8 × 10−18 PD (Suzuki et al. 1997)
N2 (A 3Σ+u) + N2 298 ⩽3 × 10−18 Compilation (Herron 1999)
N2 (A 3Σ+u) + H2 298 3 × 10−15 FP-CL (Black et al. 1969)
N2 (A 3Σ+u) + H2 298 ⩽7 × 10−15 P-CL (Young et al. 1969)
N2 (A 3Σ+u) + H2 298 3 × 10−15 FP-CL (Callear & Wood 1971)
N2 (A 3Σ+u) + H2 240–370 1.9 × 10−15 a FP-CL (Slanger et al. 1973)
N2 (A 3Σ+u) + H2 298 2.4 × 10−15 PD-ES (Levron & Phelps 1978)
N2 (A 3Σ+u) + H2 298 3.8 × 10−15 DF-LIF (Hack et al. 1988)
N2 (A 3Σ+u) + H2 298 5.8 × 10−15 DF-LIF (Bohmer & Hack 1989)
N2 (A 3Σ+u) + H2 298 3.5 × 10−15 Compilation (Herron 1999)
N2 (A 3Σ+u) + CH4 298 3 × 10−15 FP-CL (Black et al. 1969)
N2 (A 3Σ+u) + CH4 298 ⩽7 × 10−15 P-CL (Young et al. 1969)
N2 (A 3Σ+u) + CH4 298 ⩽2 × 10−14 DF-CL (Meyer et al. 1971)
N2 (A 3Σ+u) + CH4 298 ⩽1.7 × 10−15 FP-CL (Callear & Wood 1971)
N2 (A 3Σ+u) + CH4 300–360 3.2 × 10−15 a FP-CL (Slanger et al. 1973)
N2 (A 3Σ+u) + CH4 298 ⩽1 × 10−14 DF-ES (Clark & Setser 1980)
N2 (A 3Σ+u) + CH4 298 3.0 × 10−15 Compilation (Herron 1999)
N2 (A 3Σ+u) + C2H2 298 1.6 × 10−10 P-CL (Young et al. 1969)
N2 (A 3Σ+u) + C2H2 298 1.6 × 10−10 FP-CL (Callear & Wood 1971)
N2 (A 3Σ+u) + C2H2 298 2.5 × 10−10 DF-CL (Meyer et al. 1971)
N2 (A 3Σ+u) + C2H2 298 2.0 × 10−10 DF-LIF (Bohmer & Hack 1991)
N2 (A 3Σ+u) + C2H2 298 2.0 × 10−10 Compilation (Herron 1999)
N2 (A 3Σ+u) + C2H2 298 (1.40 ± 0.02) × 10−10 for C2H2 LP-LIF (Umemoto 2007)
    (1.45 ± 0.03) × 10−10 for C2D2    
N2 (A 3Σ+u) + C2H4 298 1.2 × 10−10 FP-CL (Black et al. 1969)
N2 (A 3Σ+u) + C2H4 298 1.5 × 10−10 P-CL (Young et al. 1969)
N2 (A 3Σ+u) + C2H4 298 1.1 × 10−10 FP-CL (Callear & Wood 1971)
N2 (A 3Σ+u) + C2H4 298 1.6 × 10−10 DF-CL (Meyer et al. 1971)
N2 (A 3Σ+u) + C2H4 298 0.64 × 10−10 DF-LIF (Dreyer & Perner 1973)
N2 (A 3Σ+u) + C2H4 298 1.2 × 10−10 DF-ES (Clark & Setser 1980)
N2 (A 3Σ+u) + C2H4 298 1.2 × 10−10 DF-CL (Cao & Setser 1985)
N2 (A 3Σ+u) + C2H4 298 1.0 × 10−10 DF-LIF (Thomas et al. 1987)
N2 (A 3Σ+u) + C2H4 298 1.2 × 10−10 DF-ES (Ho & Golde 1991)
N2 (A 3Σ+u) + C2H4 298 1.1 × 10−10 Compilation (Herron 1999)
N2 (A 3Σ+u) + C2H4 298 (0.97 ± 0.04) × 10−10 for C2H4 LP-LIF (Umemoto 2007)
    (0.93 ± 0.04) × 10−10 for C2D4    
N2 (A 3Σ+u) + C2H6 298 ⩽5 × 10−15 P-CL (Young et al. 1969)
N2 (A 3Σ+u) + C2H6 298 3.6 × 10−13 FP-CL (Callear & Wood 1971)
N2 (A 3Σ+u) + C2H6 298 ⩽2 × 10−14 DF-CL (Meyer et al. 1971)
N2 (A 3Σ+u) + C2H6 298 2.9 × 10−13 PR-AS (Dreyer & Perner 1973)
N2 (A 3Σ+u) + C2H6 300–370 2.2 × 10−13 a FP-CL (Slanger et al. 1973)
N2 (A 3Σ+u) + C2H6 298 2.3 × 10−13 Compilation (Herron 1999)
N2 (A 3Σ+u) + C3H8 298 1.3 × 10−12 FP-ES (Callear & Wood 1971)
N2 (A 3Σ+u) + C3H8 298 1.3 × 10−12 Compilation (Herron 1999)
N (2D) + N2 298 ⩽6 × 10−15 FP-CL (Black et al. 1969)
N (2D) + N2 298 1.6 × 10−14 DF-RA (Lin & Kaufman 1971)
N (2D) + N2 298 2.3 × 10−14 FP-RA (Husain et al. 1972)
N (2D) + N2 298 1.5 × 10−14 FP-RA (Husain et al. 1974)
N (2D) + N2 298 ⩽1.8 × 10−14 FP-CL (Black et al. 1969)
N (2D) + N2 198–372 1.8 × 10−14a FP-CL (Slanger & Black 1976)
N (2D) + N2 298 (1.3 ± 0.2) × 10−14 PR-RA (Sugawara et al. 1980)
N (2D) + N2 213–294 2.4 × 10−14a PR-RA (Suzuki et al. 1993b)
N (2D) + N2 298 1.7 × 10−14 Compilation (Herron 1999)
N (2P) + N2 400 6 × 10−14 DF-RA (Lin & Kaufman 1971)
N (2P) + N2 298 ⩽3 × 10−16 DF-RA (Husain et al. 1972)
N (2P) + N2 298 1.0 × 10−16 FP-RA (Husain et al. 1974)
N (2P) + N2 298 (3.3 ± 0.3) × 10−17 PR-RA (Sugawara et al. 1980)
N (2P) + N2 298 ⩽4 × 10−16 PR-RA (Umemoto et al. 1985)
N (2P) + N2 298 3.3 × 10−17 Compilation (Herron 1999)
N (2D) + H2 298 5 × 10−12 FP-CL (Black et al. 1969)
N (2D) + H2 298 1.7 × 10−12 FP-RA (Husain et al. 1972)
N (2D) + H2 298 2.1 × 10−12 FP-RS (Husain et al. 1974)
N (2D) + H2 298 2.7 × 10−12 FP-CL (Black et al. 1975)
N (2D) + H2 298 3.5 × 10−12 DF-ESR (Fell et al. 1981)
N (2D) + H2 298 2.3 × 10−12 DF-RF (Piper et al. 1987)
N (2D) + H2 298 1.8 × 10−12 DF-RA (Whitefield & Hovis 1987)
N (2D) + H2 213–300 2.4 × 10−12 a PR-RA (Suzuki et al. 1993b)
N (2D) + H2 298 2.3 × 10−12 LP-LIF (Umemoto et al. 1998a)
N (2D) + H2 298 2.2 × 10−12 Compilation (Herron 1999)
N (2P) + H2 298 3.0 × 10−15 FP-RA (Husain et al. 1972)
N (2P) + H2 298 1.9 × 10−15 FP-RA (Husain et al. 1974)
N (2P) + H2 298 ⩽8 × 10−16 DF-ES (Young & Dunn 1975)
N (2P) + H2 298 1.4 × 10−14 PR-RA (Umemoto et al. 1985)
N (2P) + H2 213–300 1.4 × 10−14a PR-RA (Suzuki et al. 1993b)
N (2P) + H2 298 1.9 × 10−15 Compilation (Herron 1999)
N (2D) + CH4 298 3 × 10−12 FP-CL (Black et al. 1969)
N (2D) + CH4 298 4.6 × 10−12 DF-ESR (Fell et al. 1981)
N (2D) + CH4 298 3.3 × 10−12 for CH4 LP-LIF (Umemoto et al. 1998b)
    2.0 × 10−12 for CD4    
N (2D) + CH4 223–292 5.4 × 10−12 for CH4a PR-RA (Takayanagi et al. 1999)
    3.2 × 10−12 for CD4a    
N (2D) + CH4 298 4.0 × 10−12 Compilation (Herron 1999)
N (2P) + CH4 298 7.8 × 10−14 PR-RA (Umemoto et al. 1985)
N (2P) + CH4 223–292 9.3 × 10−14 for CH4b PR-RA (Takayanagi et al. 1999)
    6.0 × 10−14for CD4b    
N (2P) + CH4 298 8.55 × 10−14 Compilation (Herron 1999)
N (2D) + C2H2 298 1.1 × 10−10 DR-ESR (Fell et al. 1981)
N (2D) + C2H2 223–293 6.5 × 10−11 for C2H2c PR-RA (Takayanagi et al. 1998b)
    6.25 × 10−11for C2D2c    
N (2D) + C2H2 298 6.5 × 10−11 Compilation (Herron 1999)
N (2P) + C2H2 295 3.2 × 10−11 PR-RA (Umemoto et al. 1985)
N (2P) + C2H2 223–293 2.3 × 10−11 for C2H2c FP-RA (Takayanagi et al. 1998b)
    2.0 × 10−11 for C2D2c    
N (2P) + C2H2 298 2.3 × 10−11 Compilation (Herron 1999)
N (2D) + C2H4 298 1.2 × 10−10 FP-CL (Black et al. 1969)
N (2D) + C2H4 298 3.7 × 10−11 PR-RA (Sugawara et al. 1980)
N (2D) + C2H4 298 8.3 × 10−11 DR-ESR (Fell et al. 1981)
N (2D) + C2H4 298 4.3 × 10−11 Compilation (Herron 1999)
N (2D) + C2H4 230–292 4.3 × 10−11 for C2H4c PR-RA (Sato et al. 1999)
    3.8 × 10−11 for C2D4c    
N (2P) + C2H4 298 2.8 × 10−11 PR-RA (Sugawara et al. 1980)
N (2P) + C2H4 298 3.2 × 10−11 PR-RA (Umemoto et al. 1985)
N (2P) + C2H4 230–292 3.0 × 10−11 (±10%) for C2H4c PR-RA (Sato et al. 1999)
    3.0 × 10−11 (±10%) for C2D4c    
N (2P) + C2H4 298 3.0 × 10−11 Compilation (Herron 1999)
N (2D) + C2H6 298 2.7 × 10−11 DF-ESR (Fell et al. 1981)
N (2D) + C2H6 298 2.1 × 10−11 LP-LIF (Umemoto et al. 1998b)
N (2D) + C2H6 298 1.9 × 10−11 Compilation (Herron 1999)
N (2P) + C2H6 298 5.4 × 10−13 FP-CL (Umemoto et al. 1985)
N (2D) + C3H8 298 4.6 × 10−11 DF-ESR (Fell et al. 1981)
N (2D) + C3H8 298 3.1 × 10−11 LP-LIF (Umemoto et al. 1998b)
N (2D) + C3H8 298 2.9 × 10−11 Compilation (Herron 1999)
N (2P) + C3H8 298 1.9 × 10−12 FP-CL (Umemoto et al. 1985)

Notes. aRate constant value at 300 K. bRate constant value at 292 K. cRate constant value at 298 K.

Download table as:  ASCIITypeset images: 1 2 3

There is a considerable scattering of the measured rate constants, due to the experimental challenge of measuring such a slow reaction. Dreyer & Perner (1973), as well as Levron & Phelps (1978) also measured the rate constant for v = 1, but it remains very small. Herron (1999) recommended an upper limit of 3 × 10−18 cm3 s−1, disregarding the values which are above 10−17 cm3 s−1, as Vidaud et al. (1976) used a discharge flow system which does not produce only the N2 (A3Σ+u) state and Suzuki et al. (1997) corrected later their first measurement (Suzuki et al. 1993a). This very small rate constant demonstrates that the quenching of this excited state by collision with N2 is not efficient.

2.3.2. Reaction N2 (A 3Σ+u) + H2

Kinetics:

Recommended value at 300 K: k = 3.5 × 10−15 cm3 s−1 (±60%)

The reaction of N2 (A3Σ+u) with molecular hydrogen has been studied by many groups (see Table 3), but there is no recent study. The rate constant values are in relatively good agreement and the recommended value is the same as the one proposed by Herron (1999), who eliminated the three oldest values. Slanger et al. (1973) measured the temperature dependence of the rate constant as k = 2.2 × 10−10 exp (−3500/T) cm3 s−1 over the 240–370 K temperature range. However for the excited vibrational levels, there are large discrepancies between the measurements made by Hack et al. (1988) and Bohmer & Hack (1989) who observe a level-off of the rate constant value at v = 1 and higher vibrational levels, and measurements of Golde et al. (Golde et al. 1989) who observe a continuous increase of the rate constant from v = 2 and higher vibrational levels. This discrepancy could be due to experimental uncertainties. Bohmer & Hack (1989) observed H atom release, but did not quantify it. It seems likely that the reaction is mainly quenching of the nitrogen excited state, due to the low value of the rate constant, but also produces some dissociation of H2. This is confirmed by quasi-classical trajectory calculations (Sperlein & Golde 1989), showing that this reaction leads to electronic quenching and vibrational relaxation, but also to dissociative energy transfer with branching ratios which vary with the vibrational level of N2 (A).

2.3.3. Reaction N2 (A 3Σ+u) + CH4

Kinetics:

Recommended value at 300 K: k = 3.0 × 10−15 cm3 s−1 (±60%)

Our recommended rate constant value is the same one recommended by Herron (1999) which is based on the measurements of Slanger et al. (1973). Slanger et al. (1973) measured the temperature dependence of the rate constant as k = 1.3 × 10−10 exp (−3170/T) cm3 s−1 over the 300–360 K temperature range. Several authors measured the rate constant as a function of the vibrational level of N2 (A3Σ+u) (Clark & Setser 1980; Golde et al. 1989; Piper et al. 1985; Thomas et al. 1983), which can be higher than for v = 0 by a factor of 100 and can be interpreted as vibrational quenching into the vibrational ground state. For v = 0, the products have not been measured. Golde et al. (1989) detected H atom products, but could not quantify this reaction channel.

2.3.4. Reaction N2 (A 3Σ+u) + C2H2

Kinetics:

Recommended values at 300 K: k = 1.40 × 10−10 cm3 s−1 (±60%) for C2H2

k = 1.45 × 10−10 cm3 s−1 (±60%) for C2D2.

This reaction has a high rate constant and the different measurements are in rather good agreement (see Table 3). We recommend the most recent rate constant value measured by Umemoto (2007), with an uncertainty that takes into account the scattering of the rate constants measured by other authors. Umemoto also studied the reaction with C2D2 which has almost the same rate constant as for C2H2. Bohmer & Heck (1991) did not observe any dependence of the rate constant with the vibrational energy of N2 (A). Both the high rate constant and the absence of vibrational dependence are explained by the fact that, as opposed to reactions with N2, H2, and CH4, this reaction is not mainly quenching of N2 (A) but produces the dissociation of C2H2, as shown by Umemoto et al. (Umemoto 2007).

Products:

Recommended yields: (N2 + C2H + H)/(N2 + C2 + H2), 0.52/0.48 for the reaction with C2H2

(N2 + C2D + D)/(N2 + C2 + D2), 0.33/0.67 for the reaction with C2D2.

Umemoto (2007) measured the H atom yield and D atom yield to be equal to 0.52 and 0.33, for the reactions with C2H2 and C2D2, respectively. They conclude that the presence of an isotope effect in the H/D atom yield suggests that the H/D release competes with the H2/D2 molecule release, the latter being favored in the case of C2D2. As the rate constant has a high value, the quenching is most probably a negligible channel, compared to the collision-induced C2H2 dissociation.

2.3.5. Reaction N2 (A 3Σ+u) + C2H4

Kinetics:

Recommended values at 300 K: k = 9.7 × 10−11 cm3 s1 (±60%) for C2H4

k = 9.3 × 10−11 cm3 s−1 (±20%) for C2D4.

As for acetylene, this reaction has a high rate constant and the different measurements are in rather good agreement (see Table 3). We recommend the rate constant value of the most recent work by Umemoto (2007), with an uncertainty which takes into account the scattering of the rate constants measured by other authors. Umemoto also studied the reaction with C2D4 which has almost the same rate constant as the reaction with C2H4. Dreyer & Perner (1973) and Thomas et al. (1987) measured a weak dependence of the rate constant with the vibrational energy of N2 (A). Similarly to the case of acetylene, both the high rate constant and the quasi-absence of vibrational dependence are explained by the fact that this reaction produces the dissociation of C2H4, as shown by Umemoto (2007).

Products:

Recommended yields: (N2 + C2H3 + H)/(N2 + C2H2 + H2), 0.30/0.70 for the reaction with C2H4

(N2 + C2D3 + D)/(N2 + C2D2 + D2), 0.13/0.87 for the reaction with C2D4.

Umemoto (2007) measured the H and D atom yields to be equal to 0.30 and 0.13, respectively. Umemoto also concluded that this isotope effect in the H/D atom yield suggests that the H/D release competes with the H2/D2 release, the latter being favored in the case of C2D4.

2.3.6. Reaction N2 (A 3Σ+u) + C2H6

Kinetics:

Recommended value at 300 K: k = 2.3 × 10−13 cm3 s−1 (±60%)

Our recommended rate constant value is the same as the one recommended by Herron (1999) which is based on the measurements of Slanger et al. (1973), slightly adjusted. It is higher by two orders of magnitude than that for the reaction with methane. Slanger et al. (1973) measured the temperature dependence of the rate constant as k = 1.6 × 10−10 exp (−1980/T) cm3 s−1 over the 300–370 K temperature range. The products are not known, but the value of the rate constant suggests that dissociation of the target molecule is a possible reaction channel in addition to quenching. Several authors (Clark & Setser 1980; Dreyer & Perner 1973) measured the rate constant as a function of the vibrational level of N2 (A3Σ+u), which can be higher than for v = 0 by a factor of 100 and can be interpreted as vibrational quenching into the vibrational ground state.

2.3.7. Reaction N2 (A 3Σ+u) + C3H8

Kinetics:

Recommended value at 300 K: k = 1.3 × 10−12 cm3 s−1 (±60%).

The only existing data obtained by Callear & Wood (1971) give a rate constant 1.3 × 10−12 cm3 s−1. The uncertainty can be estimated to be about the same as for the reaction of N2 (A) with C2H4 which is of the same order of magnitude. As for C2H6, the products are not known, but the value of the rate constant suggests that, in addition to quenching, the dissociation of the target molecule is a more important channel than in the case of the reaction with C2H6.

3. NITROGEN ATOMS IN THE N (4S), N (2D), AND N (2P) STATES

N2 dissociation produces N atoms. As discussed in detail below, the atom yields comprise about 50% N (4S) ground-state atoms and about 50% long-lived N (2D) and N (2P) metastable state atoms. The energies of the long-lived N (2D) and N (2P) metastable states lie 2.38 eV and 3.57 eV above the N (4S) ground state, respectively. The N (4S) ground state is almost non-reactive (see below), as opposed to the N (2D) and N (2P) metastable states, which are reactive with stable molecules. There are many more experimental and theoretical studies concerning the N (2D) state than the N (2P) state. Generally, the N (2P) state is much less reactive than the N (2D) state, with rate constants lower than for the N (2D) by several orders of magnitude. Therefore, according to most authors, the main reaction channel of N (2P) reactions is quenching into the N (2D) state (or maybe also N (4S)), as opposed to the N (2D) metastable state which is very reactive.

3.1. Production of N Atoms

Atomic nitrogen can be produced by N2 dissociation induced by vacuum ultraviolet (VUV) photolysis, electron impact, dissociative photoionization, and N+2 dissociative recombination.

3.1.1. N Atom Production by UV Photodissociation of N2

The dissociation processes of molecular nitrogen are not very well known. The absorption of VUV photons by the N2 molecule induces dipole-allowed excitation from the ground gerade state to several singlet ungerade states with high energy content. These high-energy states are strongly bound, but can undergo predissociation through the crossing with dissociative triplet states. The photoabsorption spectrum of N2 consists of very highly structured bands and as the solar spectrum is also highly structured in this spectral region, the correct calculation of atmospheric penetration requires a very high spectral resolution. The first band system of N2 (Lyman–Birge–Hopfield bands from the ground state 1Σ+g to the excited 1Πg state) is in the region between 8.26 and 12.4 eV, but this transition is forbidden by electric dipole selection rules and the absorption cross sections are very small. Let us note that those bands are nevertheless observed in the airglow emission spectra of Titan (Ajello et al. 2008). The dissociation energy of N2 is 9.76 eV. In principle, therefore, the forbidden transitions to vibrationally excited levels of the a1Πg state can provide enough energy to dissociate the molecule. However, the a1Πg state does not correlate with N (4S) + N (4S) and predissociation via spin–orbit coupling to the ground state of N2 molecules has never been observed. Absorption of UV light becomes significant only near 12.4 eV and the N2 spectrum shows a strong banded structure between 12.4 and 18.8 eV and a continuum above 18.8 eV. The first electric dipole allowed and intense transition is toward the b1Πu valence state (E < 12.5 eV). This state correlates with N (2D) + N (2D) but the energy necessary to reach the dissociation limit is 14.5 eV. Nonetheless, the b1Πu state can undergo predissociation at lower energies via intersystem crossing to the C 3Πu and C 3u states, producing N (4S) + N (2D) above 12.1 eV (Lewis et al. 2005). Predissociation competes with spontaneous emission, which is indeed observed for most vibrational levels of the b1Πu state (Itikawa 2006). Sprengers et al. (2004) measured a predissociation yield of only ∼28% for the v = 1 level of the N2 b1Πu state. Higher valence singlet states populated by allowed transition at higher energy can also undergo predissociation through intersystem crossing (spin–orbit coupling) and can lead to atomic nitrogen in the second electronically excited state 2P above 13.3 eV.

Because of the different reactivity of ground 4S and excited 2D and 2P atomic nitrogen and the long radiative lifetime of the 2D and 2P states, it would be quite important to characterize the N2 predissociation/dissociation yield to quantify the relative concentration of the three states and their role in the atmospheric chemistry of Titan. Unfortunately, the N (2P), N (2D), and N (4S) production yields have been measured only for a few specific rovibrational of several electronic states of N2 by Helm & Cosby (1989) and Walter et al. (1993), not including the b1Πu state (the first state accessible by an allowed transition). The observations are sparse and do not allow the derivation of a general model that can describe quantitatively the predissociation/dissociation products of N2. Nevertheless, some observations of the authors can help in drawing more general conclusions. They have observed that the predissociation of the above-mentioned levels always produces one ground state atom and one excited atom, either in the 2D or 2P states, the energetic limit of the N (2D) + N (2D) channel (14.5 eV) being too high to be reached in their experiments. Remarkably, no dissociation was observed to the ground state N (4S) + N (4S).

Bakalian (2006) proposed in 2006 a simplified scheme of the N2 dissociation for modeling the production of hot nitrogen atoms in the Martian thermosphere. More recently, Lavvas et al. (2011), in a work concerning the energy deposition in the upper atmosphere of Titan, summarized most of the recent information on the electronic transitions of N2. This paper includes detailed calculations of high-resolution photoabsorption and photodissociation cross sections computed using a coupled-channel Schrödinger equation quantum-mechanical model. These authors suggest the following scheme for the N2 photodissociation:

Equation (3a)

Equation (3b)

Equation (3c)

The choices of energy threshold were dictated by the fact that E = 12.1 eV corresponds to the asymptotic energy of the channel (3a), E = 14.5 eV corresponds to the asymptotic energy of channel (3c), while E = 13.9 eV corresponds to the energy derived by Walter et al. (1993) above which the fragmentation into channel (3b) dominates over that into channel (3a). Nevertheless, this observation was related to the excited states populated in the experiment by Walter et al. (1993) and not the most intense transition to b1Πu. Therefore, this scheme might be oversimplified and supplementary high-resolution absolute experimental cross sections complemented by dissociation branching ratios are highly desirable.

In conclusion, it is difficult to recommend a general N2 photodissociation scheme for the modeling of the atmosphere of Titan because of the fragmentation of the available information, both regarding the predissociation probabilities for all the involved molecular electronic states and their product branching ratios. We suggest using the scheme of Lavvas et al. (2011), until further experimental and theoretical work emerges.

3.1.2. N Atom Production by Electron Impact Dissociation of N2

In the upper atmosphere of Titan, N2 molecules are subject to a significant bombardment by energetic (200 eV) magnetospheric electrons (Strobel & Shemansky 1982) and by lower energy electrons produced both by photoionization processes and by magnetospheric electrons impact creating secondary electrons. Electron impact can induce ionization, dissociative ionization, excitation, and dissociation. A comprehensive review on the processes that follow electron collision with N2 (including elastic scattering, rotational and vibrational excitation, electronic excitation, dissociation and ionization) is given, and the dependence on the electron energy examined, in (Itikawa 2006). Since there are no selection rules, the collision of N2 molecules with electrons can populate a plethora of N2 states, including low-energy triplet states. One important effect is the observation of emission from many excited N2 states, as well as from its neutral and ion fragments in highly excited atomic levels. The fluorescence yield from the various states depends on the electron energy. The electron-induced dissociation cross section reaches a maximum of 1.2 × 10−16 cm2 at 40–100 eV and, at these energies, is comparable to the ionization cross section.

With the present knowledge, it is reasonable to assume that the N2 dissociation processes by electron impact produce as many N atoms in the 4S ground state as in the two 2D and 2P metastable states, with a much higher yield of 2D state compared to 2P state.

3.1.3. N Atom Production by Other Processes

N atoms are also formed by dissociative ionization of N2 by photons or electrons (see Section 5.1 below). The excited electronic states of N+2, at energies higher than 26.3 eV, dissociate mainly into N+ (3P) + N (2D) and somewhat into other dissociation channels (Aoto et al. 2006; Nicolas et al. 2003a). So globally it can be estimated that N atoms coming from the N2 dissociative ionization are mainly in the 4S ground state, with less than 10% being in the N (2D) excited state. Nitrogen atoms in the 2D state were also observed following electron impact dissociative ionization when using high electron energies (Van Brunt & Kieffer 1975).

Finally, electron–ion recombination of N+2 produces fast N atoms in the ground and metastable states, with a branching ratio which can be estimated to 0.7/0.05/0.25 for N(2D)/N (2P)/N(4S) (see Section 4.3.8 below).

3.2. Lifetime and Quenching

The radiative lifetime of the N (2D) metastable state is quite long (13.6 and 36.7 hr, for the two sub-states 2D3/2 and 2D5/2, respectively; Ralchenko et al. 2011), because this transition is strongly forbidden. In addition, collisional deactivation of N (2D) by N2 is a slow process (see below) and therefore the main fate of N (2D) above 800 km is chemical reaction with other constituents of Titan's atmosphere.

The radiative lifetime of the N (2P) metastable state is also long (11.1 and 10.5 s, for the two sub-states 2P1/2 and 2P3/2, respectively; Ralchenko et al. 2011), compared to the collision frequency in Titan's atmosphere conditions. Quenching by reaction with N2 is a rather slow process (see below) and it is not clear whether this quenching produces N (4S) or N (2D), see the review by Herron (1999).

3.3. N Atom Chemical Reactions

Until the late 1970s, N atoms in their (4S) ground state were believed to induce reactions with hydrocarbons (Herron 1966; Herron & Huie 1968; Sato et al. 1979), the reaction with H2 being endothermic. However, it was later demonstrated that the weak reactivity with acetylene and ethylene, observed at temperatures from about 320 K and above, were not due to this species, but either to excited states still present in the experiments or to reactions with free radicals created together with N atoms (Michael 1979, 1980; Michael & Lee 1977; Umemoto et al. 1986). It is now admitted that N (4S) atoms are neither reactive with hydrocarbon molecules (Sato et al. 1999), nor with N2 and H2 and these reactions are therefore not considered in this review paper. However, N (4S) atoms are reactive with hydrocarbon radicals, in particular with CH3, this reaction being the main loss channel of N (4S) in Titan's atmosphere (Yelle et al. 2010). We will therefore review the reactions of N (4S) atoms with some important radicals from Titan's atmosphere: CH2, CH3, C2H3, and C2H5. We will also discuss the reaction of N (4S) with H atoms.

Laboratory experiments on N (2D) and N (2P) reactions have been rather sparse; remarkably, reliable kinetic data on the reactions of N (2D) with the hydrocarbons present in the atmosphere of Titan became available only in the late 1990s. This lack of data arises because, in general, experimental studies of these reactive systems are problematic due to the difficulty in producing a sufficiently large amount of the electronically excited atoms without other interfering states or species. In his review, Herron (1999) reported a comprehensive compilation of the data on N (2D) and N (2P) reactions. Most of the evaluations performed in this review paper are still valid, especially as far as the kinetic data are concerned. However, a survey of the more recent literature reveals that significant improvements have been made regarding the identification of the reaction mechanisms and products for some of these reactions; these new data arising from studies of the reaction dynamics. In this context, starting from the evaluation of Herron (1999), we will compile together previous and more recent data exclusively for the N (2D) and N (2P) reactions with the main constituents of Titan's atmosphere.

3.3.1. Reactions of N (4S) Ground State with Radicals

3.3.1.1. Reaction N (4S) + H

Kinetics:

Estimated value at 150 K: k = 1.0 × 10−19 cm3 s−1

This reaction can produce NH by radiative association or three body collisions. The rate constant for three body collisions has been measured by Brown (1973) giving a value of 4.8 × 10−32 cm6 s−1 (±35%). However, this reaction will not occur under Titan's atmospheric conditions as, at the altitudes where N atoms are formed (>1000 km), the pressure is not sufficient to allow triple collisions. The radiative association rate constant can be estimated from theoretical calculations for the analogous reaction of Li with H giving LiH (Gianturco & Giorgi 1997). It gives an estimated value of about 1.0 × 10−19 cm3 s−1 at 150 K, with a large uncertainty. This value is very low, so this reaction is very unlikely to occur in Titan's atmosphere.

3.3.1.2. Reaction N (4S) + CH2

Kinetics:

Estimated value at 300 K: k = 8.0 × 10−11 cm3 s−1

The reaction of N (4S) with CH2 has two exothermic channels:

N (4S) + CH2 → HCN + H − 5.29 eV

            → HNC + H − 4.72 eV.

This reaction has not been studied experimentally. However, there is no barrier for this reaction, according to the ab initio calculations of Herbst et al. (2000), so it should be efficient. The N (4S) + CH2 (3B1) entrance channel correlates adiabatically to potential surfaces of sextet, quadruplet, and doublet spin multiplicity and products correlate only to doublet potential surfaces. So this leads to a reduction of the calculated capture rate constant by a factor equal to 1/6. In these conditions, Herbst et al. calculated the rate constant with its temperature variation to be 8.0 × 10−11 × (T/300)0.17 cm3 s−1 (Herbst et al. 2000).

Products:

Estimated yields: (HCN + H)/(HNC + H), 0.50/0.50

The main products from the ab initio calculations of Herbst et al. (Herbst et al. 2000) are HCN + H. The isomerization energy of HCN into HNC is 1.93 eV (DePrince & Mazziotti 2008). There is so much excess energy available in the title reaction (5.29 eV) that the HCN product is able to undergo efficient isomerization after production. It leads to a near equal branching ratio between the two reaction channels (HNC + H) and (HCN + H), according to Herbst et al. (2000).

3.3.1.3. Reaction N (4S) + CH3

Kinetics:

Recommended value at 300 K: k = 8.5 × 10−11 cm3 s−1 (±25%)

The N (4S) + CH3 reaction has been studied experimentally over the temperature range 200–423 K by Marston et al. (1989a). These authors measured a rate constant equal to (8.5 ± 2.0) × 10−11 cm3 s−1 at 298 K, with a strong and complex negative temperature dependence, which seems to deviate from the Arrhenius law. The authors propose two different expressions for the global rate constant: an Arrhenius law, k (T) = 4.3 × 10−10 exp (−420/T) cm3 s−1 or a more complex expression, k (T) = 6.2 × 10−11 + 2.2 × 10−9 exp (−1250/T) cm3 s−1. Their results suggest that the rate constant in the 150–170 K range will not be substantially lower than the value measured at 200 K, which is (6.4 ± 2.1) × 10−11 cm3 s−1. So we recommend the use of a constant value of 6.2 × 10−11 cm3 molecule−1 s−1 between 150 K and 200 K.

Products:

Recommended yields: (H2CN + H)/(HCN + H + H), 0.90/0.10.

The N (4S) + CH3 reaction has three exothermic channels:

N (4S) + CH3 → H2CN + H −1.59 eV

            → HCN + H + H −0.50 eV

            → HCN + H2 −5.02 eV.

Product branching ratios have been obtained for the N(4S) + CH3 and N(4S) + CD3 reactions (Marston et al. 1989b), leading to H2CN + H (85%–100%) with some HCN production (0–15%). The authors suggest that HCN formation is associated with H2. However, the (HCN + H2) production is spin-forbidden and needs an intersystem crossing process to occur. Additionally, recent ab initio calculations (Cimas & Largo 2006) found almost 100% of H2CN production, in good agreement with previous calculations (Nguyen et al. 1996). As H2CN may have enough internal energy to overcome the dissociation barrier for C–H dissociation, some (HCN + H + H) may be produced (Nguyen et al. 1996). Moreover, as Marston et al. (1989b) used an excess of nitrogen atoms, some HCN molecules may come either from the reaction H2CN + N → HCN + NH or more likely from the reaction H2CN + N → CH2 + N2 followed by N + CH2 → HCN + H. Taking in account the various uncertainties, we choose to recommend a branching ratio equal to 90% for H2CN + H formation and 10% for HCN + H + H. As Marston et al. (1989b) observed no isotopic effect, we recommend to take the same branching ratio for the reaction with CD3.

3.3.1.4. Reaction N (4S) + C2H3

Kinetics:

Recommended value at 300 K: k = 7.7 × 10−11 cm3 s−1 (±40%)

Payne et al. (1996) studied this reaction at 298 K, measuring the rate constant to be equal to (7.7 ± 2.9) × 10−11 cm3 s−1.

Products:

Recommended yields: (CH2CN + H)/(C2H2 + NH), 0.83/0.17.

This reaction has three exothermic channels:

N (4S) + C2H3 → CH2CN + H −3.11 eV

            → CH2 + HCN −2.57 eV

            → C2H2 + NH −1.96 eV.

Payne et al. (1996) measured the branching ratio of this reaction at 298 K to be (CH2CN + H)/(C2H2 + NH) equal to 0.83/0.17. They did not observe the (CH2 + HCN) channel. Theoretical calculations of Sun et al. (2004) are in relatively good agreement, predicting (CH2CN + H) to be the main exit channel. However, these calculations show the possible (CH2 + HCN) contribution and the existence of a barrier in the entrance channel for the direct H atom abstraction. We recommend the values of the experimental branching ratios, hoping that further experimental work could check the predicted formation of the (CH2 + HCN) channel.

3.3.1.5. Reaction N (4S) + C2H5

Kinetics:

Recommended value at 300 K: k = 1.1 × 10−10 cm3 s−1 (±25%)

The rate constant for this reaction has been measured at 298 K by Stief et al. (1995) to be equal to (1.1 ± 0.3) 10−10 cm3 s−1. Based on the comparison with the N (4S) + CH3 reaction, these authors suppose that the temperature dependence of the rate constant down to 200 K would be small or negligible.

Products:

Recommended yields: (C2H4 + NH)/(H2CN + CH3), 0.65/0.35

The main exothermic channels of the reaction N (4S) + C2H5 are

N (4S) + C2H5 → H2CN + CH3 −2.07 eV

            → C2H4 + NH −1.93 eV

            → CH3CHN + H −1.83 eV.

Stief et al. (Stief et al. 1995) measured the branching ratio of the reaction of N (4S) with deuterated C2D5. They observed two reactive channels, (C2D4 + ND) being the major channel, with a branching ratio (C2D4 + ND)/(D2CN + CD3) equal to 0.65/0.35. No uncertainties are given for the branching ratio. These authors consider implicitely that the branching ratio should be the same for the reaction with C2H5. They did not observe the (CH3CHN + H) products, which could also be formed according to the theoretical calculations by Yang et al. (2005). Stief et al. (1995) estimate the branching ratio by measuring the C2H4/C2H5 ratio by comparison with a known C2H4 concentration, the C2H5 being estimated from initial F atom concentration (F + C2H6 → C2H5 + HF) and also from the C2H5 decay. Their branching ratio measurements are very sensitive to uncertainty in the initial C2H5 concentration. The C2H4 + NH production might be overestimated, particularly at low temperatures, as this channel is calculated to involve a barrier in the entrance reactive channel for the direct H atom abstraction or to have a much higher transition state than the H2CN + CH3 exit channel (Yang et al. 2005).

3.3.2. Reactions of N (2D) and N (2P) Metastable States with Molecules

3.3.2.1. Reaction N (2D) + N2

Kinetics:

Recommended value at 300 K: k = 1.7 × 10−14 cm3 s−1 (±40%)

N2 being the most abundant component in the atmosphere of Titan, the role of the physical quenching of N (2D) by its parent molecule needs to be assessed. Herron has suggested 1.7 × 10−14 cm3 s−1 as the recommended room temperature rate constant, which is an average value of the most accurate available data (Husain et al. 1974; Lin & Kaufman 1971; Slanger & Black 1976; Sugawara et al. 1980; Suzuki et al. 1993b) (see Table 3). Such a low value was explained by the inefficiency of crossing between doublet and quartet N3 potential energy surfaces (Donovan & Husain 1970). Interestingly, both Suzuki et al. (1993b) and Slanger & Black (1976) investigated the temperature dependence of this rate constant over the range 213–294 and 198–372 K, respectively. The temperature dependence obtained in the two experiments is, however, quite different being k = 5.4 × 10−11 exp (−1620/T) cm3 s−1 for Suzuki et al. (1993b) and k = 1.0 × 10−13 exp (−(510 ± 156)/T) cm3 s−1 for Slanger & Black (1976). Notwithstanding the disagreement in the rate expression, it is relevant to note that the rate constant decreases with decreasing temperature in both cases. It seems to us reasonable to recommend to use the temperature dependence measured by Slanger & Black (1976) in the 198–372 K range, to extrapolate the rate constant at Titan's atmospheric temperatures.

3.3.2.2. Reaction N (2P) + N2

Kinetics:

Recommended value at 300 K: k = 3.3 × 10−17 cm3 s−1 (±60%)

The recommended value for the rate constant is the one from Sugarawa et al. (1980), as proposed by Herron (1999) and is lower than the rate constant of the N (2D) state by several orders of magnitude. According to Suzuki et al. (1993b), the main reaction channel of N (2P) with N2 involves quenching of the metastable state, but it most probably produces N (2D) atoms, as the quenching into N (4S) + N2 is spin forbidden.

3.3.2.3. Reaction N (2D) + H2

Kinetics:

Recommended value at 300 K: k = 2.2 × 10−12 cm3 s−1 (±25%)

This reaction is the best investigated reaction involving the 2D state of atomic nitrogen, so far. The room temperature rate constant recommended by Herron is the average value of the most reliable available data (see Table 3). The rate expression as a function of the temperature has been determined by Suzuki et al. (1993b) in the 213–300 K range and the reported expression is k = 4.6 × 10−11 exp (−880/T) cm3 s−1.

Products: NH (X 3Σ) + H.

The reaction dynamics has been extensively investigated and it is actually one of the few insertion reactions for which a comparison between state-of-the-art quantum-dynamical calculations and detailed experimental data has been possible (Balucani et al. 2002, 2006). This work has allowed a complete characterization of the reaction mechanism. The only exothermic reactive channel is the one leading to NH in its X 3Σ electronic ground state and atomic hydrogen. Of potential relevance for the atmosphere of Titan is the product energy partitioning, that has been determined in the studies of Umemoto et al. (2000) and Balucani et al. (2006), which show that NH vibrational states are populated up to v = 3 (Umemoto et al. 2000). Let us note that the vibrational excitation of NH can significantly change its capability to undergo subsequent reactions. The determination of the product energy release can also provide important information to establish whether H atoms are formed with enough translational energy to escape the atmosphere of the satellite.

3.3.2.4. Reaction N (2P) + H2

Kinetics:

Recommended value at 300 K: k = 1.9 × 10−15 cm3 s−1 (±60%)

No new study has been performed since 1999, so the recommended value for the rate constant is the same as the one proposed by Herron (1999; see Table 3). It is much lower than for the N (2D) state and the outcome of the reaction is probably not the production of NH, but quenching giving N (2D) (or eventually N(4S) atoms), as for the reaction with N2. Suzuki et al. (1993b) measured the temperature dependence of the rate constant as 3.5 × 10−13 exp(−950/T) cm3 s−1 over the 213–300 K temperature range.

3.3.2.5. Reaction N (2D) + CH4

Kinetics:

Recommended values at 300 K: k = 4.0 × 10−12 cm3 s−1 (±40%) for CH4

k = 2.6 × 10−12 cm3 s−1 (±40%) for CD4.

Since the collisional deactivation of N (2D) by N2 is a slow process, the reaction of N (2D) with methane has been soon recognized as an important pathway. The rate constant at 298 K recommended by Herron (4.0 × 10−12 cm3 s−1) is the average of the values determined by Fell et al. (1981), Umemoto et al. (1998b) and Takayanagi et al. (1999) (see Table 3). The data by Black et al. (1969) were disregarded as they differ too much with respect to the other measurements. Umemoto et al. (1998b) and Takayanagi et al. (1999) also studied the reaction with CD4, measuring at 298 K slightly lower rate constants for CD4 than for CH4. Takayanagi et al. (1999) also measured the temperature dependence of the rate constant over the range 223–292 K as k = 7.1 × 10−11 exp (−750/T) cm3 s−1 and 3.3 × 10−11 exp (−700/T) for the reactions with CH4 and CD4, respectively.

Products:

Recommended yields: (CH2NH and/or CH3N + H)/(NH + CH3), (0.8 ± 0.2/(0.2 ± 0.1) for CH4

(CD2ND and/or CD3N + D)/(ND + CD3), (0.8 ± 0.2/(0.2 ± 0.1) for CD4.

In addition to kinetic data, much work has been done on this system at the level of reaction dynamics experiments and theoretical calculations in recent years. The first hint of the reaction mechanism and primary products came from the ab initio calculations by Kurosaki et al. (1998) of the relevant potential energy surface. According to those calculations, the exothermic channels are:

Equation (3d)

Equation (3e)

Equation (3f)

Equation (3g)

Equation (3h)

All of these products can be formed after N (2D) has inserted into one of the C-H bonds of methane forming a first intermediate, CH3NH, which is bound by about 4.43 eV, with respect to the reactants. The insertion mechanism, through a small barrier, has been confirmed by recent ab initio calculations (Ouk et al. 2011). The CH3NH intermediate can directly dissociate into the products CH2NH + H (3d), CH3N + H (3e), and NH + CH3 (3f), or can rearrange to the isomeric form CH2NH2, which is the global minimum of the PES (4.68 eV lower in energy with respect to the reactants). CH2NH2 can also fragment and the possible products are essentially CHNH2 + H (3f), and again CH2NH + H. A first spectroscopic dynamical study by Umemoto et al. (1997) identified NH (from channel (3g)) and H (from channels (3d–f)) as primary reaction products with an absolute yield of 0.2 ± 0.1 and 0.8 ± 0.2, respectively. Later these authors measured the vibrational population of NH produced in this reaction by a laser induced fluorescence technique. The vibrational distribution was determined to be 10.0 (v = 0); 8.0 ± 1.0 (v = 1); 5.0 ± 0.7 (v = 2); 2.5 ± 0.5 (v = 3), in very good agreement with the theoretical calculations of Pederson et al. (1999) and Honvault & Launay (1999). Casavecchia et al. (Balucani et al. 2009; Casavecchia et al. 2001) have used the crossed molecular beam technique with mass spectrometric detection to investigate which of the three possible CH3N isomers (CH2NH, CHNH2 and CH3N) is actually produced in conjunction with H. According to the experimental results obtained in a wide range of ECM collision energies (from about 0.21 to 0.62 eV, it was established that two isomers are actually formed: methylene-imine (CH2NH) and methyl-nitrene (CH3N). Notably, the relative branching ratio for the formation of CH2NH with respect to CH3N isomers was found to change considerably with ECM, from 0.6 at ECM ∼ 0.21 to 0.04 at ∼0.62 eV.

So the assumption that only NH + CH3 would be the products of the N (2D) + CH4 reaction is not correct. This reactive channel accounts only for about 20% of the products in the room temperature laboratory experiment and we can assume that the product branching ratios are about the same for the reaction with CD4. By analyzing the trend of the branching ratio as a function of the available energy (Balucani et al. 2009), we can presume that the yield of NH + CH3 will be even smaller under the low temperature conditions of Titan's atmosphere. Therefore, the nitrogen chemistry that relies on the dominance of the CH3 + NH channel in the photochemical models of the atmosphere of Titan should be reconsidered. Furthermore, the crossed molecular beam results suggest that the reaction of N (2D) with CH4 is an active route for formation of methylene-imine (CH2NH), a closed-shell molecule containing a new unsaturated CN bond. This reaction demonstrates that C-N bonds can be generated directly by a reaction involving an active form of N2. An indirect identification of CH2NH in the atmosphere of Titan has indeed been reported by Vuitton et al. (2006, 2007) and Yelle et al. (2010).

3.3.2.6. Reaction N (2P) + CH4

Kinetics:

Recommended values at 300 K: k = 8.5 × 10−14 cm3 s−1 (±25%) for CH4

k = 6.0 × 10−14 cm3 s−1 (±25%) for CD4

No new studies have been performed since 1999, so the recommended value for the rate constant is the same as the one proposed by Herron (1999; see Table 3). As for N(2D) reaction, Takayanagi et al. (1999) measured a rate constant for the reaction with CD4, which was found to be slightly lower than for CH4. Both Umemoto et al. (1985) and Takayanagi et al. (1999) suppose that the main reaction channel is quenching of N (2P), most probably into N (2D) state. Takayanagi et al. (1999) also measured the temperature dependence of the rate constant over the range 223–292 K as k = 5.0 × 10−13 exp (−490/T) cm3 s−1 and 3.1 × 10−13 exp (−480/T) for the reactions with CH4 and CD4, respectively.

3.3.2.7. Reaction N (2D) + C2H2

Kinetics:

Recommended values at 300 K: k = 6.5 × 10−11 cm3 s−1 (±25%) for C2H2

k = 6.25 × 10−11 cm3 s−1 (±25%) for C2D2.

The rate constant has been measured by two groups (Fell et al. 1981; Takayanagi et al. 1998b; see Table 3) measured the rate constants over the temperature range 223–293 K and reported values of k = 1.6 × 10−10 exp (−270/T) cm3 s−1 and 1.4 × 10−10 exp(−240/T) for C2H2 and C2D2, respectively. So the H/D isotope effect was found to be very small. The recommended rate constant value of 6.5 × 10−11 cm3 s−1 for C2H2 is the same as the one proposed by Herron (1999), based on the results from Takayanagi et al. (1998b). It is quite significantly larger than the value postulated by Yung (1987).

Products:

Recommended yields at 300 K (and at 150 K): (HCCN + H)/(cyclic-HCCN + H), (0.9 ± 0.1)/(0.1 ± 0.05)

Ab initio calculations of the potential energy surface have been reported (Balucani et al. 2000a), according to which the initial approach is N (2D) addition to the unsaturated π-system of C2H2, which leads to the formation of a three-member cyclic HC(N)CH intermediate. HC(N)CH can then either decompose to cyclic-HC(N)C + H or isomerize to HCCNH and/or H2CCN (cyanomethyl). Both HCCNH and H2CCN can dissociate into HCCN + H. Crossed beam experiments have been performed by Balucani et al. (2000a) and the results are consistent with both HCCN + H and cyclic-HC(N)C + H formation. Unfortunately, the difference in the energetics of the two channels does not allow a clear discrimination between these products, but some arguments are in favor of HCCN formation. RRKM estimates (Balucani et al. 2000a) based on the ab initio potential energy surface calculations confirmed that the channels leading to HCCN + H and cyclic-HCCN + H are the main pathways, with branching ratios σ(HCCN + H)/σ (cyclic-HCCN + H) estimated to be 86:14 and 77:23 at ECM = 0.135 and 0.42 eV, respectively. By extrapolating this variation, we thus recommend a branching ratio of (0.9 ± 0.1)/(0.1 ± 0.05) at 300 K, as well as at 150 K.

3.3.2.8. Reaction N (2P) + C2H2

Kinetics:

Recommended values at 300 K: k = 2.3 × 10−11 cm3 s−1 (±40%) for C2H2

k = 2.0 × 10−11 cm3 s−1 (±40%) for C2D2.

No new measurement has been performed since 1999, so the recommended rate constant value is the same as the one proposed by Herron (1999; see Table 3). Takayanagi et al. (1998b) measured the rate constants over the temperature range 223–293 K and reported values of k = 1.0 × 10−10 exp (−440/T) cm3 s−1 and 7.1 × 10−11 exp (−380/T) for C2H2 and C2D2, respectively. The products of this reaction have not been measured, but the measured rate constant, which is much larger than for the reactions of N (2P) with H2 or CH4, suggests that the main process is chemical reaction and not N (2P) quenching. However, the products are hard to predict without an experimental or theoretical study.

3.3.2.9. Reaction N (2D) + C2H4

Kinetics:

Recommended values at 300 K: k = 4.3 × 10−11 cm3 s−1 (±25%) for C2H4

k = 3.8 × 10−11 cm3 s−1 (±25%) for C2D4.

The room temperature rate constant recommended by Herron is 4.3 × 10−11 cm3 s−1, in excellent agreement with the more recent determination of Sato et al. (1999) who reported a value of 4.1 × 10−11 cm3 s−1 at T = 292 K (see Table 3). Sato et al. (1999) also measured the temperature dependence of the rate constants over the 230–292 K temperature range, as k = (2.3 ± 0.3) × 10−10 exp (−500 ±50/T) cm3 s−1 and (2.4 ± 0.5) ×10−10 exp (−550 ± 50/T) for C2H4 and C2D4, respectively.

Products:

Recommended yields:

(CH2NCH + H)/(c-CH2(N)CH + H)/(CH2CNH + H)/(HCN/HNC + CH3)/(CH3CN/CH3NC + H)/(CH2NC/CHNCH + H2)/(CH2N/CHNH + CH2), 0.67/0.23/0.05/0.02/0.01/0.01/0.01

Table 4 summarizes the product branching ratios of the N (2D) + C2H4 reaction.

According to early ab initio calculations of the relevant potential energy surface (Takayanagi et al. 1998a), the initial step is the addition of N (2D) to the π bond of the C2H4 molecule with formation of cyclic H2C(N)CH2 which can either decompose directly to 2H-azirine (cyclic-CH2(N)CH) and H (2S) or isomerize (via H-migration) to HC(NH)CH2. The latter intermediate can decompose to 1H-azirine or 2H-azirine and H (2S), or isomerize to CH2-CHNH (2A) which can, in turn, isomerize to CH3CNH (2A') or decompose to ketene-imine, CH2CNH (1A'), and H (2S). CH3CNH (2A') is the only intermediate which correlates with the CH3CN + H products. RRKM calculations can help us to understand, which are the major channels among the above possibilities. According to the estimates at 0 K made by Takayanagi et al. (1998a), the channel leading to (2H-azirine (cyclic-CH2(N)CH) + H) is by far the most dominant, accounting for 84.8% of the total reaction, followed by the channel leading to (ketene-imine (CH2CNH) + H) (13.2%), CH3 + HNC (1.2%) and CH3CN + H (0.8%). The yields of the other energetically allowed reaction channels were found to be negligible. Balucani et al. (2000b) have performed a series of crossed beam experiments at ECM = 0.345 eV. The experimental data are consistent with the formation of both 2H-azirine, as well as other isomers. In that paper, it was pointed out that a large fraction of 2H-azirine is formed with enough internal energy to spontaneously tautomerize to the most stable isomer acetonitrile, CH3CN, even in a collision-free environment. A more recent crossed beam study (Lee et al. 2011), associated with a more complex determination of the potential energy surface, has pointed out that the radical CH2NCH is also formed, as well as the CH2CN radical. In general, the channels corresponding to one H-loss made a contribution 5.7 times smaller than the channels corresponding to the loss of two hydrogen atoms. RRKM calculations on the new potential energy surface, however, kept on predicting the dominance of 2H-azirine over the other products (87%; Lee et al. 2011). Finally, a very recent work has led to the construction of a different and more accurate potential energy surface (Rosi et al. 2012) where a larger basis set was used. RRKM calculations have been conducted at 170 K, typical temperature of Titan's atmosphere, by using the new electronic structure calculations (Balucani et al. 2012). These new results are different from those of Takayanagi et al. (1998a) and Lee et al. (2011), possibly because of the incompleteness of the previous potential energy surfaces (see Table 4). Only the channels with a significant yield have been reported in Table 4. The possibility that the primary products 2H-azirine or CH2NCH isomerize to acetonitrile (CH3CN) has been considered within the RRKM estimates. Differently from what was previously supposed, neither 2H-azirine (cyclic-CH2(N)CH) nor the CH2NCH radicals can significantly rearrange into acetonitrile at 170 K (Balucani et al. 2012). Therefore, the reaction N (2D) + C2H4 cannot be a formation route of acetonitrile (CH3CN), as previously suggested and our recommended product yields for this reaction are those calculated by Balucani et al. (2012) at 170 K. Collision-induced isomerization of 2H-azirine and CH2NCH radical into CH3CN is, therefore, only possible in dense high-temperature environments.

Table 4. Product Branching Ratios of the N (2D) + C2H4 Reaction

CH2NCH + H c-CH2(N)CH + H CH2CNH + H HNC/HCN + CH3 CH3NC/CH3CN + H CH2NC/CH2CN/CHNCH + H2 CH2N/CHNH + CH2 References
... 0.848 0.132 0.012/... .../0.008 .../∼0/... ∼0/... Takayanagi et al. 1998b, a
0.0440 0.8691 0.0228 0.0009/0.0039 0.0418/0.0003 0.0021/∼0/0.0117 0.0034/... Lee et al. 2011, b
0.675 0.230 0.052 0.0118/0.0022 0.0111/0.0006 0.008/∼0/... 0.0049/0.0017 Balucani et al. 2012, c

Notes. aCalculated yields at 0 K. bCalculated yields at 0.22 eV collision energy (equivalent to 2500 K). cCalculated yields at 170 K.

Download table as:  ASCIITypeset image

The most probable fate in Titan's atmosphere of the highly strained cyclic molecule 2H-azirine, as well as of the CH2NCH radical, is further reaction with other molecular species present in the upper atmosphere of Titan. These molecular species could then contribute to the formation of N-rich organic macromolecules of Titan's atmosphere.

3.3.2.10. Reaction N (2P) + C2H4

Kinetics:

Recommended values at 300 K: k = 3.0 × 10−11 cm3 s−1 (±25%) for C2H4

k = 3.0 × 10−11 cm3 s−1 (±25%) for C2D4.

Since the review of Herron (1999), Sato et al. (1999) published an experimental study of this reaction with both C2H4 and C2D4, showing no isotope effect (see Table 3). These authors measured the rate constant as a function of temperature in the 230–292 K range reporting the following temperature dependence: k = (1.4 ± 0.5) × 10−10 exp (−455 ± 90/T) cm3 s−1 and (1.3 ± 0.2) × 10−10 exp (−435 ± 50/T) for C2H4 and C2D4, respectively. Their value at 292 K is in good agreement with previous data, so the recommended rate constant value is the same as the one proposed by Herron (1999). The reaction products have not been identified, but the value of the rate constant suggests that chemical reaction is the dominant channel versus N (2P) quenching (Sato et al. 1999), as for the N (2P) + C2H2 reaction.

3.3.2.11. Reaction N (2D) + C2H6

Kinetics:

Recommended value at 300 K: k = 1.9 × 10−11 cm3 s−1 (±25%)

The kinetic of the N (2D) + C2H6 reaction has been investigated by two groups (Fell et al. 1981; Umemoto et al. 1998b) and the room temperature value recommended by Herron is 1.9 × 10−11 cm3 s−1 (see Table 3). Considering the rate constant value at room temperature and by comparison with the reactions with other hydrocarbons, there is almost no doubt that this reaction has a small barrier (few hundreds of Kelvin). This implies that the rate constant value will be smaller at the temperatures of Titan's atmosphere.

Products:

Recommended yields:

(CH2=NH + CH3)/(CH3CH=NH + H)/(NH + C2H5)/(CH2=CHNH2 + H)/(3CH2=NH* + CH3), 0.79/0.12/0.06/0.02/0.01

Table 5 summarizes the product branching ratios of the N (2D) + C2H6 reaction.

Table 5. Product Branching Ratios of the N (2D) + C2H6 Reaction

Products Branching ratio
CH2=NH + CH3 0.79
CH3CH=NH + H 0.12
NH + C2H5 0.06
CH2=CHNH2 + H 0.02
3CH2=NH* + CH3 0.01

Download table as:  ASCIITypeset image

Balucani et al. (2010) have studied the reaction of N (2D) with ethane, both experimentally at two different collision energies and theoretically by calculating the statistical product branching ratios, at temperatures relevant to Titan's conditions. The main reaction channels involve the C–C bond breaking giving CH2NH + CH3 and H atom transfer giving NH + C2H5. At all energies, the dominant channel is dissociation into CH2NH + CH3 and its branching ratio only varies from 0.79 at the lowest energy (ECM = 0.008 eV) to 0.74 at the highest one (ECM = 0.325 eV). The recommended yields are the values calculated at a collision energy of ECM = 0.015 eV, corresponding to a temperature of 175 K, which is characteristic for Titan's strastosphere.

3.3.2.12. Reaction N (2P) + C2H6

Kinetics:

Recommended value at 300 K: k = 5.4 × 10−13 cm3 s−1 (±25%)

The only experimental study of this reaction was performed by Umemoto et al. (1985) who measured a small rate constant of 5.4 × 10−13 cm3 s−1, suggesting that the main channel is quenching of N (2P) and not any reactive process, as for the reaction with methane. So N (2P) seems to be mainly quenched in collisions with saturated hydrocarbons, whereas the reaction of N (2P) with unsaturated hydrocarbons, such as C2H2 and C2H4, leads to chemical reactions.

3.3.2.13. Reaction N (2D) + C3H8

Kinetics:

Recommended value at 300 K: k = 2.9 × 10−11 cm3 s−1 (±25%)

The kinetics of the N (2D) + C3H8 reaction has been investigated by the same two groups mentioned for the reaction with C2H6 (Fell et al. 1981; Umemoto et al. 1998b; see Table 3). The room temperature value recommended by Herron is 2.9 × 10−11 cm3 s−1. Considering the rate constant value at room temperature and by comparison with the reactions with other hydrocarbons, there is almost no doubt that this reaction has a small barrier (a few hundreds of Kelvin). This implies that the rate constant value will be smaller at the temperatures of Titan's atmosphere.

No information is available on the nature of the products and we can expect that the favored mechanism will proceed through N (2D) insertion into one of the C–H bonds, maybe giving CH2NH + C2H5 or CH3CHNH + CH3, by analogy with the N insertion into C2H6. It could also produce NH associated with C3H7 or C2H5CHNH + H. But it would be too speculative to propose branching ratio values without theoretical calculations.

3.3.2.14. Reaction N (2P) + C3H8

Kinetics:

Recommended value at 300 K: k = 1.9 × 10−12 cm3 s−1 (±25%)

The only experimental study of this reaction was done by Umemoto et al. (1985) who measured a rate constant equal to 1.9 × 10−12 cm3 s−1. There is no information about the products. The measured rate constant is higher than for the other CH4 and C2H6 saturated hydrocarbons, but lower than those for the unsaturated C2H2 and C2H4 hydrocarbons. Hence, it is hard to deduce for sure whether this reaction proceeds via quenching of the N (2P) state or by chemical reaction, but quenching is more likely, as for the reactions of N (2P) with the other saturated hydrocarbons, CH4 and C2H6.

4. N+2 SINGLY CHARGED IONS

4.1. Production

Ionization of N2 by photons or electrons produces N+2 ions in the ground state X2Σ+g, as well as in the A2Πu, B2Σ+u, and C2Σ+u valence electronic excited states. For most reactions, only N+2 (X, v) ground-state molecular ions will be considered in this review. The photoionization cross section of N2 has been measured from threshold up to 107 eV by Samson et al. (1987) and partial ionization cross sections into electronic and vibrational states have been measured by Woodruff & Marr (1977). The ionization threshold of N2 is at 15.58 eV, but above this energy, the absorbed photon energy does not only result in the ionization of N2, but also can induce the formation of dissociative neutral Rydberg states. Shaw (1992) measured the absolute absorption cross section and the photoionization quantum efficiency from threshold up to 25.6 eV. The N+2 formation cross section by electron impact has been measured by Tian & Vidal (1998), from threshold up to 600 eV (see the comment of Stebbings & Lindsay (2001) on the accuracy of these cross sections). The N+2 formation cross section is a maximum (∼1.85 × 10−16 cm2) for an electron energy of 80–100 eV (Itikawa 2006). We note that the N+2 quartet state (4Σ+u) can also be produced by electron impact, whereas this process is forbidden by photoionization. Ionization of N2 molecules produces N+2 in different vibrational levels. The N+2 (X 2Σ+g) state is produced mainly in the ground vibrational level by direct ionization (Potts & Williams 1974). But vibrationally excited N+2 (X, v = 1–4) states are also populated, both by autoionization in the case of photoionization and by radiative cascades from excited electronic states. Their relative population is difficult to estimate because it strongly depends on the photon or electron energy. Partial cross sections into individual vibrational levels of the N+2 ground state have been measured by Woodruff & Marr (1977) for photon energies between 16 and 40 eV.

4.2. Lifetimes and Quenching

The A2Πu and B2Σ+u excited electronic states of N+2 decay by radiative emission to the ground state. The A2Πu state has a lifetime which decreases from 13.9 μs to 7.3 μs when its vibrational energy increases (Peterson & Moseley 1973) and the lifetime of the B2Σ+u state is equal to 67 ns (Wuerker et al. 1988). In Titan's ionosphere, these two states have a lifetime which is too short compared to the time between collisions (∼1 s at the peak of the ionosphere) for their reactivity to be significant. The N2 dissociative ionization threshold is at 24.3 eV, in the N+2 (C2Σ+u) state region for vibrational levels v ⩾ 3. The radiative lifetime of N+2 (C2Σ+u, v ⩽ 2) levels is equal to 0.09 μs and about 5 × 10−9 s for predissociated vibrational levels v ⩾ 4 (Fournier et al. 1971). N+2 ions in the (C2Σu, v ⩾ 3) states and in the higher inner valence states dissociate into N+ + N. Thus relevant reactions in Titan's atmosphere are only due to N+2 ions in their (X 2Σ+g) ground state. The excited vibrational states of N+2 (X, v = 1–4) can only relax by collisional quenching. Böhringer et al. (1983) predict that vibrational quenching in planetary atmospheres will generally be efficient, k > 1 × 10−12 cm3 s−1 for almost all ions and neutral gases.

4.3. Reactions

Anicich & McEwan made several compilations of all ion–molecule reactions which may play a role in Titan's atmospheric chemistry (Anicich 1993; Anicich & McEwan 1997; McEwan & Anicich 2007; McEwan et al. 1998). They give the reaction rate constants at 300 K, and most of these reactions have only been studied at room temperature. In this review, we have updated the most recent of those compilations (McEwan & Anicich 2007) for N+2 reactions with the main neutral constituents of Titan's atmosphere, adding a discussion of the reaction mechanisms, as well as information about isotope effects and their dependence with collision and internal energy.

4.3.1. Reaction N+2 + N2

Kinetics

Recommended value at 300 K: k = 5.0 × 10−10 cm3 s−1 (±25%)

The different rate constant data for the N+2 + N2 reaction can be found in Table 6, which summarizes all published rate constant data for N+2, N+ (3P, 1D) and N++ reactions. The second and fourth columns indicate the temperature or collision energy and the experimental method used, respectively.

Table 6. Rate Constant Measurements for the Ion N+2, N+(3P,1S) and N++ Reactions

Reaction T (K) or k Method Reference
ECM(eV) (cm3 s−1)
N+2 + N2 300 K 6.6 × 10−10 (±8%) ICR (McMahon et al. 1976)
N+2 + N2 300 K 5.0 × 10−10 (±25%) SIFT (Adams & Smith 1981)
N+2 + N2 300 K 4.24 × 10−10 (±20%) for v = 0 SIFT (Frost et al. 1994)
    5.32 × 10−10 (±20%) for v = 1    
    5.13 × 10−10 (±20%) for v = 2    
N+2 + N2 300 K 4.0 × 10−10 (±20%) for v = 0 SIFT (Kato et al. 1998)
    6.3 × 10−10 (±30%) for v = 1    
    5.8 × 10−10 (±30%) for v = 2    
    6.4 × 10−10 (±30%) for v = 3    
    5.5 × 10−10 (±30%) for v = 4    
N+2 + N2 300 K 5.80 × 10−10 (±25%) Compilation (Anicich & McEwan 1997)
N+2 + H2 300 K 1.8 × 10−9 (±30%) SIFT (Tichy et al. 1979)
N+2 + H2 300 K 2.1 × 10−9 (±20%) SIFT (Smith et al. 1978)
N+2 + H2 < 200 K 1–1.2 × 10−9 CRESU (Rowe et al. 1989)
N+2 + H2 < 15 K 0.8–2.0 × 10−9 FJFR (Randeniya & Smith 1991b)
N+2 + H2 300 K 2.0 × 10−9 for H2 SIFDT (Schwarzer et al. 1991)
    1.5 × 10−9 for D2    
N+2 + H2 300 K 1.7 × 10−9 (±30%)a GIB (Tosi et al. 1992)
N+2 + H2 300 K 1.72 × 10−9 (±30%) for H2 GIB (Schultz & Armentrout 1992)
    1.27 × 10−9 (±30%) for HD    
    1.15 × 10−9 (±30%) for D2    
N+2 + H2 300 K 1.71 × 10−9 (±5%) for (X, v = 0) SIFT-LIF (de Gouw et al. 1995)
    1.69 × 10−9 (±3%) for (X, v = 1)    
    1.80 × 10−9 (±6%) for (X, v = 2)    
    1.68 × 10−9 (±3%) for (X, v = 3)    
    1.47 × 10−9 (±5%) for (X, v = 4)    
N+2 + H2 300 K 1.1 × 10−9 (±20%) for (X, v = 0–4, J = 2)a GIB (Knott et al. 1995)
N+2 + H2 300 K 2.00 × 10−9 (±15%) for H2 Compilation (McEwan & Anicich 2007)
    1.25 × 10−9 (±20%) for D2    
N+2 + CH4 300 K 1.01 × 10−9 (±17%) Opt (Herod et al. 1971)
N+2 + CH4 300 K 1.28 × 10−9 PIMS (Warneck 1972)
N+2 + CH4 300 K 1.06 × 10−9 (±5%) IonTrap (Li & Harrison 1978)
N+2 + CH4 300 K 1.30 × 10−9 (±30%) SIFT (Tichy et al. 1979)
N+2 + CH4 298 K 9.80 × 10−10 (±10%) ICR (Huntress et al. 1980)
N+2 + CH4 300 K 1.00 × 10−9 (±20%) SIFT (Smith et al. 1978)
N+2 + CH4 70 K 1.20 × 10−9 (±30%) CRESU (Rowe et al. 1989)
N+2 + CH4 8–15 K 1.90 × 10−9 (±50%) FJFR (Randeniya & Smith 1991a)
N+2 + CH4 0.36 eV 1.15 × 10−9 (±20%)a GIB (Nicolas 2002)
N+2 + CH4 0.1–3 eV 1.53–2.83 × 10−9(± 20%)a GIB (Nicolas et al. 2003b)
N+2 + CH4 298 K 1.00 × 10−9 SIFT (Anicich et al. 2004)
N+2 + CH4 300 K 1.14 × 10−9 (±15%) Compilation (McEwan & Anicich 2007)
N+2 + C2H2 300 K 8.80 × 10−10 (±11%) Opt (Dreyer & Perner 1971)
N+2 + C2H2 300 K 4.30 × 10−10 PIMS (Warneck 1972)
N+2 + C2H2 298 K 4.00 × 10−10 (±15%) ICR (McEwan et al. 1998)
N+2 + C2H2 0.08 eV 1.7 × 10−10 (±20%)a GIB (Nicolas 2002)
N+2 + C2H2 298 K 5.50 × 10−10 SIFT (Anicich et al. 2004)
N+2 + C2H2 300 K 4.0 × 10−10(± 15%) Compilation (McEwan & Anicich 2007)
N+2 + C2H4 298 K 7.10 × 10−10 SIFT (Anicich et al. 2004)
N+2 + C2H4 298 K 1.30 × 10−9 (±15%) ICR (McEwan et al. 1998)
N+2 + C2H4 298 K 1.30 × 10−9 (±15%) Compilation (McEwan & Anicich 2007)
N+2 + C2H6 298 K 1.44 × 10−9 (±30%) SIFDT (Praxmarer et al. 1998)
N+2 + C2H6 298 K ∼1.30 × 10−9(±30%) ICR (McEwan et al. 1998)
N+2 + C2H6 0.1 eV 7.4 × 10−10 (±20%)a for C2D6 GIB (Nicolas 2002)
N+2 + C2H6 298 K 1.44 × 10−9 (±30%) ICR (McEwan & Anicich 2007)
N+2 + C3H8 298 K 1.30 × 10−9 (±30%) SIFDT (Praxmarer et al. 1998)
N+ + N2 300 K 1.8 × 10−10(± 40%)b FA (Fehsenfeld et al. 1974)
N+ + N2 300 K 3.3 × 10−10 (±25%)b ICR (Anicich et al. 1977)
N+ + N2 300 K 1.6 × 10−10 (±20%)b, c GIB (Freysinger et al. 1994)
N+ + N2 300 K 1.9 × 10−10(± 20%)b, c GIB-TOF (Glosik et al. 2000)
N+ + H2 300 K 5.6 × 10−10(± 30%) FA (Fehsenfeld et al. 1967)
N+ + H2 298 K 4.8 × 10−10(± 10%) ICR (Kim et al. 1975)
N+ + H2 300 K 6.4 × 10−10(± 20%) SIFT (Adams & Smith 1976)
N+ + H2 300 K 6.2 × 10−10(± 30%) SIFT (Tichy et al. 1979)
N+ + H2 300 K 4.8 × 10−10(± 20%) SIFT (Smith et al. 1978)
N+ + H2 300 K 3.7 × 10−10 (±20%) for H2 SIFT (Adams & Smith 1985)
    3.5 × 10−10 (±20%) for HD    
    1.3 × 10−10 (±20%) for D2    
N+ + H2 67 K 2.3 × 10−10 (±30%) CRESU (Marquette et al. 1985)
N+ + H2 11–20 K 0.3–4 × 10−12 (±40%) Ion Trap (Luine & Dunn 1985)
N+ + H2 300 K 3.9 × 10−10 (±20%) for H2 GIB (Ervin & Armentrout 1987)
    2.7 × 10−10 (±20%) for HD    
    1.7 × 10−10 (±20%) for D2    
N+ + H2 163 K 3.0 × 10−10 (±30%) for p-H2 CRESU (Marquette et al. 1988)
N+ + H2 300 K 3.73 × 10−10 for H2 Calculations (Gerlich 1989)
    1.2 × 10−10 for D2    
N+ + H2 15–100 K 0.2–4 × 10−10 for nH2d GIB (Gerlich 1993)
N+ + H2 300 K 4.16 × 10−10 Calculations (Ge et al. 2006)
N+ + H2 300 K 5.0 × 10−10 (±20%) for H2 Compilation (McEwan & Anicich 2007)
    3.1 × 10−10 (±20%) for HD    
    1.5 × 10−10 (±20%) for D2    
N+ (3P) + CH4 298 K 1.35 × 10−9 (±10%) ICR (Anicich et al. 1977)
N+ (3P) + CH4 300 K 1.1 × 10−9 (±30%) SIFT (Tichy et al. 1979)
N+ (3P) + CH4 300 K 9.4 × 10−10 (±20%) SIFT (Adams et al. 1980)
N+ (3P) + CH4 300 K 1.20 × 10−9 (±10%) DT (Dheandhanoo et al. 1984)
N+ (3P) + CH4 8 K 8.2 × 10−10 (±35%) CRESU (Rowe et al. 1985)
N+ (3P) + CH4 300 K 1.15 × 10−9 (±15%) Compilation (Anicich 1993)
N+ (3P) + CH4 298 K 1.1 × 10−9(± 20%)     SIFT (Anicich et al. 2004)
N+ (3P) + CH4 0.2–2 eV 1.5 × 10−9 (±25%)a GIB (Alcaraz et al. 2004)
    for the 15N+ + CD4 reaction    
N+ (3P) + CH4 300 K 1.15 × 10−9 (±15%) Compilation (McEwan & Anicich 2007)
N+ (1D) + CH4 300 K 1.1 × 10−9 (±15%)e SIFT (Tichy et al. 1979)
N+ (1D) + CH4 0.2–2 eV 1.5 × 10−9 (±25%)a TPEPICO (Alcaraz et al. 2004)
    for the 15N+ + CD4 reaction    
N+ (3P) + C2H2 298 K 1.50 × 10−9 (±15%) ICR (McEwan et al. 1998)
N+ (3P) + C2H2 298 K 1.30 × 10−9 (±20%) SIFT (Anicich et al. 2004)
N+ (3P) + C2H2 298 K 1.50 × 10−9 (±15%) compilation (McEwan & Anicich 2007)
N+ (3P) + C2H4 300 K 1.54 × 10−9 (±20%) SIFT (Rakshit 1980)
N+ (3P) + C2H4 300 K 1.60 × 10−9 (±20%) for C2H4 SIFT (Smith & Adams 1980)
    1.50 × 10−9 (±20%) for C2D4    
N+ (3P) + C2H4 298 K 1.30 × 10−9 (±15%) ICR (McEwan et al. 1998)
N+ (3P) + C2H4 298 K 1.10 × 10−9 (±20%) SIFT (Anicich et al. 2004)
N+ (3P) + C2H4 298 K 1.3 × 10−9 (±15%) Compilation (McEwan & Anicich 2007)
N+ (1D) + C2H4 300 K 3.80 × 10−9 (±30%) SIFT (Rakshit 1980)
N+ (1D) + C2H4 0.2 eV 1.7 × 10−9 (±30%)a GIB (Alcaraz et al., unpublished)
    for the 15N++ C2D4 reaction    
N+++ N2 300 K 2.8 × 10−9 (±20%) Penning trap (Church & Holzscheiter 1980)
N+++ N2 1.8 eV 2.1 × 10−9 (±10%) rf ion trap (Fang & Kwong 1997)
N+++ H2 0.34 eV 3.38 × 10−11 (±10%) rf ion trap (Fang & Kwong 1997)

Notes. aConverted from cross sections by using the expression k = σ〈v〉. bRate constants related to the isotope exchange channel measured with isotopically labelled reactants 15N+ + 14N2 or 14N+ + 15N2. cExtrapolated to 300 K and converted from cross sections by using the expression k = σ〈v〉. dThe range of the rate constant values are given here for normal H2, but there are also values for para-H2 with different ortho-H2 admixtures in the paper of Gerlich(Gerlich 1993). eThis rate constant corresponds to both the reactivity of the excited state and its quenching.

Download table as:  ASCIITypeset images: 1 2 3

Collisions of N+2 ions (v = 0–4) with N2 can lead to vibrational deactivation and charge transfer (CT) processes. The formation of N+3, which is endothermic by 4.97 eV, has been studied from thermal energies up to 25 eV energy by Tosi et al. (2001) and has been interpreted as resulting from reactions of the N+2 (4Σ+u) quartet state. The symmetric charge transfer reaction has been studied using isotopic labeling, i.e., by measuring the 15N+2 (v = 0) + 14N2 or the 14N+2 (v = 0) + 15N2 reaction. This charge transfer is an efficient process with a rate constant equal to about half the Langevin rate constant (8.3 × 10−10 cm3 s−1). It is hard to predict the value at 150 K, so we recommend the same value as at 300 K with a 70% uncertainty in order to include the maximum Langevin rate value. The recommended rate constant value for N+2 (X, v = 0) at 300 K is 5.0 × 10−10 cm3 s−1, which is the mean value of the literature values (see Table 6). Frost et al. (1994) and Kato et al. (1998) measured the rate constant as a function of the vibrational energy of N+2. Kato et al. explain the observed increase of the rate constant with vibrational energy (about 6.0 × 10−10 (±30%) cm3 s−1 for v = 1–4, instead of 4.0 × 10−10 (±20%) cm3 s−1 for v = 0), as the result of vibrational relaxation in addition to symmetric charge transfer. Vibration and translation inter-conversion processes can occur coincidently with charge transfer during collisions between N+2 and N2, and several vibrational quanta can efficiently be converted to product kinetic energy during exothermic charge transfer reactions (McAfee et al. 1981; Sohlberg 1999).

4.3.2. Reaction N+2 + H2

Kinetics

Recommended values at 300 K: k = 1.56 × 10−9 cm3 s−1 (±15%) for H2 and 1.15 × 10−9 cm3 s−1 (±20%) for D2

Recommended values at 150 K: k = 1.3 × 10−9 cm3 s−1 (±20%) for H2 and 9.6 × 10−10 cm3 s−1 (±20%) for D2.

This reaction has been extensively studied over the years. All rate constant measurements at 300 K are in very good agreement (see Table 6) and give a value nearly equal to the Langevin rate (1.56 × 10−9 cm3 s−1), which is therefore our recommended value. Rowe et al. (1989) and Randeniya & Smith (1991) have measured the rate constant as a function of temperature below 300 K. When temperature decreases, the rate constant goes through a minimum and then increases to reach about 2 × 10−9 cm3 s−1 at about 8 K (Randeniya & Smith 1991). At 150 K, it is about 1.3 × 10−9 cm3 s−1, which is our recommended value for Titan's ionosphere. Knott et al. (1995) measured the cross section as a function of collision energy and found two regimes: it follows an E−0.5CM law for ECM = 8–90 meV and an E−0.33CM law for ECM = 90 meV to 2 eV. These authors did not observe a variation of the rate constant with the vibrational level of N+2. Similar behavior is observed at higher collision energies (Henri et al. 1988; Koyano et al. 1987). Isotopic effects have been investigated in guided ion beam experiments (Schultz & Armentrout 1992). These authors found that there is no significant isotopic effect on the rate constant for the reactions of N+2 with H2, HD, or D2, which are equal to the Langevin rate within experimental errors. The rate constants decrease from H2 to HD to D2, but the ratio of the measured rate constant divided by the Langevin rate remains equal to about 1. In a selected ion flow drift tube (SIFDT) experiment, Schwarzer et al. (1991) observed the same trend in kinetics for the reactions of N+2 with H2 and D2. Given the above, the recommended rate constant values are the Langevin rate constants at 300 K for both reactions with H2 and D2. At 150 K, the recommended rate constant values are lower. For the reaction with H2, it is the one measured by Randeniya & Smith (1991) and for the reaction with D2, it is a value calculated from the one at 300 K, decreased in the same proportion.

Products

Recommended yields at 300 K: (N2H+ + H)/(H+2 + N2), 0.99/0.01.

Reactive collisions between N+2 and H2 result either in hydrogen transfer to give N2H+ + H or in charge transfer to give H+2 + N2, the former channel being strongly favored at 300 K. The hydrogen transfer giving N2H+ + H products has been the subject of extensive experimental investigations. Three of the most recent papers describe guided ion beam experiments (Knott et al. 1995; Schultz & Armentrout 1992; Tosi et al. 1992), while two other experiments focused on the dependence of the reactivity on the N+2 internal state (electronic and vibrational) (de Gouw et al. 1995; Uiterwaal et al. 1995). For ground-state reactants, the reaction proceeds at thermal energies with the rate predicted by the Langevin classical ion–molecule capture theory (1.56 × 10−9 cm3 s−1). Vibrational excitation of the N+2 appears to have only minor effects on the reaction rates (de Gouw et al. 1995; Knott et al. 1995). As far as the electronic excitation is concerned, the reaction rate for N+2 in the first electronically excited A2Πu state is 50% lower with respect to the X state (Uiterwaal et al. 1995). This difference in rate is not relevant for Titan, as the A state has a lifetime which is much shorter than the average collision time in Titan's ionosphere. When HD is used as reactant, the production of N2D+ is slightly favored over N2H+, for collision energies below 0.1 eV cm. In particular, at thermal energies the branching ratio is N2D+:N2H+ = 0.52: 0.48 (Schultz & Armentrout 1992), while the formation of N2H+ is favored at energies above 0.1 eV.

The charge transfer process giving H+2 + N2 products is only a minor product channel (Schultz & Armentrout 1992; Uiterwaal et al. 1995), and has been investigated as a function of the collision energy for ground-state reactants (Schultz & Armentrout 1992), and at thermal energies as a function of the internal state of N+2 (Uiterwaal et al. 1995). The charge transfer cross section is about two orders of magnitude lower than that for the hydrogen abstraction reaction, according to Schultz et al. (Schultz & Armentrout 1992) and one order of magnitude lower, according to Uiterwaal et al. (1995). This small cross section has been explained by the absence of crossing between the potential energy surfaces (Schultz & Armentrout 1992).

4.3.3. Reaction N+2 + CH4

Kinetics:

Recommended values at 300 K: k = 1.18 × 10−9 cm3 s−1 (±15%) for CH4

k = 1.10 × 10−9 cm3 s−1 (±20%) for CD4.

The kinetics of this reaction has been extensively studied (see Table 6). The latest determinations at 300 K are all in good agreement, and are satisfyingly represented by the value (1.14 × 10−9 cm3 s−1 (±15%)) in the compilation of McEwan & Anicich (2007). Moreover, this value is very equal within experimental uncertainties to the Langevin rate (1.18 × 10−9 cm3 s−1), which is reasonable for such a reaction. It is thus our recommended value.

Rowe et al. (1989) and Randeniya & Smith (1991) studied this reaction at low temperatures, 70 and 8–15 K, respectively. In spite of a slight increase of the rate constant at low temperatures, 1.20 × 10−9 at 70 K and 1.90 × 10−9 at 8–15 K, the difference with the value of the rate constant at 300 K is not significant. The rate constant seems effectively unchanged over the temperature range 8–300 K, so the recommended value is the same at 300 K and 150 K. The study of Nicolas et al. (2003b) shows, however, a substantial increase of the rate constant with collision energy, and in the thesis report of Nicolas (2002), a variation of less than 20% in the rate constant is observed with vibrational energy (v = 0–4). A small isotopic effect has been observed by Nicolas et al. (2003b), who measured, in the 0.1–3 eV ECM range, rate constants equal to (1.53–2.83) × 10−9 cm3 s−1 and (1.43–1.64) × 10−9 for the reaction with CH4 and CD4, respectively, as expected for the Langevin rates, because the reaction relative mass and the polarization are not the same for CH4 and CD4. The Langevin rate for the reaction with CD4 is 1.10 × 10−9 cm3 s−1, calculated with the polarization of CD4 which is lower by 0.034 than the one from CH4 according to Bell (1942) and Wong et al. (1991). So the recommended rate constant values at 300 K and 150 K for the reaction with CD4 are the Langevin rates, which are independent of temperature.

Products

Recommended yields at 300 K and 150 K:

(CH+3 + H + N2)/(CH+2 + H2 + N2)/(N2H+ + CH3), 0.86/0.09/0.05 for CH4

(CD+3+ D + N2)/(CD+2+ D2 + N2)/(N2D++ CD3), 0.88/0.07/0.05 for CD4.

Table 7 summarizes the product branching ratio data for the N+2 + CH4 reaction.

Table 7. Product Branching Ratio Data for the N+2 + CH4 Reaction

CH+3 + H + N2 CH+2 + H2 + N2 N2H+ + CH3 Reference
0.89 0.11 ... Tichy et al. 1979
0.92 0.08 ... Huntress et al. 1980
0.93 0.07 ... Smith et al. 1978
0.80 0.20 ... Randeniya & Smith 1991
0.91 0.09 ... Anicich 1993
0.80 0.05 0.15 McEwan et al. 1998
0.86 (±2%) for CH4 0.09 (±2%) for CH4 0.05 (±2%) for CH4 Nicolas et al. 2003b, a
0.88 (±2%) for CD4 0.07 (±2%) for CD4 0.05 (±2%) for CD4  
0.88 0.12 ... Anicich et al. 2004
0.91 0.09 ... McEwan & Anicich 2007
0.84 0.09 0.07 C. Alcaraz et al., unpublished
0.83 ± 0.02 0.17 ± 0.02 ... Gichuhi & Suits 2011

Note. aMeasured at ECM = 0.1 eV.

Download table as:  ASCIITypeset image

All studies are in agreement concerning the identity of the dissociative charge transfer products, CH+2 and CH+3, with CH+3 formation being the major channel. The nondissociative charge transfer product CH+4 is not observed (less than 0.2% in the study of Nicolas et al. (2003b) because a large amount of internal energy is transferred to the primary CH+4 product which then dissociates. However, there is an important disagreement concerning the production of N2H+. Only three studies relate its formation (McEwan et al. 1998; Nicolas et al. 2003b; C. Alcaraz et al., unpublished). Note that an older work of Harrison & Myher (Harrison & Myher 1967; Li & Harrison 1978) studied specifically this reaction channel. They found a partial rate constant of 2.3 × 10−10 cm3 s−1 for the production of N2D+ in the reaction with CD4, but underline it only represents 2% maximum of the products. Their results are not directly comparable with other values in the literature, but do confirm the production of this ion. The detection of N2H+ is difficult because there is a mass overlap with the secondary product C2H+5 (from the reaction CH+3 + CH4) and the natural isotope of N+2, 14N15N+. In their work, Nicolas et al. (2003b) studied very precisely, at collision energies between 0.1 and 3 eV, the production of N2H+ and N2D+ in the reactions of N+2 with CH4 and CD4, respectively. Their time of flight analysis allowed them to unambiguously distinguish between N2H+ and C2H+5, and the uncertainty of their determination is very small. The measurements of Nicolas et al. (2003b) were confirmed in unpublished experiments of some of the authors of the present paper (C. Alcaraz et al., unpublished). The measured branching ratios are almost constant with collision energy. So it is reasonable to expect the same values at 300 K. C. Alcaraz et al. (unpublished) also studied the variation of the branching ratios with vibrational energy of N+2 (X2Σ+g, v = 0, 4)). The effect is very weak, 2%–3% for the main product CH+3. Recently, Gichuhi et al. (Gichuhi & Suits 2011) measured the branching ratio of this reaction at 45 ± 5 K. Their mass resolution was not sufficient to distinguish N2H+ from the intense N+2 parent ion peak. But for the two other products, their branching ratio is very close to the measurements made at 300 K, in good agreement with the previous measurements of Randeniya & Smith. (1991) at even lower temperatures (8–15 K). In conclusion, the product branching ratio of this reaction seems to be constant as a function of temperature and vibrational excitation of N+2 parent ions within experimental uncertainties. Our recommended yields are the values from Nicolas et al. (2003b).

4.3.4. Reaction N+2 + C2H2

Kinetics

Recommended value: k = 4.15 × 10−10 cm3 s−1 (±25%)

All published rate constant data are shown in Table 6. The study of Dreyer et al. (Dreyer & Perner 1971) involves quite particular conditions, such as a high pressure of N2 (about 40 Torr, and up to 160 Torr). The results of Anicich et al. (2004) have also been obtained at rather high pressure in order to detect the products of multiple successive reactions and not only the primary products. We thus recommend an average value (k = 4.15 × 10−10 cm3 s−1)) for the two other experimental determinations at ∼300 K (k = 4.15 × 10−10 cm3 s−1). This value is significantly smaller than the Langevin rate (1.16 × 10−9 cm3 s−1). It can be explained by the fact that it is mainly a nonresonant charge transfer reaction (see below). Nicolas et al. (Nicolas 2002) measured a lower rate constant at ECM = 0.08 eV (928 K), which indicates that most probably the rate constant decreases with temperature, as expected such a nonresonant charge transfer. This decrease is compatible with the typical variation of the cross section with collision energy for such reactions, i.e., proportional to E−1/2CM. Using this expression, the reaction cross section value extrapolated down to 150 K is 39.6 Å2, giving a rate constant of 1.03 × 10−9 cm3 s−1 at 150 K. This should be checked experimentally, but until then, we recommend a rate constant value of 1.0 × 10−9 cm3 s−1 at 150 K, with an uncertainty of 50% which includes the value at 300 K.

Products

Recommended yields: (C2H+2 + N2)/(N2H+ + C2H), 0.94/0.06.

Table 8 summarizes the product branching ratio data of the N+2 + C2H2 reaction.

Table 8. Product Branching Ratio Data for the N+2 + C2H2 Reaction

C2H+2 + N2 HCN+ + HCN N2H+ + C2H Reference
0.94 ... 0.06 Nicolas 2002, a
1 ... ... Anicich et al. 2004
0.37 0.03 0.60 McEwan et al. 1998
1 ... ... Warneck 1972
1 ... ... Gichuhi & Suits 2011, b

Notes. aMeasured at ECM = 0.08 eV. bMeasured at T = 45 ± 5 K.

Download table as:  ASCIITypeset image

C2H+2 is the only significant product observed by Nicolas (2002), which is in agreement with Anicich et al. (2004), Warneck (1972) and with the recent work at low temperature (45 ± 5 K) of Gichuhi et al (Gichuhi & Suits 2011). The disagreement with the study of McEwan et al. (1998) is assumed by Anicich et al. (2004) to be due to the presence of impurities in the instrument, so these results are disregarded. The erroneous old branching values are still present in the compilation of McEwan & Anicich (2007), probably due to an error. The N2H+ product observed by Nicolas (2002) is a very minor product and was probably not detected in other studies, due to experimental difficulties. In conclusion, we recommend the branching ratios measured by Nicolas (2002). They should be nearly identical at 150 K.

4.3.5. Reaction N+2 + C2H4

Kinetics

Recommended value at 300 K: k = 1.30 × 10−9 cm3 s−1 (±15%)

The disagreement between the determinations of Anicich et al. (2004, 7.10 × 10−10 cm3 s−1) and McEwan et al. (1998, 1.30 × 10−9 cm3 s−1) which are the only published rate constant data for this reaction, is not explained by the authors (see Table 6). However, the experimental conditions are substantially different, with a higher pressure in the case of Anicich et al. (2004) inducing secondary reactions. The inherent uncertainty in this higher pressure procedure being more significant, we recommend the rate constant determined by McEwan et al. (1998), as in the recent compilation of McEwan & Anicich (2007). This value is equal to the Langevin rate (1.29 × 10−9 cm3 s−1) within experimental errors. As the Langevin rate does not depend on temperature, the value will be the same at 150 K.

Products

Recommended yields: (C2H+3 + H + N2)/(C2H+2 + H2 + N2)/(N2H+ + C2H3), 0.67/0.23/0.10.

Table 9 summarizes the product branching ratio data for the N+2 + C2H4 reaction.

Table 9. Product Branching Ratio Data for the N+2 + C2H4 Reaction

C2H+3 + H + N2 C2H+2 + H2 + N2 HNC+ + ? HCNH+ +? N2H+ + C2H3 Reference
0.50 0.20 0.10 0.10 0.10 McEwan et al. 1998
0.64 0.36 ... ... ... Anicich et al. 2004
0.74 ± 0.02 0.26 ± 0.02 ... ... ... Gichuhi & Suits 2011, a

Note. aMeasured at T = 45 ± 5 K.

Download table as:  ASCIITypeset image

In the study of McEwan et al. (1998), C2D4 was used in addition to C2H4. They could conclude that the direct charge transfer product fragmented into C2H+3 + H and C2H+2 + H2, as for the study of Gichuhi et al. (Gichuhi & Suits 2011) at 45 ± 5 K. The chemical products HCNH+ and HCN+, observed by McEwan et al. (1998), which would result from an N2 triple bond breaking, are, however, very unlikely to be produced in this reaction, which is confirmed by the results of Anicich et al. (2004), who did not detected them. The production of N2H+, observed by McEwan et al. (1998) was not confirmed in the work of Anicich et al. (2004). Nevertheless, the production of this ion cannot be definitively excluded. That is, N2H+ signals could have been masked in the work of Anicich et al. (2004), as a mass overlap with 14N15N+ complicates this ion's detection. Gichuhi et al. (Gichuhi & Suits 2011) could also not have detected N2H+ product ions, due to a lack of mass resolution. We thus recommend the branching ratios measured at low temperature by Gichuhi et al. (Gichuhi & Suits 2011), normalized to 100%, after adding 10% of N2H+ ions, i.e., C2H+3/C2H+2/N2H+ yields being equal to 0.67/0.23/0.10.

4.3.6. Reaction N+2 + C2H6

Kinetics

Recommended value at 300 K: k = 1.30 × 10−9 cm3 s−1 (±30%) with C2H6

k = 1.25 × 10−9 cm3 s−1 (±30%) with C2D6.

The three different rate constant measurements made at 298 K for this reaction (see Table 6) are in good agreement and very close to the Langevin rate (1.30 × 10−9 cm3 s−1), so this is our recommended value. As the Langevin rate is independent of temperature, we recommend the same value at 150 K. Nicolas et al. (2002) measured a lower value at ECM = 0.1 eV for the N+2 + C2D6 reaction. It seems to indicate a small isotope effect, as expected for the Langevin rate. So we recommend the Langevin rates for both reactions.

Products

Recommended yields: C2H+5/C2H+4/C2H+3/C2H+2/CH+3/CH+4, 0.14/0.27/0.32/0.18/0.08/0.01.

For clarity, the neutrals associated with the ions are not indicated in this paragraph due to a too high number of product channels. However, they are reported in Table 17 which summarizes our recommended values for all ion–molecule reactions. Table 10 summarizes the product branching ratio data for the N+2 + C2H6 reaction.

Table 10. Product Branching Ratio Data for the N+2 + C2H6 Reaction

C2H+6 C2H+5 C2H+4 C2H+3 C2H+2 CH+3 CH+4 HCNH+ Reference
... 0.15 0.30 0.35 0.20 ... ... ... Praxmarer et al. 1998
Detected Detected Detected Detected ... ... ... Detected McEwan et al. 1998
... 0.17 0.35 0.37 ... 0.10 0.01 ... Nicolas 2002, a

Note. aMeasured at ECM = 0.1 eV for the N+2 + C2D6 reaction.

Download table as:  ASCIITypeset image

The results of Praxmarer et al. (1998) and those of Nicolas (Nicolas 2002) are in good agreement for the products C2H+5, C2H+4, and C2H+3, even though the experiments of Nicolas were not performed at thermal energy (∼0.04 eV), but at ECM = 0.1 eV. CH+3 was not detected in the experiments of Praxmarer et al. (1998) because of their lower sensitivity, and C2H+2 was not detected in those of Nicolas because of a mass overlap between C2D+2 and N+2. So the main product ions are the result of dissociative charge transfer, as expected. Both experiments suggest that the branching ratios do not depend much on collision energy and are not sensitive to an isotope effect when using C2H2 or C2D2. The HCNH+ product detected by McEwan et al. (1998) is very unlikely to be produced directly by the reaction as for the N+2 with C2H4 reaction. In conclusion, we recommend branching ratio values which combine the results from the studies of Praxmarer et al. (1998) and Nicolas (2002), i.e., 0.14/0.27/0.32/0.18/0.08/0.01 for C2H+5/C2H+4/C2H+3/C2H+2/CH+3/CH+4, as well as for the reaction with deuterated ethane. In the experimental study of Nicolas (2002), cross sections were observed to be almost constant with the variation of the photon energy which produces N+2 reactant ions. This indicates that the reaction does not depend much on the vibrational energy of N+2 ions.

4.3.7. Reaction N+2 + C3H8

Kinetics

Recommended value at 300 K: k = 1.30 × 10−9 cm3 s−1 (±30%)

There is only one experimental study of this reaction (Praxmarer et al. 1998). This value is close to the Langevin rate (1.42 × 10−9 cm3 s−1), so it will be about the same value at 150 K.

Products

Recommended yields C3H+5/C2H+5/C2H+4/C2H+3, 0.13/0.30/0.17/0.40.

For clarity, the neutrals associated with the ions are not indicated in this paragraph due to a too high number of product channels. However, they are reported in Table 17 which summarizes our recommended values for all ion–molecule reactions. Products of this reaction have been measured by only one group (Praxmarer et al. 1998). They have observed only dissociative charge transfer products with branching ratios C3H+5/C2H+5/C2H+4/C2H+3 equal to 0.13/0.30/0.17/0.40.

Further studies would be required to confirm this experimental study.

4.3.8. Dissociative Recombination Reaction: N+2 + e

Kinetics

Recommended value: k (N+2, v = 0) = 2.2 × 10−7 (Te/300)−0.39 cm3 s−1 (±25%).

The dissociative recombination of N+2 has been extensively studied, both experimentally (Canosa et al. 1991; Cunningham & Hobson 1972; Geoghegan et al. 1991; Kasner 1967; Kella et al. 1996; Mahdavi et al. 1971; Mehr & Biondi 1969; Mul & McGowan 1979; Noren et al. 1989; Oddone et al. 1997; Peterson et al. 1998; Queffelec et al. 1985; Sheehan & St. Maurice 2004; Zipf 1980) and theoretically (Guberman 1991, 2003, 2012; Guberman et al. 1993). The interest in the reaction arises because of its involvement in Earth's atmosphere photochemistry (Fox & Dalgarno 1985; Torr & Torr 1979), and more generally in the chemistry of terrestrial planet ionospheres (Fox 1992, 1993; Fox et al. 1993). All the previous studies on the N+2 recombination process are summarized in the review papers on ion dissociative recombination reactions by Florescu-Mitchell & Mitchell (Florescu-Mitchell & Mitchell 2006) and more recently by Johnsen (2011). This recombination process is usually very exothermic and therefore tends to be an important source of kinetically and internally excited fragments which contribute to heating. Furthermore, the energetic nascent products may have enough energy to escape the atmosphere. As a consequence, the dissociative recombination of N+2 for example, may contribute to 15N/14N isotope enrichment at Titan (Lammer et al. 2000).

From an experimental point of view, two classes of studies have been carried out: those measuring directly the rate constant k (Canosa et al. 1991; Cunningham & Hobson 1972; Geoghegan et al. 1991; Kasner 1967; Mahdavi et al. 1971; Mehr & Biondi 1969; Zipf 1980) and those which determined cross sections as a function of the relative energy of the N+2 ions and electrons (Kella et al. 1996; Mul & McGowan 1979; Noren et al. 1989; Oddone et al. 1997; Peterson et al. 1998; Sheehan & St. Maurice 2004). The latter were then able to derive a rate constant by averaging the cross section over a Maxwellian energy distribution. Three main aspects had to be considered: the electron temperature (Te) dependence of k, the role of the vibrational state of the ions and finally the state of the nitrogen atoms formed through this process.

Two extensive determinations of the direct electron temperature dependence of k are available in the literature. In 1969, Mehr & Biondi (1969) studied the dissociative recombination of N+2 (v = 0) within the temperature range 300 K ⩽ Te ⩽ 5000 K using an afterglow apparatus, then in 1972, Cunningham & Hobson (1972) studied this process in a reduced temperature range, 700–2700 K, using a shock tube experiment. Both experiments were in good agreement with each other and led to a temperature dependence of k (Te) ∼ Tne (with n = 0.39 or 0.37, respectively). All other rate constant determinations were carried out at Te = 300 K (Canosa et al. 1991; Geoghegan et al. 1991; Mahdavi et al. 1971) or within a reduced temperature range (Kasner 1967). Interestingly, more recent studies from Peterson et al. (1998) and Sheehan & St. Maurice (2004) who measured cross sections using merged beam techniques led to the same electron temperature dependence of k (k(Te) ∼ T−0.39e) for Te < 1200 K (Sheehan & St. Maurice 2004), but for a population of ions in several vibrational states among which the ground state accounted for about 50%. This observation suggests that the vibrational excitation of the ions has probably little influence on the electron temperature dependence below 1200 K. It is worth indicating however that recent merged beam experiments demonstrated that the electron temperature dependence may be non-monotonic. Thus, the authors proposed a power law such as k (Te) ∼ T−0.56e for 1200 K ⩽ Te ⩽ 4000 K indicating that vibrationally excited ions present a steeper temperature dependence within the range 1200–4000 K, which is of interest for the ionosphere of Titan at altitudes greater than 1600 km (Keller et al. 1992; Wahlund et al. 2005). No uncertainty is given with respect to the power-law exponent, but an error of 20% seems reasonable. A theoretical work by Guberman (2003) concluded that the electron temperature dependence is slightly different according to the temperature range. For the ground state v = 0, he proposed three different power laws: T−0.20e for 200 K ⩽ Te ⩽ 400 K; T−0.35e for 700 K ⩽ Te ⩽ 900 K; and T−0.50e for 1800 ⩽ Te ⩽ 2000 K (Guberman 2003; S. L. Guberman 2006, private communication). His set of results however is in good agreement with the global single temperature dependence T−0.39e found experimentally and recommended by Sheehan & St. Maurice (2004) for the v = 0 state. Very recently, Guberman (2012) revisited this reaction including for the first time in calculations both major and minor dissociative routes for the v = 0 state. He was able to derive from his calculation a rate coefficient for the N+2 ground vibrational level in the temperature range 100–3000 K which was well fitted by the following expression: (2.2 ± +0.2−0.4× 10−7 × (Te/300)−0.40 cm3 s−1, in perfect agreement with the previous recommendation for the temperature dependence. Guberman, however, points out in his paper that an improved representation of k (N+2, v = 0) can be obtained by separating the temperature range into two sub-ranges: k (N+2, v = 0) = 2.2 × 10−7 (Te/300)−0.22 cm3 s−1 for 100 K ⩽ Te ⩽ 600 K and k (N+2, v = 0) = 1.1 × 10−7 (Te/1800)−0.51 cm3 s−1 for 600 K ⩽ Te ⩽ 3000 K.

Considering now the absolute rate constant k, analysis of the literature shows that several experiments (Canosa et al. 1991; Cunningham & Hobson 1972; Geoghegan et al. 1991; Mehr & Biondi 1969) dealt with N+2 (v = 0) ions. At 300 K, a rate constant of 2.2 × 10−7 cm3 s−1 matches all these studies (Sheehan & St. Maurice 2004) with a typical 25% uncertainty. Therefore, it seems reasonable to recommend the following expression for the rate constant for the dissociative recombination of the ground state:

Finally, it is also worth mentioning that the dissociative recombination of 15N14N+ (v = 0) has been found to be rather similar to that of 14N+2 (v = 0). The only available work is a quantum-theoretical calculation from Guberman (2003) who indicated that at room temperature the rate constant is 2.6 × 10−7 cm3 s−1 and at 1000 K 1.6 × 10−7 cm3 s−1.

For ions in vibrational excited states, however, there are no experiments that allow clear conclusions, because ions are formed in a mixture of different vibrational states (Peterson et al. 1998; Sheehan & St. Maurice 2004; Zipf 1980). In 1980, Zipf (1980) made a study in which he identified the v = 0, 1, and 2 levels by laser-induced fluorescence and reported rate constants at 300 K, which were increasing with the excitation level. However, it was recognised later that his experiment reflected the effective recombination for N+2 ions with a vibrational temperature near 1500 K. Reanalysis of this work by Bates & Mitchell (1991) led to the qualitative conclusion that, at 300 K, the rate constant for the dissociative recombination of N+2 (v = 1, 2) should be much smaller than for the ground state. From a theoretical point of view, work is presently in progress (Guberman 2012). It is worth mentioning however that in the ionosphere of Titan, vibrational excited N+2 ions should be quenched prior dissociative recombination by collisions with N2. The rate constant for the N+2 (X, v) vibrational relaxation can be estimated to be efficient and equal to a few 10−10 cm3 s−1 (see the above discussion about the N+2 reaction with N2). Although it is about three orders of magnitude less than the dissociative recombination rate constant, the relative abundance of electrons with respect to N2 in the ionosphere of Titan is too small (∼1 ppm) to favor this latter process.

Products

Recommended yields: N (4S)/N (2D)/N (2P), 0.25, 0.7, 0.05 (±25%.)

Table 11 summarizes the product branching ratio data for the N+2 + e recombination reaction.

Table 11. Product Branching Ratio Data for the N+2 + e Recombination Reaction

N (4S) + N (4S) N (2D) + N (4S) N (2P) + N (4S) N (2D) + N (2D) Reference
... ... ... 0.9a Queffelec et al. 1985
... 0.46 ± 0.06 0.08 ± 0.06 0.46 ± 0.06 Kella et al. 1996
... 0.37 ± 0.08 0.11 ± 0.06 0.52 ± 0.04 Peterson et al. 1998
... 0.7 0.03 0.27 Guberman 2003

Note. a0.9 corresponds to the total N (2D) yield, without distinguishing between the N (2D) + N (4S) and N (2D) + N (2D) channels.

Download table as:  ASCIITypeset image

Determination of the products issued from the dissociative recombination of N+2 is also of great interest. Four channels are exothermically accessible:

Equation (4a)

Equation (4b)

Equation (4c)

Equation (4d)

A fifth channel, namely N (2D) + N (2P) (endothermic by 0.13 eV), is only accessible for vibrational excited ions. From an experimental point of view, only three studies have been carried out (Kella et al. 1996; Peterson et al. 1998; Queffelec et al. 1985). The pioneering work by Queffelec et al. (1985) using a flowing afterglow apparatus with Te ∼ 1000 K showed that the yield of N (2D) was higher than 0.85 for N+2(v = 0, 1) ions, whereas the yield of N (4S) atoms was lower than 0.05, indicating that the main channel is clearly N (2D) + N (2D) (∼0.90). More recent experiments using storage rings working at close to 0 eV relative collision energy confirmed that this channel was important but also indicated a significant branching ratio for the channel N (4S) + N (2D) which cannot be reconciled with the previous afterglow study. For ground state 15N14N+, Kella et al. (1996) found branching ratios of 0.46 ± 0.06, 0.08 ± 0.06, and 0.46 ± 0.06 for channels (4b), (4c), and (4d), respectively, whereas Peterson et al. (1998) indicated values of 0.37 ± 0.08, 0.11 ± 0.06, and 0.52 ± 0.04 for 14N+2 ions. In both experiments, channel (4a) was not observed in agreement with a previous theoretical work by Guberman (1991). However, the agreement between theory and experiments remains poor. The more recent study by Guberman (2003) indicated branching ratios of 0.7, 0.03, and 0.27 for channels (4b), (4c), and (4d), respectively. This topic deserves further study to resolve the present discrepancies. However, experiments clearly indicate that channel leading to two N (2D) atoms accounts for at least 50% of the whole process. With the current state of knowledge, we recommend the following percentages for the different electronic states of N atoms: 0.7/0.05/0.25 for N (2D)/N (2P)/N (4S).

It is now interesting to compare the loss rate of N+2 arising from the dissociative recombination process with that from other possible destruction routes. Neutral molecular nitrogen, methane and to a less extent H2, are obviously the main potential competitors because of their predominance in the ionosphere of Titan. However, the only possible loss of N+2 by the reaction with N2 would be via an association reaction (Rowe et al. 1984) which will be much too slow considering the small nitrogen concentration (∼109 cm−3 at ∼1200 km altitude; Waite et al. 2005). On the other hand, the reactions at room temperature of N+2 with CH4 and H2 have been found to be 1.14 × 10−9 and 1.7 × 10−9 cm3 s−1, respectively (see above). These values are close to the Langevin rate constant and therefore it is expected that they will not change significantly with temperature. Although being two orders of magnitude less efficient than the dissociative recombination of N+2, the concentrations of CH4 and H2 are much higher than the electron density. At ∼1200 km for example the abundance of methane (Waite et al. 2005) is about 2 × 107 cm−3, that of H2 (Waite et al. 2005) is 4 × 106 cm−3, whereas the electron density (Wahlund et al. 2005) is typically 2 × 103 cm−3. Considering a rate constant k (N+2 + e) = 1.4 × 10−7 cm3 s−1 at Te = 1000 K corresponding to a typical electron temperature at this altitude, the loss rate for N+2 through the reaction with methane is about 70 times greater than through dissociative recombination process, and the one resulting from the reaction with H2 is 30 times greater. As a conclusion, it appears that the dissociative recombination of N+2 is a minor destruction route for this ion, at the 10−2 level, although it cannot completely be disregarded.

5. N+ SINGLY CHARGED IONS

5.1. Production

In ionospheres, N+ ions are mainly produced by dissociative ionization of N2. The N2 dissociative ionization threshold is at 24.29 eV and the different dissociation onsets are given in the Table 1. The only long-lived metastable atomic ion state which has been observed to arise from photon or electron impact on N2 is the N+ (1D) state, which lies in 1.9 eV above the N+ (3P) ground state (see Table 1).

Samson et al. (1987) measured the N+ production cross section by photo-dissociative ionization of N2, from threshold up to 107 eV photon energy. The maximum cross section is ∼3.3 × 10−18 cm2 at about 47 eV photon energy. Above the double ionization threshold at 42.88 eV, there is a small contribution from N++2 doubly charged ions, which have the same mass-to-charge ratio (m/z) as N+ fragment ions. Tian & Vidal (1998) measured the N2 electron impact ionization cross sections to form N+2, N+, and N++2 ions from threshold up to 600 eV electron energy. They found that the N+ production cross section reaches a maximum value of 6.6 × 10−17 cm2 at 120 eV electron energy (see the review from Itikawa 2006). The N2 dissociative ionization threshold is at 24.29 eV, in the N+2 (C 2Σ+u) inner valence state region. As dissociative ionization occurs only in the region of inner valence states (C, F, E, G, and H bands), it is not a very efficient process and it only represents less than 5% of the total ionization of N2 for photon energies up to 30 eV. For photons energies higher than 44.5 eV, N+ ions are also produced by dissociation of N++2 doubly charged ions. Above 80 eV photon energy, the contribution of multiple-ionization processes to the total ionization is about 4% (Wehlitz 2010). By electron impact, this contribution reaches a maximum of 6% for 150 eV electron energy (Tian & Vidal 1998).

Most N+ ions produced by the N+2 dissociation are in the 3P ground state. Detailed photoionization experiments showed that in a narrow N+2 internal energy range, 60% of N+ ions are in the long-lived 1D excited state. But under ionospheric conditions, it was estimated that globally about 15% of N+ ions are in the 1D state and 85% are in the ground 3P state (Nicolas et al. 2003a). No other excited states have been observed so far to be populated by dissociative photoionization of N2 (Nicolas et al. 2003a), but it cannot be totally excluded that some N+ (1S) metastable state is produced at high internal energy. Dissociative ionization of N2 produces N+ ions which have some kinetic energy. The N+2 → N+ + N process releases a mean value of about 1 eV (Nicolas et al. 2003a), i.e., N+ ions with about 0.5 eV kinetic energy. The N++2 → N+ + N+ process releases much more kinetic energy: due to the Coulomb repulsion, the charge separation releases a large amount of kinetic energy, shared equally between the two N+ partners. This kinetic energy is important as it can potentially modify some reaction rate constants, as well as participate in the process of atmospheric escape. The kinetic energy released in the dissociation has been measured by Brehm & De Frenes (1978) for electron impact and by Besnard et al. (Besnard et al. 1988) for photoionization. The distribution of kinetic energy released can be described by a broad Gaussian, centered at 8 eV, with FWHM of 6 eV. Thermodynamically, Besnard et al. (1988) showed that the dissociation leads systematically to atomic ions in their ground state, which puts a lower limit of 4.8 eV on the kinetic energy release.

5.2. Lifetimes and Quenching

Two excited states of N+ ions are metastable, the N+(1D) and N+(1S) states, which have a lifetime of 258 s and 31.5 s, respectively (Wiese & Fuhr 2007). The quenching of N+ (1D) metastable states by collisions with N2 is negligible due to its spin forbidden nature (Freysinger et al. 1994). So when N+ (1D) is formed in ionospheres, these species will relax by reactions with other gases.

The kinetic energy of N+ ions produced by dissociative ionization of N2 is efficiently relaxed by collisions with N2 (Glosik et al. 2000). However, the efficiency decreases with the relative kinetic energy. From the work of Glosik et al. (2000), we can estimate the kinetic energy relaxation for the two populations of N+ ions. For N+ coming from N+2 dissociation (∼0.5 eV kinetic energy), each collision leads to a loss of 50 ± 10% of their kinetic energy. For the more energetic N+ ions coming from N++2 dissociation, this loss is of 15 ± 7% of their kinetic energy. So, for example, at the peak of Titan's ionosphere, where there are about 50 collisions on average with N2 before N+ ions collide with another gas, the kinetic energy of the N+ ions from single ionization will be relaxed when reacting with other gases but N2. This relaxation will not be complete for N+ ions coming from N++2 dissociation, but these are minor ions compared to the N+ ions coming from N+2 dissociation. It is therefore a rather good approximation to take rate constants at thermal energy in the models of Titan's ionosphere. However, a rigorous treatment should precisely take into account the relaxation of N+ kinetic energy as a function of altitude for N+ ions coming from both N+2 and N++2 dissociation.

5.3. Reactions

Before reviewing the different reactions of N+ atoms, let us note that experimental studies had to face the problem of contamination of the N+ reactant ions by the presence of some N++2 doubly charged molecular ions, which have the same m/z as the N+ atomic ions and which are very reactive. So the results of many older studies had to be reviewed and corrected at a later date.

5.3.1. Reaction N+ + N2

Kinetics

Recommended value at 300 K: k = 2.1 × 10−10 cm3 s−1 (±25%)

When working with isotopically labeled reagents the following three reactive channels can be distinguished in a mass spectrometer:

  •   
    Isotope exchange reaction 15N+ + 14N214N+ (m/z 14) + 15N14N
  •   
    Atom abstraction 15N+ + 14N214N + 15N14N+ (m/z 29)
  •   
    Electron transfer 15N+ + 14N215N + 14N+2 (m/z 28).

All published rate constant data for this reaction are shown in Table 6. They correspond to the isotope exchange N+ + N2 reaction (15N+ + 14N2 or 14N+ + 15N2 reaction giving 14N+ + 15N14N or 15N+ + 15N14N products, respectively). At thermal energies (300 K) the only energetically allowed reaction channel is the isotope exchange process producing 14N+ because the atom abstraction and electron transfer channels, giving 15N14N+ and 14N+2, respectively, are endothermic by 1.043 eV. Therefore, the rate constant at 300 K only concerns the isotope exchange process. The recommended rate constant value is the mean value of the different measurements (2.1 × 10−10 cm3 s−1). It is significantly lower than the Langevin rate (1.02 × 10−9 cm3 s−1). In guided ion beam (GIB) experiments, Freysinger et al. (1994) and Glosik et al. (2000) measured the kinetic energy dependence of this cross section via experiments using isotopically labeled reactants.

Products at 300 K (and 150 K): 14N+ + 14,15N2 isotope exchange channel

Since the isotope exchange channel does not change the species involved in the reaction, one might argue that it does not eventually influence the ion composition. This is not the case as, reactive collision with molecular nitrogen may generate cold N+ product ions, as it has been shown that most of the kinetic energy of the incoming 15N+ is converted into internal excitation of the neutral N2 product (Glosik et al. 2000). According to Freysinger et al. (1994), the rate constant of this reaction does not vary with temperature.

Products at ECM > 2.7 eV: N+2 + N Electron Transfer/Atom Abstraction Channel

These products, which result from both an electron transfer process and an atom abstraction process, can be distinguished by using suitable isotopic labeling (Fehsenfeld et al. 1974; Freysinger et al. 1994; Maier & Murad 1971; Mark & Gerlich 1996). For the scope of the present review, we treat the overall N+2 production as a unique reaction. In the oldest studies (Maier & Murad 1971; Phelps 1991), there were contributions from N++2 and N+ metastable states in the N+ reactant ions, so the apparent threshold was not due to N+ (3P) ground state ions. According to the N+ production mode, Freysinger et al. (1994) and Mark & Gerlich (1996) could measure the reaction of pure N+ (3P) state ions; details are discussed in these papers. GIB investigations have shown that the reaction has an apparent threshold of 2.7 ± 0.2 eV, well above the thermochemical onset of 1.043 eV (Freysinger et al. 1994; Mark & Gerlich 1996). Except very close to threshold, the cross section σ varies with collision energy according to the following law:

where E is the collision energy in eV (Mark & Gerlich 1996).

According to Freysinger et al. (1994) and Mark & Gerlich (1996), this reaction produces mainly N+2 (X) + N (2D) above 3.43 eV, the thermochemical onset for the production of the N (2D) metastable state in this reaction. The effective threshold of the charge transfer reaction is probably too large to make this channel significant under the conditions of Titan's ionosphere. Let us note that this reaction is exothermic by 0.9 eV when N+ reactant ions are in the N+ (1D) metastable state, but the first spin-allowed channel giving N (2D) + N+2 (X) is endothermic by 1.53 eV.

5.3.2. Reaction N+ + H2

Kinetics

Recommended values at 300 K: k = 3.8 × 10−10 cm3 s−1 (±20%) for H2

k = 3.1 × 10−10 cm3 s−1 (±20%) for HD

k = 1.2 × 10−10 cm3 s−1 (±20%) for D2.

This reaction has been extensively studied both experimentally (see Table 6) and theoretically. It is almost a case study, as its reactive channel giving NH+ + H is very slightly endothermic by 17 meV (Gerlich 1989; Tosi et al. 1994) and by 29 meV for the reaction of N+ + D2 giving ND+ + D (Tosi et al. 1994). As a consequence, this reaction is very sensitive to the temperature (or collision energy) and to the rotational temperature of H2. This sensitivity is because the activation energy can be overcome at 300 K, but not at very low temperatures and for para-H2; this last fact is important, for example, in some regions of the interstellar medium. However in the temperature range of Titan's ionosphere (∼150–200 K), the reaction is still efficient (Marquette et al. 1988). The charge transfer channel producing H+2 + N is endothermic by 0.89 eV and is only energetically possible for the reaction of N+ (1D) metastable state, which lies 1.9 eV above the ground state; however, the reaction is spin-forbidden and has never been unambiguously observed. Therefore, the rate constant values in Table 6, which summarizes the rate constant data for the N+ + H2 reaction, are for the H transfer reaction producing NH+ + H.

The recommended values for the rate constant at 300 K have been estimated from the more recent experimental results (Adams & Smith 1985; Ervin & Armentrout 1987) for the reactions of N+ with H2, HD and D2. Experimental values are in very good agreement with theoretical calculations (Ge et al. 2006; Gerlich 1989), within experimental uncertainties.

This reaction has been investigated by a variety of techniques. The most detailed measurements (Gerlich 1993; Marquette et al. 1988; Tosi et al. 1994) have shown that the reaction is strongly influenced by the internal energy of the reactants: kinetic and rotational energies are equivalent in promoting the reaction, while the fine structure energy seems inefficient for surmounting the reaction barrier. These findings explain why the interpretation of low energy measurements is quite difficult, without a full control of the internal states.

Several authors measured the rate constant at very low temperatures (see Gerlich 1993; Luine & Dunn 1985; Marquette et al. 1988 and references therein). In the 40–100 K range, Gerlich (1993) could fit the data with the expression k = 2 × 10−10 exp (−41/T). Below 40 K, the variation becomes k = 1.1 × 10−10 exp (−26/T). At 15 K and for para-H2, when the available energy is not sufficient to overcome the reaction barrier, the rate constant decreases by several orders of magnitude, being equal to 5 × 10−13 cm3 s−1 (Gerlich 1993). The reactions of N+ with H2, HD, and D2 have been studied as a function of translational energy and hydrogen temperature in a guided ion beam mass spectrometer (Sunderlin & Armentrout 1994). These authors found that the cross sections depend significantly upon the rotational temperature of H2 for collision energies below 0.3 e V. The branching ratio for formation of NH+ and ND+ in the reaction with HD at 0.02 eV is 1:3 at 305 K and 1:13 at 105 K. These effects are consistent with rotational energy driving the reaction. Effects due to the presence of the N+ (1D) metastable state have been observed in a previous experiment of the same group (Ervin & Armentrout 1987). However, since the experiment is not state selected, it was only possible to estimate qualitatively the variation of the rate constant, which might be 50% larger than for ground-state N+ ions.

5.3.3. Reaction N+ (3P) + CH4

For this reaction, we will review separately the reaction of N+ (3P) ground state and the reaction of N+ (1D) metastable state, which lies 1.9 eV above the ground state.

5.3.3.1. Reaction N+ (3P) + CH4

Kinetics:

Recommended value at 300 K: k = 1.15 × 10−9 (±20%) cm3 s−1

We recommend for this reaction the mean value of the experimental values measured at 300 K, i.e., 1.15 × 10−9 cm3 s−1 (±20%), as in the compilation of McEwan & Anicich (2007) (see Table 6). This value matches the Langevin rate (1.38 × 10−9 cm3 s−1) within experimental uncertainties. The rate constant for the reaction with CD4 has been measured by Alcaraz et al. (2004) to be, within experimental uncertainties, equal to the Langevin rate (1.30 × 10−9 cm3 s−1, calculated with the polarization of CD4 which is lower by 0.034 than the one from CH4 according to Bell 1942 and Wong et al. 1991), so it is our recommended value.

Rowe et al. (1985) studied this reaction at a low temperature, 8 K, and found a value of the rate constant slightly lower than the value at 300 K. Similarly, Alcaraz et al. (2004) studied the reaction at collision energies higher than 0.2 eV (i.e., T > 1500 K) and found a higher value for the rate constant. These studies seem to confirm an increase of the rate constant with temperature or with collision energy. So we recommend a rate constant value of 1.0 × 10−9 cm3 s−1 at 150 K. In Titan's ionosphere, the N+ ions are thermalized by collisions with N2, so the value at 150 K should be used.

Products:

Recommended yields: (CH+3 + NH)/(CH+4 + N)/(HCNH+ + H2)/(HCN+ + H2 + H), 0.50/0.05/0.35/0.10.

Table 12 summarizes the product branching ratio data for the N+ (3P) + CH4 reaction.

Table 12. Product Branching Ratio Data for the N+ (3P) + CH4 Reaction

Reaction CH+3+ NH CH+4+ N HCNH++ H2 HCN++ H2+ H Reference
N+ (3P) + CH4 0.53 0.04 0.32 0.10 (Anicich et al. 1977)
N+ (3P) + CH4 0.42 0.06 0.38 0.14 (Tichy et al. 1979)
N+ (3P) + CH4 0.51 0.03 0.40 0.06 (Adams et al. 1980)
N+ (3P) + CH4 0.52 0.06 0.33 0.09 (Dheandhanoo et al. 1984)
N+ (3P) + CH4 0.50 0.05 0.36 0.10 (Anicich 1993) (compilation)
N+ (3P) + CH4 0.53 0.05 0.32 0.10 (McEwan et al. 1998)
N+ (3P) + CH4 0.38 0.03 0.44 0.15 (Anicich et al. 2004)
15N+ (3P) + CD4* 0.51 0.11 0.29 0.09 (Alcaraz et al. 2004)
N+ (3P) + CH4 0.50 0.05 0.36 0.10 (McEwan & Anicich 2007) (compilation)
N+ (1D) + CH4 0.20 0.10 0.40 0.30 (Tichy et al. 1979)
15N+ (1D) + CD4* 0.09 0.40 0.33 0.18 (Alcaraz et al. 2004)

Note. *Measured at ECM = 0.36 eV.

Download table as:  ASCIITypeset image

The values of the branching ratios determined by McEwan et al. (1998) are in agreement with the previous studies. The measurements by Alcaraz et al. (2004), at higher collision energy (ECM = 0.36 eV) and with isotopic labeling, give similar results, except for slight differences for CH+4 and HCN+. This study indicates that there is no major isotopic effect for this reaction, so the branching ratios are the same for the reactions with CH4 and CD4. A joint experimental and theoretical study showed, in the case of the reaction with CD4, that both DCND+ and CD2N+ isomers can be formed (Zabka et al. 2010). The most recent determination of reaction ion products by Anicich et al. (2004) gives different branching ratios, especially for CH+3. Note that their measurements were performed under higher pressure conditions in order to detect the products of multiple successive reactions and not only the primary products. The uncertainty on this procedure being more significant, these authors recommended later (McEwan & Anicich 2007) branching ratios close to those measured by McEwan et al. (1998) at low collision energy, i.e., CH+3/CH+4/HCNH+/HCN+, 0.50/0.05/0.36/0.10. These are also our recommended values, slightly adjusted to have a sum equal to 1.

At higher collision energy, it has been shown that the formation of the products CH+3 and CH+4 was substantially favored at ECM = 0.05–0.53 eV (Dheandhanoo et al. 1984), and at ECM = 0.36–2 eV (Alcaraz et al. 2004). At 2 eV, the branching ratios change to 0.76/0.18/0.06/0.0. This can easily be explained, as CH+3 and CH+4 products resulting from charge transfer should be favored when collision energy is increased, compared to the HCNH+ and HCN+ products resulting from a collision complex with rearrangement of the molecular bonding. In contrary, with thermalized N+ ions at 150 K in Titan's ionosphere, one can expect, as suggested by Dheandhanoo et al. (1984), that the reaction N+ + CH4 would favor the products HCN+ and HCNH+. Up to now, no experiment at low temperature has been performed to confirm this tendency.

5.3.3.2. Reaction N+ (1D) +CH4

Kinetics:

Recommended value at 300 K: k = 1.15 × 10−9 cm3 s−1 (±20%)

Tichy et al. (1979) and Alcaraz et al. (2004) measured the rate constant of the N+ (1D) + CH4 reaction (see Table 6). Both studies showed that the rate constant values are similar to their respective values found for the reaction of the N+ (3P) ground state and close to the Langevin rate within experimental uncertainties. So it seems that the rate constant for this reaction does not change with the electronic excitation of N+ and we recommend the same rate constant value as for the N+ (3P) state. Note that in the study of Alcaraz et al. (2004), the metastable state of N+ (1D) is strictly controlled and corresponds to 100% of the N+ reactant ions. In the case of Tichy et al. (1979), the quantity of this metastable state (about 30%) is not directly measured, but only deduced. Moreover, the metastable state is not clearly identified, Tichy et al. (1979) suggesting a possible contribution from N+ (1S).

Products:

Recommended yields: (CH+3+ NH)/(CH+4 + N)/(HCNH+ + H2)/(HCN+ + H2 + H), 0.09/0.40/0.33/0.18.

As previously noted, a pure sample of the N+(1D) metastable state being generated and identified in the study of Alcaraz et al. (2004), we recommend the product branching ratios from that work (see Table 12).

In comparison with the reactivity of N+ (3P), the ratio (CD+3 + CD+4)/(DCN+ + DCND+) decreases from 0.62/0.38 to 0.49/0.51, corresponding to a slight increase in the yield of the nitrogen-containing compounds. The strongest effect is a complete inversion of the ratio CD+3/CD+4, from 4.4/1.0 to 1.0/4.7. The present Titan ionospheric models do not take into account the N+ (1D) metastable state. Such an inversion in the reaction branching ratios could affect some model predictions and it would be worthwhile to evaluate precisely this effect.

5.3.4. Reaction N+ (3P) + C2H2

Kinetics:

Recommended value at 300 K: k = 1.42 × 109 (±20%) cm3 s1

This rate constant value measured in two different studies (Anicich et al. 2004; McEwan et al. 1998; see Table 6) is very close to the Langevin rate (1.42 × 10−9 cm3 s−1), as for the reaction of N+ with methane, so the recommended value is the Langevin rate. By analogy with the reactions with methane and ethylene (see below), it is reasonable to estimate that the reaction of N+ (1D) occurs also at the Langevin rate and are the same at 300 K and 150 K.

Products:

Recommended yields: (C2H+2 + N)/(CNC+ + H2)/(CHCN+ + H), 0.70/0.15/0.15.

Table 13 summarizes the product branching ratio data for the N+ (3P) + C2H2 reaction.

Table 13. Product Branching Ratio Data for the N+ (3P) + C2H2 Reaction

C2H+2 + N HCNH+ + C CNC+ + H2 CHCN+ + H Reference
0.70 ... 0.15 0.15 McEwan et al. 1998
0.65 0.07 0.06 0.22 Anicich et al. 2004
0.70 ... 0.15 0.15 McEwan & Ancich 2007
        (compilation)

Download table as:  ASCIITypeset image

For this reaction, the same authors made measurements of the branching ratios using two different techniques, the Ion cyclotron resonance (ICR) technique (McEwan et al. 1998) and later the selected ion flow drift tube (SIFT) technique (Anicich et al. 2004). Anicich et al. (2004) identified a new product ion, HCNH+, but with the high pressure conditions of this experiment, there are more secondary reactions and therefore larger uncertainties on the resulting experimental outputs. As for the reaction of N+ with methane (see above), McEwan et al. (McEwan & Anicich 2007) recommended later their oldest product yields, which are also our recommended values. At 150 K, the yield of C2H+2 ions coming from charge transfer might be a little smaller than for the other products coming from a chemical reaction which require a long-lived complex intermediate, but it should be checked. No study exists for the reaction of N+ (1D) metastable state with acetylene, but it is reasonable to estimate that the charge transfer products dominate, as for the reactions of N+ (1D) with methane and ethylene.

5.3.5. Reaction N+ + C2H4

For this reaction, we will separately review the reaction of the N+ (3P) ground state and the reaction of the N+ (1D) metastable state.

5.3.5.1. Reaction N+ (3P) + C2H4

Kinetics:

Recommended values at 300 K: k = 1.58 × 10−9 cm3 s−1 (±20%) for C2H4

k = 1.55 × 10−9 cm3 s−1 (±20%) for C2D4.

The experimental determinations of Rakshit (1980), Smith & Adams (1980), and McEwan et al. (1998) are in relatively good agreement, given their uncertainties (see Table 6). The most recent determination of Anicich et al. (2004) suggests a lower value of the rate constant. This discrepancy with the previous works can be explained by the different experimental conditions used in this later study, as for the reactions of N+ with methane and acetylene. This value of k is equal to the Langevin value (1.58 × 10−9 cm3 s−1) within uncertainties, therefore most probably independent of temperature as for the reactions of N+ with methane and acetylene. Note that Smith & Adams (1980) did not observe any significant isotope effect for this reaction. So the recommended values are the Langevin rates for both reactions.

Products:

Recommended yields: C2H+2/C2H+3/C2H+4/HCN+/HCNH+/CHCN+/CH2CN+, 0.12/0.32/0.38/0.02/0.10/0.01/0.05.

Table 14 summarizes the product branching ratio data for the N+ + C2H4 reaction. For clarity, the neutrals associated with the ions are not indicated in Table 14 due to a too high number of product channels. However, they are reported in Table 17 which summarizes our recommended values for all ion–molecule reactions.

All studies give product branching ratios which are in fair agreement for the N+ (3P) + C2H4 reaction, except the study of Rakshit (1980), which is surprising. No details are given on the instrument mass resolution in this study and one might suspect that the C2H+4 production seen by Rakshit (1980) is in fact the sum of the three species (C2H+4, C2H+3, and C2H+2) observed by the other groups (sum about 70%–87% according to the above studies). Moreover, the adduct CH3CNH+ is a priori unstable under single collision conditions. C. Alcaraz et al. (unpublished) used 15N and D isotopic labeling in order to avoid mass overlap, as in the studies of McEwan et al. (1998) and of Smith & Adams (1980). The yields reported by Anicich et al. (2004) are questionable for the same reasons already mentioned above. The studies of McEwan et al. (1998) and of Smith & Adams (1980) deduced branching ratios by analyzing both reactions with C2H4 and C2D4, assuming no isotopic effect. Except from the study of Rakshit (1980), all studies are in good agreement for the branching ratios of the three main species C2H+2, C2H+3, and C2H+4 within uncertainties. Disagreements are however observed for the N-bearing species HCN+, HCNH+, CHCN+, and CH2CN+. HCN+ and HCNH+ are not observed in the study of Anicich et al. (2004), because of the mass overlap with the main products C2H+3 and C2H+4. McEwan et al. (1998) do not observe CH2CN+, whereas Anicich et al. (2004), Smith & Adams (1980), and C. Alcaraz et al. (unpublished) revealed branching ratios between 0.05 and 0.10 for this product. Alcaraz et al. only observed 1% of HCN+ versus 10%–15% in the cases of Smith & Adams (1980) and of McEwan et al. (1998). These differences are significant and hard to explain. Only the work of C. Alcaraz et al. (unpublished) allows a direct measurement of the branching ratios, thanks to complete isotopic labeling. We therefore recommend this determination of the branching ratios for the N-bearing species. For the reaction with deuterated ethylene C2D4, it seems reasonable to estimate that the yields are the same as for the reaction with C2H4.

Other branching ratio values should however be employed if the N+ ions pertain some kinetic energy. Indeed, C. Alcaraz et al. (unpublished) also measured the variation of the reaction cross sections with collision energy between 0.2 and 6.4 eV in the center-of-mass frame. The HCNH+, CHCN+, and CH2CN+ production cross sections strongly decrease with collision energy, and the C2H+2 and C2H+4 cross sections slightly increase. A complex evolution of the cross section for forming the C2H+3 product has been observed: an important decrease between 0.05 and 0.3 eV collision energy, followed by an increase, similar to the decrease of C2H+2 and C2H+4. These observations seem to confirm a charge transfer mechanism forming C2H+2 and C2H+4, and a long lifetime complex involved in the formation of HCNH+, CHCN+, and CH2CN+. The formation of C2H+3 could be explained by the competition of two formation routes: the pathway C2H+3 + NH at low collision energy, and the dissociative charge transfer C2H+3 + H + N at high collision energy. The collision energy dependence of the cross section for forming HCN+ presents a different pattern from the other N-bearing species: a very minor product at low collision energy, it becomes the most important N-bearing species at high energy. From the cross sections, one can deduce the corresponding branching ratios. C. Alcaraz et al. (unpublished) thus observed an evolution of the branching ratios with the collision energy: C2H+4/C2H+3/C2H+2/HCN+/HCNH+/CHCN+/CH2CN+ from 0.38/0.32/0.12/0.02/0.10/0.01/0.05 to 0.53/0.23/0.17/0.02/0.03/0.01/0.01 when the collision energy increases from 0.2 to 6.4 eV in the center-of-mass frame. The most noticeable effect concerns the relative increase of C2H+4 production.

5.3.5.2. Reaction N+ (1D) + C2H4

Kinetics

Recommended value at 300 K: k = 1.58 × 10−9 cm3 s−1 (±30%)

Rakshit (1980) studied the reactivity of a mixture of N+ ground and metastable states (see Table 6). The measured value for the rate constant is very surprising (3.80 × 10−9 cm3 s−1), as it is about twice the Langevin value (1.58 × 10−9 cm3 s−1). C. Alcaraz et al. (unpublished) performed preliminary experiments for this reaction, generating reactant ions involving just the metastable N+ (1D) state. Their measured rate constant is equal to the Langevin rate within experimental uncertainties, as for the N+ (3P) ground state reaction with ethylene. Therefore this is our recommended rate constant value.

Products

Recommended yields: C2H+2/C2H+3/C2H+4/(HCN+ + HCNH+ + CHCN+ + CH2CN+), 0.26/0.22/0.46/0.06.

Table 14 shows the product branching ratio data for the N+ (1D) + C2H4 reaction.

Table 14. Product Branching Ratio Data for the N+ (3P) + C2H4 Reaction

Reaction C2H+2 C2H+3 C2H+4 HCN+ HCNH+ CHCN+ CH2CN+ CH3CNH+ Reference
N+ (3P) + C2H4 0.10 0.30 0.25 0.15 0.10   0.10 ... (Smith & Adams 1980)
N+ (3P) + C2H4 ... ... 0.76 ... ... ... ... 0.24 (Rakshit 1980)
N+ (3P) + C2H4 0.10 0.25 0.35 0.10 0.15 0.05 ... ... (McEwan et al. 1998)
N+ (3P) + C2H4 0.13 0.44 0.30 ... ... 0.04 0.10 ... (Anicich et al. 2004)
N+ (3P) + C2H4 0.12 0.32 0.38 0.02 0.10 0.01 0.05 ... (Alcaraz et al., unpublished)*
N+ (3P) + C2H4 0.10 0.25 0.35 0.10 0.15 0.05 ... ... (McEwan & Anicich 2007) (compilation)
N+ (1D) + C2H4 ... ... 0.76 ... ... ... ... 0.24 (Rakshit 1980)
N+ (1D) + C2H4 0.26 0.22 0.46 ... 0.06** ... ... ... Alcaraz et al., unpublished)*

Notes. *Extrapolation at low collision energy of cross section measurements for 15N++C2D4. **Estimated yield for the sum of N-bearing products (HCN+, HCNH+, CHCN+ and CH2CN+).

Download table as:  ASCIITypeset image

The branching ratios found by Rakshit (1980) are the same as for the ground-state reactivity and have to be considered with care for the same reason as described above. In contrast, the preliminary results of C. Alcaraz et al. (unpublished) show a strong variation of the product branching ratios between the N+ (3P) and N+ (1D) reactions. These experiments reveal a significantly higher yield of C2H+4, C2H+3, and C2H+2 product ions, than for the N+ (3P) ground-state reaction (+20%, −30%, and +120%, respectively) and much weaker signals from N-bearing species. These observations indicate that the charge transfer process is most probably the dominant reaction mechanism for the N+ (1D) state reaction with C2H4. With the present knowledge, we recommend the product yields estimated from the measurements of Alcaraz et al., but further studies are necessary to precise these values.

5.3.6. Reaction N+ (3P) + C2H6

Kinetics

Estimated value at 300 K: k = 1.60 × 10−9 cm3 s1

The kinetics of this reaction has never been measured, to our knowledge. However, the rate constants for the reactions of N+ with methane, ethylene, and acetylene are all very close to the Langevin rate. By analogy, the rate constant value is therefore estimated to be equal to the Langevin rate, i.e., 1.60 × 10−9 cm3 s−1. By analogy with the reactions with methane and ethylene, it is also reasonable to estimate that the reaction of N+ (1D) occurs at the Langevin rate.

Products

Recommended yields: (C2H+5+ NH)/(C2H+4+ NH2)/(C2H+3+ NH3)/(HCNH++ CH4), 0.10/0.55/0.25/0.10

The ion products and their respective branching ratios have been identified and quantified by McEwan et al. (1998) so they are our recommended values. No study exists for the reaction of N+ (1D) metastable state with ethane.

5.3.7. Reaction N+ (3P) + C3H8

Kinetics

Estimated value at 300 K: k = 1.8 × 10−9 cm3 s−1 (±20%)

The rate constant of this reaction, measured at 300 K by Dryahina et al. (2011) with the SIFT technique, was found to be equal to 2.0 × 10−9 cm3 s−1 (±20%), very close to the Langevin rate (1.8 × 10−9 cm3 s−1), which is our recommended value. It is thus expected to be the same value at 150 K. By analogy with the reactions with methane and ethylene, it is also reasonable to estimate that the reaction of N+ (1D) occurs at the Langevin rate.

Products

Recommended yields: C2H+3/C2H+4/C2H+5/C3H+5/C3H+6/C3H+7/C3H+8, 0.12/0.25/0.36/0.11/0.05/0.09/0.02.

The ion products of this reaction have been unambiguously identified by Dryahina et al. (2011) by using isotope labeling for both the reacting ion (14N+ or 15N+) and the propane target molecule (C3H8 or C3D8). Their measured product ion yields are our recommended values. No N-bearing ion products were found. The neutrals associated with the other product channels are less clear and discussed in the paper from Dryahina et al. (2011). According to these authors, they can result from processes which are either dissociative charge transfer and/or hydride (H) transfer producing NH neutrals. The production of NH2 and NH3 associated with the measured hydrocarbon ions is also possible, based on exoergicity considerations. No study exists for the reaction of N+ (1D) metastable state with propane, but it is reasonable to estimate that the (dissociative) charge transfer products dominate, as for the reactions of N+ (1D) with methane and ethylene.

5.3.8. Radiative Electron Recombination Reaction: N+ + e

Kinetics

Recommended value: k = 3.5 × 10−12 (Te/300)−0.7 cm3 s−1.

The radiative recombination rate of an electron with N+ has never been measured to our knowledge. However, calculations have been made for the recombination rate constant as a function of electron energy and these theoretical studies are reviewed in a compilation report (Kato & Asano 1999). The calculated rate constant varies between 10−12 and 10−11 cm3 s−1 in the 1–10 eV electron energy rang. Beyond 10 eV, it decreases linearly with electron energy as the electron energy increases. This study is consistent with the previous estimated value of 3.5 × 10−12 (Te/300)−0.7 cm3 s−1 reported in the compilation of Schmidt et al. (1988). In conclusion, in ionospheric conditions, N+ will be essentially quenched by reactions with molecules, rather than by radiative recombination with electrons.

6. N++2 DOUBLY CHARGED IONS

Ionization by solar UV photons, as well as by solar wind electrons, can produce N++2 doubly charged ions (dications). Double ionization processes can represent up to a few percent of the total ion yield (see below). Nevertheless, to date, dications have almost never been considered in ionospheric models. Most of the electronic states of molecular doubly charged ions are dissociative due to Coulomb explosion, but ground-state N++2 ions are stable under ionospheric conditions. They have not been detected in any ionosphere to date, even if they have been predicted for more than 35 years to be present in the ionosphere of the Earth (Avakyan 1979; Prasad & Furman 1975; Simon et al. 2005), and more recently in Titan (Lilensten et al. 2005) and Venus (Gronoff et al. 2007). N++2 densities have been calculated to be about 10−4 of the total ion density in the case of Titan. In this review paper, we present the production processes, the lifetime and the reactivity of N++2 ions: the latter is still less than adequately understood as there have been very few experimental studies to date. For more details, see the recent review article by Thissen et al. (2011) on doubly charged ions in planetary atmospheres.

6.1. Production

The double ionization threshold of molecular nitrogen, together with the spectroscopy of the states located up to 11 eV above threshold have been measured very accurately by Ahmad et al. (2006) using coincidence methods. The double-ionization threshold of 42.88 ± 0.01 eV is found to be in excellent agreement with previous experimental (Dawber et al. 1994; Hochlaf 1996) and theoretical (Senekowitsch et al. 1991) works. Ahmad et al. (2006) show that the double ionization at threshold is nondissociative and becomes dissociative into N+ + N+ above 44.5eV, with a lifetime in the microsecond time range.

The cross sections for the production of N++2 doubly charged molecular ions are very difficult to measure, due to the mass to charge overlap with the N+ fragment. Moreover, it is essential to distinguish between double ionization producing mainly dissociative doubly charged ions and production of long-lived N++2 ions.

The cross sections for the production of doubly charged N++2 ions by photoionization have only very recently been measured experimentally (Wehlitz 2010), by selecting the thermal N++2 ions among the N+ energetic ions, and detailed results are not yet published. In order to predict the N++2 density which could be expected in the ionospheres of Earth (Simon et al. 2005), Venus (Gronoff et al. 2007), or Titan (Lilensten et al. 2005), these authors have used an empirical method based on the observation of a proportionality between electron impact ionization (Bahati et al. 2001), and photoionization (Samson 1990), as Samson showed that the double-photoionization cross section is proportional to the electron-impact ionization of the monocation multiplied by the total photoabsorption of the neutral species. In these studies, it was proposed to use a proportionality constant of 0.133 in order to obtain the photoionization cross section (Gronoff et al. 2007). The measured cross sections by Wehlitz (2010) show that this estimate was excellent and that the double-ionization cross section increases from threshold up to 100 eV photon energy, where it represents about 4% of total ionization.

For electron impact double-ionization cross sections, Märk (1975) performed a study using a mass spectrometer with sufficient mass resolution and sensitivity in order to detect accurately the 14N15N++ doubly charged ions, which appear as a peak at a mass-to-charge ratio (m/z) of 14.5. Märk provides absolute cross sections for electron impact double ionization from threshold up to 170 eV electron energy. The maximum cross section is 1.3 × 10−18 cm2 (±5%) at 125 eV electron energy. This result is in very good agreement with the previous results of Halas & Adamczyk (1972) which extend to 600 eV, but has been criticized by Straub et al. (1996) and Tian & Vidal (1998) at a later date. The latter authors showed that the measurements made by Märk (1975) were probably underestimated by 20%. However, there is an excellent agreement between the measurements of Märk (1975) and the very recent ones of Ferreira et al. (2012), who measured the ratio of the double- to single-ionization cross sections by electron impact up to 400 eV electron energy.

Here we should add a comment concerning the cross sections for molecular N++2 production, and their application to ionospheric questions. Laboratory measurements are based on the detection of an ionic signature corresponding to the presence in the spectrometer, at a specific time, of doubly charged ions. However, molecular dications such as N++2 are often metastable species toward direct dissociation into N+ and N+. Therefore, an instrument probing the ionic population at longer times will inevitably observe a smaller amount of dications, the measured cross section being dependent on the instrument. The instruments of Märk (1975) and Halas & Adamczyk (1972) were probing the dications at times in the range of 1–5 ms, which is orders of magnitude shorter than the usual chemical lifetime considered in ionospheres (in the range of 1 s). One should therefore try to assess the effect of these differences in timescale. The consequent effect of population reduction is difficult to quantify, but fortunately, in the case of nitrogen, Lundqvist et al. (1996) measured specific lifetimes, which show that the transition between short lifetimes (nanosecond scale) and long lifetimes (beyond 3 ms scale) occurs abruptly between two vibrational states. As Mathur et al. (1995) have shown that N++2 ions have no or very few states with a lifetime in the range of micro or milliseconds, we propose to consider that values of absolute cross sections recorded in the microsecond time range are still valid for the ionospheric time range.

Though the density of ions in ionospheres is insufficient to require consideration of the competitive production of molecular dications by ionization of monocations, let us mention that data exist for electron impact ionization of N+2 and have been recorded by Bahati et al. (2001). To our knowledge, there is only a very recent, not yet published, work (J. M. Bizau 2011, private communication) for the equivalent process induced by photons.

6.2. Lifetime and Quenching

The metastable lifetime of N++2 toward its dissociation into N+ + N+ has been determined by Mathur et al. (1995) by monitoring the decay curves of N++2 beams in a heavy-ion storage ring. The longest component of the lifetime is measured to be of the order of 3 s. Mathur et al. performed ab initio configuration interaction calculations of the potential energy curve for the lowest-energy state of N++2 in order to compute tunneling times for each vibrational level. They find that lifetimes in the range of 3 s would correspond to v = 10 of the ground state, and therefore propose the association of the experimentally measured lifetime to the effect of interactions of N++2 with the residual gas in the ring, and consider that lifetime of v = 0 is effectively infinite. Further experimental evidence of the long lifetime of lower excited states is provided by Ahmad et al. (2006) showing that a stable dication signature is still visible in the coincident mass spectra at v = 7 of the A(1Πu) excited state of the dication, about 2.5 eV above the N++2 ground state.

Higher excited states of N++2 have short lifetimes (Ahmad et al. 2006) due to dissociation, but also resulting, for the D1Σ+u state, from fluorescence decay. Already identified in 1958 (Carroll 1958), these fluorescence bands have been assigned to the D1Σ+uX1Σ+g transition by Cossart et al., corresponding to the (0–0) band (Cossart et al. 1991) and to the (1–1) band (Cossart et al. 1985), respectively. The lifetime of the D1Σ+u state has been measured to be 6 ± 0.5 ns (Olsson & Larsson 1988). The cross section for the photo production of this emission has been determined by Ehresman et al. (2003) in the energy range from threshold (51eV) to 66 eV, where it peaks at a value of 1 × 10−20 cm2. It is interesting to note that, as shown by Ahmad et al. (2006), the emission is associated with the production of the non-dissociative ground state of N++2. Therefore, in ionospheres, the formation of the D1Σ+u state results in the formation of stable N++2 ions, which will decay by reactions with the neutral gases.

6.3. Reactions

The reactivity of molecular doubly charged ions is still poorly known, but from the existing experimental data for reactions with rare gas atoms, diatomic as well as polyatomic molecules, single charge transfer is usually the dominant product channel. However, surprisingly chemical reactions involving chemical bond rearrangement have been observed (see, for example, the review articles from Price (2003, 2007) and from Roithova & Schröder (2007a, 2007b).

6.3.1. Reaction N++2 + N2

Kinetics:

Recommended value at 300 K: k = 1.7 × 10−9 cm3 s−1 (±25%)

For the reaction of N++2 ions with N2, only very partially published results (Lilensten et al. 2005; Thissen et al. 2004) obtained by some of the co-authors are available. These results have been measured in a guided ion beam (GIB) apparatus for collision energies between 0.2 and 20 eV (Thissen et al. 2004). In order to distinguish between N++2 and N+ reactions, kinetic energy discrimination against the fast N+ ions has been used. The efficiency of this methodology has been checked by using 15N14N++ reactant ions, dications which can be unambiguously distinguished from singly charged atomic ions. The value for the rate constant at 300 K has been estimated from extrapolations of these data (Lilensten et al. 2005). It gives a value of 2.4 × 10−9 cm3 s−1 (±25%). It is nearly equal to the Langevin rate (1.66 × 10−9 cm3 s−1) within experimental uncertainties, but slightly higher, which is impossible. So the recommended rate constant value is the Langevin rate at 300 K and 150 K.

Products:

Recommended yields: (N+2 + N+2)/(N+2 + N+ + N), 0.90/0.10

Isotopic labeling has been used to identify the reaction products. In the case of the 14N++2 + 15N2 reaction at ECM = 0.2 eV, the charge transfer of one electron giving 14N+2 + 15N+2 is the main observed reaction channel (Thissen et al. 2004). But the reaction also produces some dissociative charge transfer of the target molecule giving 14N+2 + 15N+ + 15N products. Let us note that the eventual production of dissociation induced by collision giving 14N+ ions could not have been detected, as it has the same m/z ratio as the parent ions 14N++2.

6.3.2. Reaction N++2 + H2

Recommended value at 300 K: k = 3.1 × 10−9 cm3 s−1 (±25%) for H2

Estimated value at 300 K: k = 2.2 × 10−9 cm3 s−1 (±25%) for D2.

The reactions of N++2 with H2/D2 have recently been investigated (Lockyear et al. 2011). The rate constant is estimated to be equal to the Langevin rate and independent of temperature. This estimate is supported by the very good agreement with the absolute cross section measured for the NH+ + H+ + N reaction channel, multiplied by the corresponding branching ratio and converted into a partial rate constant. The uncertainty on this value is hard to estimate, but it seems reasonable to evaluate it to be equal to about 25%. It is naturally assumed that the reaction with D2 is also governed by the Langevin rate.

Products:

Recommended yields: (N+2 + H+2)/(N+ + H+2 + N)/(N+2 + H+ + H)/(N+ + H+ + (N + H))/(NH+ + H+ + N), 0.53/0.32/0.10/0.03/0.02 for H2

(N+2 + D+2)/(N+ + D+2 + N)/(N+2 + D+ + D)/(N+ + D+ + (N + D))/(ND+ + D+ + N), 0.61/0.24/0.12/0.02/0.01 for D2

Table 15 gives the product branching ratios for the N++2 + H2 and N++2 + H2 reactions.

Table 15. Product Branching Ratios for the N++2 + H2 and N++2 + D2 Reactions (X is H or D)

Reaction N+2+ X+2 N++ X+2+ N N+2+ X++ X N++ X++ (N + X) NX++ X++ N Reference
N++2 + H2 0.53 0.32 0.10 0.03 0.02 (Lockyear et al. 2011)*
N++2 + D2 0.61 0.24 0.12 0.02 0.01 (Lockyear et al. 2011)**

Notes. *Values measured for the reaction of N++2 with H2 at 0.9 eV CM collision energy. **Values measured for the reaction of N++2 with D2 at 1.8 eV CM collision energy.

Download table as:  ASCIITypeset image

The experiments of Lockyear et al. (2011) reveal that, at elevated collision energies (∼1 eV CM), charge transfer to form N+2 + H+2, N+ + H+2, and N+2 + H+ + H dominates the reactivity; a reactivity almost certainly dominated by the ground electronic state of N++2 which is the major species in the reactant beam. However, the formation of NH+ is also observed, a product which quantum-chemical investigations indicate must arise from the reaction of a triplet state of N++2. The formation of the NH+ product is assigned to the long-lived, but low-lying, 3Σ+u excited state of N++2 reacting at close to the Langevin rate; this long-lived state of N++2 is present as a minor species in the reactant ion beam. These observations indicate that for fully comprehensive modeling of the role of N++2 in Titan's ionosphere, a consideration of the electronic state distribution of the dication may well be required. Again further experimental investigations of reaction of N++2 with H2 would be highly desirable.

6.3.3. Reaction N++2 + CH4

Kinetics:

Recommended value at 300 K: k = 1.8 × 10−9 cm3 s−1 (±25%).

For the reaction of N++2 ions with CH4, the results of studies available in the literature (Lilensten et al. 2005; Thissen et al. 2004) are augmented by additional data presented for the first time in this review. These results have been measured in a GIB apparatus for collision energies between 0.2 and 16 eV for the reaction of 15N++2 with CD4 (Thissen et al. 2004). The recommended value for the rate constant at 300 K has been estimated from extrapolations of these data (Lilensten et al. 2005). It is close to the Langevin rate (2.36 × 10−9 and 2.20 × 10−9 cm3 s−1 for the reaction with CH4 and CD4, respectively) and the rate constant at 150 K is estimated to be the same.

Products:

Recommended yields: (N+2 + CD+4)/(N+2 + CD+3 + D)/(N+2 + CD+2 + D2)/(DCND+ + ND+ + D)/(DCN+ + ND+2 + D), 0.22/0.54/0.15/0.06/0.03

Isotopic labeling has been used to identify the reaction products. For the reaction of 15N++2 with CD4 at ECM = 0.2 eV (Thissen et al. 2004), the charge transfer of one electron is dominant, giving N+2 + CD+4, but also CD+3 and CD+2 product ions resulting from dissociative charge transfer. In addition, minor reactive channels giving ND+, ND+2, DCN+, and DCND+ have also been observed. Product branching ratios cannot be precisely derived from these experiments, as the derivation of such ratios would have required coincidence measurements between singly charged product ions as in the studies from Lockyear et al. (2011) in the case of the reactions of N++2 with H2 and D2. Ion pairs given in our recommended yields are the most reasonable estimated ones. These preliminary experiments reveal the need for further studies.

6.3.4. Reaction N++2 + C2H2

Kinetics:

Estimated value at 300 K: k = 2.3 × 10−9 cm3 s−1.

This reaction has never been studied. By analogy with the reactions of N++2 with N2, H2, CH4, and C2H4 (see below for C2H4), it is reasonable to estimate the rate constant to be equal to the Langevin rate, at 300 K and at 150 K.

Products:

By analogy with the reactions of N++2 with CH4 and C2H4 (see below for C2H4), it is reasonable to estimate that the main products come from the single charge transfer giving N+2 + C2H+2 and from the dissociative charge transfer giving N+2 + C2H+ + H and N+2 + C+2 + H2, which are the dissociative channels observed by Mackie et al. (2003) for the dissociative photoionization of acetylene. Minor products coming from chemical reactions, such as HCNH+ and HCN+, can also be expected, especially at low collision energy.

6.3.5. Reaction N++2 + C2H4

Kinetics:

Recommended value at 300 K: k = 2.2 × 10−9 cm3 s−1 (±25%)

The recommended value for the rate constant at 300 K has been estimated from extrapolations of the measurements made at collision energies between 0.2 and 20 eV for the reaction of 15N++2 with C2D4 (Thissen et al. 2004). It is equal to the Langevin rate (2.36 × 10−9 cm3 s−1) within experimental uncertainties, so the rate constant at 150 K is most probably the same.

Products:

Recommended yields: (N+2 + C2D+4)/(N+2 + C2D+3 + D)/(N+2 + C2D+2 + D2)/(N+2 + C2D+ + D2 + D)/(N+2 + CD+3 + CD)/(N+2 + CD+2 + CD2)/(DCND+ + ND+ + CD)/(DCN+ + ND+2 + CD), 0.26/0.41/0.18/0.01/0.03/0.06/0.02/0.03

Isotopic labeling has been used to identify the reaction products. For the reaction of 15N++2 with C2D4 at ECM = 0.2 eV (Thissen et al. 2004) as for the reaction of N++2 with methane, the main products come from single charge transfer, non-dissociative as well as dissociative, giving N+2 + (C2D+4, C2D+3, C2D+2, C2D+, CD+3, and CD+2) ions. These are the same dissociation products as those observed for the dissociative photoionization of ethylene (Mackie et al. 2003), except H+ (D+) which could not have been detected in the experiments of Thissen et al. (2004). The same minor reactive channels as for the reaction with methane have been observed giving ND+, ND+2, DCN+, and DCND+. Product branching ratios cannot be precisely derived from these experiments, as the derivation of such ratios would have required coincidence measurements between singly charged product ions as in the studies from Lockyear et al. (2011) in the case of the reactions of N++2 with H2 and D2. Ion pairs given in our recommended yields are the most reasonable estimated ones. These preliminary experiments reveal the need for further studies.

6.3.6. Reaction N++2 + C2H6

Kinetics:

Estimated value at 300 K: k = 2.6 × 10−9 cm3 s−1 (±25%).

This reaction has never been studied. By analogy with the reactions of N++2 with N2, H2, CH4, and C2H4, it is reasonable to estimate the rate constant to be equal to the Langevin rate, independent of temperature.

Products:

By analogy with the reactions of N++2 with CH4 and C2H4, it is reasonable to estimate that the main products come from the nondissociative (N+2 + C2H+6) and dissociative single charge transfer giving N+2 + (C2H+5 + H, C2H+4 + H2, C2H+3 + H2 + H, C2H+2 + 2H2, and CH+3 + CH3) products. The dissociation channels are supposed to be the same ones as those observed for the dissociative photoionization of ethane (Mackie et al. 2002; Stockbauer 1973). Minor products coming from chemical reactions, such as HCNH+, HCN+, NH+, and NH+2, can also be expected.

6.3.7. Reaction N++2 + C3H8

Kinetics:

Estimated value at 300 K: k = 2.8 × 10−9 cm3 s−1 (±25%).

This reaction has never been studied. By analogy with the reactions of N++2 with N2, H2, CH4, and C2H4, it is reasonable to estimate the rate constant to be equal to the Langevin rate, independent of temperature.

Products:

By analogy with the reactions of N++2 with CH4 and C2H4, it is reasonable to estimate that the main products come from the nondissociative (N+2 + C3H+8) and dissociative single charge transfer giving N+2 + (C3H+7 + H, C3H+5 + H2 + H, C2H+5 + CH3, C2H+4 + CH4, and C2H+3 + CH3 + H2) products. The dissociation channels are proposed to be the same ones as those observed for the dissociative photoionization of propane (Stockbauer & Inghram 1976). Minor products coming from chemical reactions, such as HCNH+, HCN+, NH+ and NH+2, can also be expected.

6.3.8. Reactions with Electrons: Dissociative Recombination and Collision-induced Dissociation

N++2 + e → N+ + N dissociative recombination

N++2 + e → N++ + N + e. collision-induced dissociation

Kinetics:

Recommended value: k = 5.8 × 10−7 (Te/300)−0.5 cm3 s−1 (±25%) for electron recombination

Absolute cross section measurements have been carried out by Seiersen et al. (2003) using the heavy-ion storage ring ASTRID in the relative collision energy range 10−4 to 10 eV. They studied the recombination of 14N15N++ and deduced from their data analysis a rate constant of

which is about 2.5 times greater than the dissociative recombination of N+2 at room temperature. Seiersen et al. did not measure the rate constant for dissociation without capture, a competing process that could be induced by the interaction with electrons, but they show that the later process is probably negligible for molecular doubly charged ions. Beyond 10.9 eV collision energy, the asymmetric dissociation into N++ + N seems to become the dominant process.

7. N++ DOUBLY CHARGED IONS

7.1. Production

In 1941, Hagstrum & Tate (1941) observed the formation of N++ doubly charged atomic ions in the dissociative ionization of N2 by electron impact. Only a very minor fraction of the N++2 ions created by electron ionization dissociate to N++ ions and an associated neutral N fragment. The photon-induced process was studied by Franceschi et al. (2007). They showed that the N++2 asymmetric dissociation has a threshold at 59.9 ± 0.2 eV and exhibits a maximum cross section of 2 × 10−20 cm2 at 85 eV. Crowe & McConkey (1973) measured the cross section for electron impact formation of N++ and showed that it peaks around 190 eV, with a cross section of 1.9 × 10−18 cm2.

N++ ions could also be formed in planetary atmospheres by double ionization of N atoms or ionization of N+ singly charged ions. The double ionization potential of an N atom is 44.135 eV (Ralchenko et al. 2011). The double photoionization cross section of N atoms has been measured by Samson et al. (Samson & Angel 1990; Samson et al. 1996). The N double photoionization cross section reaches a maximum of 2.1 × 10−19 cm2 at 60 eV photon energy and the ratio of doubly charged to singly charged ions has been measured to be 6.6% at this energy (Samson & Angel 1990).

The cross section for the photoionization of N+ ions to give N++ doubly charged ions has been measured by Kjeldsen et al. (2002). It rises sharply from the threshold at 29.60 eV and exhibits many resonances. The continuum cross-section value is 7.0× 10−18 cm2 at threshold and decreases slowly down to 3.6 × 10−18 and 1.1 × 10−18 cm2 at 50 and 80 eV photon energies, respectively. The experimental cross section is in good agreement with previous theoretical calculations (Nahar & Pradhan 1997). Tawara & Kato (1999) made a review of all measurements of the cross section for ionization of N+ by electron impact. As for the photoionization cross section of N+, it has a sharp onset and then reaches a maximum of 0.5 to 1 × 10−16 cm2, according to the different authors, at about 90 eV electron energy.

7.2. Lifetime and Quenching

The lowest metastable state of N++, about 7 eV above the 2P ground state, is the 4PJ state with J = 1/2, 3/2, and 5/2 (Ralchenko et al. 2011). These J states have lifetimes of 0.98, 13.4, and 3.2 ms, respectively (Fang et al. 1993). It is short with respect to the time between collisions under ionospheric pressure conditions, so these states will be relaxed by radiative decay to the ground state, before they can react by collisions.

The N++ ions produced by double ionization of N2 by 125 eV electron impact have a broad kinetic energy distribution of about 6 eV, centered around 4 eV, according to the old measurements of Kieffer & van Brunt (1967). It is thus likely that their kinetic energy is not, or only partially, relaxed by collisions prior to reaction.

7.3. Reactions

7.3.1. Reaction N++ + N2

Recommended value: k = 2.04 × 10−9 cm3 s−1 (±20%)

The rate constant of this reaction has been measured by two groups (Church & Holzscheiter 1980; Fang & Kwong 1997). Church & Holzscheiter (1980) measured it at 300 K and found a value of 2.8 × 10−9 cm3 s−1 (±20%). The rate constant measured by Fang & Kwong (1997) at a laboratory energy of 2.7 eV (1.8 eV CM and an equivalent temperature calculated to be 1.3 × 104 K) is very close to the value at 300 K and equal to the Langevin rate within experimental uncertainties (2.04 × 10−9 cm3 s−1). So the Langevin rate is our recommended rate constant value. According to Church & Holzscheiter (1980), the product ions are N+ ions with a mean kinetic energy of 2–3 eV, resulting from the dissociative single charge transfer giving N+ + N+ + N.

7.3.2. Reaction N++ + H2

Recommended value: k = 3.4 × 10−11 (±10%) cm3 s−1

Fang & Kwong (1997) measured the rate constant of this reaction at a laboratory energy of 2.7 eV (0.34 eV cm) for which they calculate the equivalent temperature (2.9 × 103 K). This is the only experimental study for this reaction and the products are unknown. The measured rate constant (3.4 × 10−11 (±10%) cm3 s−1) is about two orders of magnitude lower than the reaction of N++ with N2 (see above), much lower than the Langevin rate (3.2 × 10−9 cm3 s−1). Three reaction channels are energetically possible: the single charge transfer giving N+ + H+2, the dissociative charge transfer giving N+ + H+ + H, and the double charge transfer giving N + H+ + H+.

7.3.3. Radiative electron recombination reaction N++ + e

The electron recombination reactions of atomic doubly charged ions can only involve radiative recombination processes. For such reactions, there are only theoretical calculations (Nahar & Pradhan 1997; see also the review report from Kato & Asano 1999), due to the enormous difficulty in measuring such rates experimentally. All calculations give very close rate constant values as a function of the electron temperature. A typical expression was given by Aldrovandi & Pequignot (1973) as

where Te is the electron temperature in K.

8. SUMMARY AND CONCLUSIONS

In this paper, we have presented a critical review of the reactions of all nitrogen species, atomic and molecular, neutral, singly charged and doubly charged ions, with main neutral constituents of Titan's atmosphere. All recommended and estimated values for the rate constants and product yields are gathered in Tables 1618 for neutral, ion–molecule and ion–electron recombination reactions, respectively. In Table 16, recommended rate constant values of neutral reactions are given at 300 K, as well as their temperature dependence, specifying the temperature range of the associated studies. Products and their branching ratios at 300 K are indicated when they are reasonably known. In Table 17, recommended rate constant values for ion–molecule reactions are given at 300 K and 150 K. Products and their branching ratios are given at 300 K. In Table 18, rate constants of electron-ion recombination reactions are given as a function of electron temperature. Products are indicated together with recommended yields.

Table 16. Summary for Reactions of Neutral Active Nitrogen Species

Reaction Products Branchinga Ratio k (cm3 s−1) at 300 K k Temperature Dependence
with Uncertainty (Temperature Range)
N2 (A 3Σ+u) + N2 N2 (X 1Σ+g) + N2 1 ⩽3 × 10−18 ?
N2 (A 3Σ+u) + H2 N2 (X 1Σ+g) + H2 0.9 3.5 × 10−15 (±60%) 2.2 × 10−10 exp (−3500/T)
  N2 (X 1Σ+g) + H + H 0.1   (240–370 K)
N2 (A 3Σ+u) + CH4 N2 (X 1Σ+g) + CH4 0.9 3.0 × 10−15 (±60%) 1.3 × 10−10 exp (−3170/T)
  N2 (X 1Σ+g) + CH3 + H 0.1   (300–360 K)
N2 (A 3Σ+u) + C2H2 N2 (X 1Σ+g) + C2H + H 0.52 ± 5% 1.40 × 10−10 (±60%) ?
  N2 (X 1Σ+g) + C2 + H2 0.48 ± 5%    
N2 (A 3Σ+u) + C2D2 N2 (X 1Σ+g) + C2D + D 0.33 ± 5% 1.45 × 10−10 (±60%) ?
  N2 (X 1Σ+g) + C2 + D2 0.67 ± 5%    
N2 (A 3Σ+u) + C2H4 N2 (X 1Σ+g) + C2H3 + H 0.30 ± 5% 9.7 × 10−11 (±60%) ?
  N2 (X 1Σ+g) + C2H2 + H2 0.70 ± 5%    
N2 (A 3Σ+u) + C2D4 N2 (X 1Σ+g) + C2D3 + D 0.13 ± 5% 9.3 × 10−11 (±60%) ?
  N2 (X 1Σ+g) + C2D2 + D2 0.87 ± 5%    
N2 (A 3Σ+u) + C2H6 N2 (X 1Σ+g) + C2H6 ? 2.3 × 10−13 (±60%) 1.6 × 10−10 exp (−1980/T)
  N2 (X 1Σ+g) + C2H5 + H ?   (300–370 K)
  N2 (X 1Σ+g) + C2H4 + H2 ?    
N2 (A 3Σ+u) + C3H8 N2 (X 1Σ+g) + C3H8 ? 1.3 × 10−12 (±60%) ?
  N2 (X 1Σ+g) + C3H7 + H ?    
  N2 (X 1Σ+g) + C3H6 + H2 ?    
N (4S) + H NH + hν 1 1.0 × 10−19  
N (4S) + CH2 HCN + H 0.50 8.0 × 10−11  
  HNC + H 0.50    
N (4S) + CH3 H2CN + H 0.90 ± 10% 8.5 × 10−11 (±25%) 6.2 × 10−11
  HCN + H + H 0.10 ± 50%   (150–200 K)
N (4S) + CD3 D2CN + D 0.90 ± 10% 8.5 × 10−11  
  DCN + D + D 0.10 ± 50%    
N (4S) + C2H3 CH2CN + H 0.83 ± 10% 7.7 × 10−11 (±40%)  
  C2H2 + NH 0.17± 50%    
N (4S) + C2H5 C2H4 + NH 0.65 1.1 × 10−10 (±25%)  
  H2CN + CH3 0.35    
N (4S) + C2D5 C2D4 + ND 0.65 ± 50% 1.1 × 10−10  
  D2CN + CD3 0.35 ± 50%    
N (2D) + N2 N (4S) + N2 1 1.7 × 10−14 (±40%) 1.0 × 10−13 exp (−510/T)
        (198–372 K)
N (2D) + H2 NH + H 1 2.2 × 10−12 (±25%) 4.6 × 10−11 exp (−880/T)
        (213–300 K)
N (2D) + CH4 CH2NH/CH3N + H 0.8 ± 20% 4.0 × 10−12 (±40%) 7.1 × 10−11 exp (−750/T)
  NH + CH3 0.2 ± 50%   (223–292 K)
N (2D) + CD4 CD2ND/CD3N + D 0.8 2.6 × 10−12 (±40%) 3.3 × 10−11 exp (−700/T)
  ND + CD3 0.2   (223–292 K)
N (2D) + C2H2 HCCN + H 0.9 ± 10% 6.5 × 10−11 (±25%) 1.6 × 10−10 exp (−270/T)
  cyclic-HCCN + H 0.1 ± 50%   (223–293 K)
N (2D) + C2D2 DCCN + D 0.9 6.25 × 10−11 (±25%) 1.4 × 10−10 exp (−240/T)
  cyclic-DCCN + D 0.1   (223–293 K)
N (2D) + C2H4 CH2NCH + H 0.67 4.3 × 10−11 (±25%) (2.3 ± 0.3) × 10−10 exp (−500 ± 50/T)
  cyclic-CH2(N)CH + H 0.23   (230–292 K)
  CH2CNH + H 0.05    
  HCN/HNC + CH3 0.01    
  CH3CN/CH3NC + H 0.01    
  CH2NC/CHNCH + H2 0.01    
  CH2N/CHNH + CH2 0.01    
N (2D) + C2D4 CD2NCD + D 0.67 3.8 × 10−11 (±25%) (2.4 ± 0.5) × 10−10 exp (−550 ± 50/T)
  cyclic-CD2(N)CD + D 0.23   (230–292 K)
  CD2CND + D 0.05    
  DCN/DNC + CD3 0.01    
  CD3CN/CD3NC + D 0.01    
  CD2NC/CDNCD + D2 0.01    
  CD2N/CDND + CD2 0.01    
N (2D) + C2H6 CH2=NH + CH3 0.79 ± 10% 1.9 × 10−11 (±25%) ?
  CH3CH=NH + H 0.12 ± 10%    
  NH + C2H5 0.06 ± 20%    
  CH2=CHNH2 + H 0.02 ± 20%    
  3CH2=NH + CH3 0.01 ± 20%    
N (2D) + C3H8 CH2NH + C2H5 ? 2.9 × 10−11 (±25%) ?
  CH3CHNH + CH3 ?    
  NH + C3H7 ?    
  C2H5CHNH + H ?    
N (2P) + N2 N (2D) + N2 1 3.3 × 10−17 (±60%) ?
N (2P) + H2 N (2D) + H2 1 1.9 × 10−15 (±60%) 3.5 × 10−13 exp (−950/T)
        (213–300 K)
N (2P) + CH4 N (2D) + CH4 1 8.5 × 10−14 (±25%) 5.0 × 10−13 exp (−490/T)
        (223–292 K)
N (2P) + CD4 N (2D) + CD4 1 6.0 × 10−14 (±25%) 3.1 × 10−13 exp (−480/T)
        (223–292 K)
N (2P) + C2H2 ... ... 2.3 × 10−11 (±40%) 1.0 × 10−10 exp (−440/T)
        (223–293 K)
N (2P) + C2D2 ... ... 2.0 × 10−11 (±40%) 7.1 × 10−11 exp (−380/T)
        (223–293 K)
N (2P) + C2H4 ... ... 3.0 × 10−11 (±25%) (1.4 ± 0.5) × 10−10 exp (−455 ± 90/T)
        (230–292 K)
N (2P) + C2D4 ... ... 3.0 × 10−11 (±25%) (1.3 ± 0.2) × 10−10 exp (−435 ± 50/T)
        (230–292 K)
N (2P) + C2H6 N (2D) + C2H6 1 5.4 × 10−13 (±25%) ?
N (2P) + C3H8 Quenching or chemical reaction? ? 1.9 × 10−12 (±25%) ?

Note. aEstimated values are in italic characters.

Download table as:  ASCIITypeset images: 1 2

Table 17. Summary for Reactions of Nitrogen Ions

Reaction Productsa Branching Ratioa k (cm3 s−1)a at 300 K k (cm3 s−1)a at 150 K
with Uncertaintya with Uncertainty
N+2 + N2 N2 + N+2 1 5.0 × 10−10 (±25%) 5.0 × 10−10 (±70%)
N+2 + H2 N2H+ + H 0.99 ± 5% 1.56 × 10−9 (±15%) 1.3 × 10−9 (±20%)
  H+2 + N2 0.01 ± 10%    
N+2 + D2 N2D+ + D 0.99 ± 5% 1.15 × 10−9 (±20%) 9.6 × 10−10 (±20%)
  D+2 + N2 0.01 ± 10%    
N+2 + CH4 CH+3 + H + N2 0.86 ± 5% 1.18 × 10−9 (±15%) 1.18 × 10−9 (±15%)
  CH+2 + H2 + N2 0.09 ± 5%    
  N2H+ + CH3 0.05 ± 10%    
N+2 + CD4 CD+3 + D + N2 0.88 ± 5% 1.10 × 10−9 (±20%) 1.10 × 10−9 (±20%)
  CD+2 + D2 + N2 0.07 ± 5%    
  N2D+ + CD3 0.05 ± 10%    
N+2 + C2H2 C2H+2 + N2 0.94 ± 5% 4.15 × 10−10 (±25%) 1.0 × 10−9 (±50%)
  N2H+ + C2H 0.06 ± 10%    
N+2 + C2H4 C2H+3 + H + N2 0.67 ± 5% 1.29 × 10−9 (±15%) 1.29 × 10−9 (±20%)
  C2H+2 + H2 + N2 0.23 ± 5%    
  N2H+ + C2H3 0.10 ± 10%    
N+2 + C2H6 C2H+5 + H + N2 0.14 ± 5% 1.30 × 10−9 (±30%) 1.30 × 10−9 (±30%)
  C2H+4 + H2 + N2 0.27 ± 5%    
  C2H+3 + H2 + H + N2 0.32 ± 5%    
  C2H+2 + 2H2 + N2 0.18 ± 5%    
  CH+3 + CH3 + N2 0.08 ± 10%    
  CH+4 + CH2 + N2 0.01 ± 10%    
N+2 + C2D6 C2D+5 + D + N2 0.14 1.25 × 10−9 (±30%) 1.25 × 10−9 (±30%)
  C2D+4 + D2 + N2 0.27    
  C2D+3 + D2 + D + N2 0.32    
  C2D+2 + 2 D2 + N2 0.18    
  CD+3 + CD3 + N2 0.08    
  CD+4 + CD2 + N2 0.01    
N+2 + C3H8 C3H+5 + H2 + H + N2 0.13 ± 5% 1.30 × 10−9 (±30%) 1.30 × 10−9 (±30%)
  C2H+5 + CH3 + N2 0.30 ± 5%    
  C2H+4 + CH4 + N2 0.17 ± 5%    
  C2H+3 + CH4 + H + N2 0.40 ± 5%    
14N+ + 15N2 15N+ + 14N15N 1 2.1 × 10−10 (±25%) 2.1 × 10−10 (±25%)
or 15N+ + 14N2 14N+ + 14N15N 1    
N+ + H2 NH+ + H 1 3.8 × 10−10 (±20%) 3.0 × 10−10 (±30%)
N+ + HD NH+ + D 0.25 ± 5% (300 K) 3.1 × 10−10 (±20%) 2.4 × 10−10 (±30%)
  ND+ + H 0.07 ± 10% (105 K)    
    0.75 ± 5% (300 K)    
    0.93 ± 5% (105 K)    
N+ + D2 ND+ + D 1 1.2 × 10−10 (±20%) 9.5 × 10−11 (±30%)
N+ (3P) + CH4 CH+3 + NH 0.5 ± 5% 1.15 × 10−9 (±20%) 1.0 × 10−9 (±20%)
  CH+4 + N 0.05 ± 10%    
  HCNH+ + H2 0.35 ± 5%    
  HCN+ + H2 + H 0.10 ± 10%    
N+ (1D) + CH4 CH+3 + NH 0.09 ± 10% 1.15 × 10−9 (±20%) 1.0 × 10−9 (±20%)
  CH+4 + N 0.40 ± 5%    
  HCNH+ + H2 0.33 ± 5%    
  HCN+ + H2 + H 0.18 ± 5%    
N+ (3P) + C2H2 C2H+2 + N 0.70 ± 5% 1.42 × 10−9 (±20%) 1.42 × 10−9 (±20%)
  CNC+ + H2 0.15 ± 5%    
  CHCN+ + H 0.15 ± 5%    
N+ (1D) + C2H2 C2H+2 + N 0.90 1.42 × 10−9 (±20%) 1.42 × 10−9 (±20%)
  CNC+ + H2 0.05    
  CHCN+ + H 0.05    
N+ (3P) + C2H4 C2H+2 + NH2 0.12 ± 5% 1.58 × 10−9 (±20%) 1.58 × 10−9 (±20%)
  C2H+3 + NH 0.32 ± 5%    
  C2H+4 + N 0.38 ± 5%    
  HCN+ + CH3 0.02 ± 10%    
  HCNH+ + CH2 0.10 ± 10%    
  CHCN+ + H2 + H 0.01 ± 10%    
  CH2CN+ + H2 0.05 ± 10%    
N+ (3P) + C2D4 C2D+2 + ND2 0.12 1.55 × 10−9 (±20%) 1.55 × 10−9 (±20%)
  C2D+3 + ND 0.32    
  C2D+4 + N 0.38    
  DCN+ + CD3 0.02    
  DCND+ + CD2 0.10    
  CDCN+ + D2 + D 0.01    
  CD2CN+ + D2 0.05    
N+ (1D) + C2H4 C2H+2 + NH2 0.26 ± 5% 1.58 × 10−9 (±30%) 1.58 × 10−9 (±30%)
  C2H+3 + NH 0.22 ± 5%    
  C2H+4 + N 0.46 ± 5%    
  HCN+ + CH3 (0.06 ± 10%    
  HCNH+ + CH2 for the sum of    
  CHCN+ + H2 + H N bearing ions)    
  CH2CN+ + H2      
N+ (3P) + C2H6 C2H+5 + NH 0.10 ± 10% 1.6 × 10−9 (±20%) 1.6 × 10−9 (±20%)
  C2H+4 + NH2 0.55 ± 5%    
  C2H+3 + NH3 0.25 ± 5%    
  HCNH+ + CH4 0.10 ± 10%    
N+ (1D) + C2H6 C2H+5 + NH ? 1.6 × 10−9 (±30%) 1.6 × 10−9 (±30%)
  C2H+4 + NH2 ?    
  C2H+3 + NH3 ?    
  HCNH+ + CH4 ?    
         
N+ (3P) + C3H8 C2H+3 + CH4 + NH 0.12 1.8 × 10−9 ± 20% 1.8 × 10−9 ± 20%
  C2H+4 + CH3 + NH 0.25    
  C2H+5 + CH2 + NH 0.36    
  C3H+5 + NH3 0.11    
  C3H+6 + NH2 0.05    
  C3H+7 + NH 0.09    
  C3H+8 + N 0.02    
N+ (1D) + C3H8 C2H+3 + CH4 + NH ? 1.8 × 10−9 ± 20% 1.8 × 10−9 ± 20%
  C2H+4 + CH3 + NH ?    
  C2H+5 + CH2 + NH ?    
  C3H+5 + NH3 ?    
  C3H+6 + NH2 ?    
  C3H+7 + NH ?    
  C3H+8 + N ?    
         
14N++2 + 15N2 14N+2 + 15N+2 0.9 ± 25% 1.7 × 10−9 (±25%) 1.7 × 10−9 (±25%)
  14N+2 + 15N+ + 15N 0.1 ± 25%    
N++2 + H2 N+2 + H+2 0.53 ± 5% 3.1 × 10−9 (±25%) 3.1 × 10−9 (±25%)
  N+ + H+2 + N 0.32 ± 5%    
  N+2 + H+ + H 0.10 ± 5%    
  N+ + H+ + (N + H) 0.03 ± 10%    
  NH+ + H+ + N 0.02 ± 10%    
N++2 + D2 N+2 + D+2 0.61 ± 5% 2.2 × 10−9 (±25%) 2.2 × 10−9 (±25%)
  N+ + D+2 + N 0.24 ± 5%    
  N+2 + D+ + D 0.12 ± 5%    
  N+ + D+ + (N + D) 0.02 ± 10%    
  ND+ + D+ + N 0.01 ± 10%    
N++2 + CH4 N+2 + CH+4 0.22 1.8 × 10−9 (±25%) 1.8 × 10−9 (±25%)
  N+2 + CH+3 + H 0.54    
  N+2 + CH+2 + H2 0.15    
  HCNH++ NH+ + H 0.06    
  HCN+ + NH+2 + H 0.03    
N++2 + CD4 N+2 + CD+4 0.22 ± 25% 1.8 × 10−9 (±25%) 1.8 × 10−9 (±25%)
  N+2 + CD+3 + D 0.54 ± 25%    
  N+2 + CD+2 + D2 0.15 ± 25%    
  DCND++ ND+ + D 0.06 ± 25%    
  DCN+ + ND+2 + D 0.03 ± 25%    
N++2 + C2H2 N+2 + C2H+2 ? 2.3 × 10−9 2.3 × 10−9
  N+2 + C2H+ + H ?    
  N+2 + C+2 + H2 ?    
  HCNH+ + CN+ ?    
  HCN+ + HCN+ ?    
N++2 + C2H4 N+2 + C2H+4 0.26 2.2 × 10−9 2.2 × 10−9
  N+2 + C2H+3 + H 0.41    
  N+2 + C2H+2 + H2 0.18    
  N+2 + C2H+ + H2 + H 0.01    
  N+2 + CH+3 + CH 0.03    
  N+2 + CH+2 + CH2 0.06    
  HCNH+ + NH+ + CH 0.02    
  HCN+ + NH+2 + CH 0.03    
N++2 + C2D4 N+2 + C2D+4 0.26 ± 25% 2.2 × 10−9 (±25%) 2.2 × 10−9 (±25%)
  N+2 + C2D+3 + D 0.41 ± 25%    
  N+2 + C2D+2 + D2 0.18 ± 25%    
  N+2 + C2D+ + D2 + D 0.01 ± 25%    
  N+2 + CD+3 + CD 0.03 ± 25%    
  N+2 + CD+2 + CD2 0.06 ± 25%    
  DCND+ + ND+ + CD 0.02 ± 25%    
  DCN+ + ND+2 + CD 0.03 ± 25%    
N++2 + C2H6 N+2 + C2H+6 ? 2.6 × 10−9 (±25%) 2.6 × 10−9 (±25%)
  N+2 + C2H+5 + H ?    
  N+2 + C2H+4 + H2 ?    
  N+2 + C2H+3 + H2 + H ?    
  N+2 + C2H+2 + 2H2 ?    
  N+2 + CH+3 + CH3 ?    
  HCNH+ + NH+ + CH3 ?    
  HCN+ + NH+2 + CH3 ?    
N++2 + C3H8 N+2 + C3H+8 ? 2.8 × 10−9 (±25%) 2.8 × 10−9 (±25%)
  N+2 + C3H+7 + H ?    
  N+2 + C3H+5 + H2 + H ?    
  N+2 + C2H+5 + CH3 ?    
  N+2 + C2H+4 + CH4 ?    
  N+2 + C2H+3 + CH3 + H2 ?    
  HCNH+ + NH+2 + C2H4 ?    
  HCN+ + NH+ + C2H6 ?    
N++ + N2 N+ + N+ + N 1 2.04 × 10−9 (±20%) 2.04 × 10−9 (±20%)
N++ + H2 N+ + H+2 ? 3.4 × 10−11 (±10%) 3.4 × 10−11 (±20%)
  N+ + H+ + H ?    
  N + H+ + H+ ?    

Note. aEstimated products and estimated rate constant and branching ratio values are in italic characters.

Download table as:  ASCIITypeset images: 1 2 3

Table 18. Summary for Electron–Ion Recombination Reactions of Nitrogen Ions

Reaction Products Recommended Yield Rate Constant
(cm3 s−1)
N+2 + e N (4S) 0.25 ± 25% 2.2 × 10−7 (Te/300)−0.39 (±25%)
  N (2D) 0.7 ± 25%  
  N (2P) 0.05 ± 25%  
N+ + e N + hv 1 3.5 × 10−12 (Te/300)−0.7*
N++2 + e N+ + N 1 5.8 × 10−7 (Te/300)−0.5 (±25%)
N++ + e N+ + hv 1 2.2 × 10−12 (Te/10000)−0.639*

Note. *Calculated value (see the text).

Download table as:  ASCIITypeset image

In Tables 16 and 17, we propose estimated values for rate constants and product yields, when no data are available and when we could reasonably make appropriate estimates, using our expertise for chemical reactions, as planetary scientists need them for their atmospheric chemical models. However we would like to underline the importance to check such estimates by detailed experimental or theoretical work. These estimated values are given in italic characters in Tables 1618 in order to clearly distinguish them from measured or calculated values. Uncertainties are given for all data (except calculated and some estimated ones). Most uncertainties are estimated based on our knowledge of the experimental techniques used. In the literature, uncertainties are often not given by the authors, or if they are, they are mostly estimated by the authors from experimental scatter, without taking into account any systematic errors, which are much more difficult to evaluate.

We also included in this paper a review of the main production processes, lifetimes, and quenching of the active nitrogen species. For all reactions, temperature and/or collision energy dependence of the rate constants are discussed in the light of the available information from the literature. However for many reactions, this information is still unavailable, in particular at low temperatures, where it is essential for the modeling of the chemistry of Titan's atmosphere. We also note that the variation of the reaction cross sections of N atoms and N+ atomic ions with collision energy up to a few eV is very important for some applications, in particular for the studies of atmospheric escape.

For neutral reactions, in summary, the most reactive nitrogen species is the N (2D) metastable state and all quenching reactions are slow. The reactions of the N2 (A 3Σ+u) state with N2, H2 and CH4 produce only or mainly the quenching into the N2 (X) ground state, but with C2H2 and C2H4, they produce the dissociation of the target molecule. Quenching is the main channel for reactions of the N (2P) higher metastable state with N2, H2 and saturated hydrocarbons. This quenching is predicted to produce the N (2D) lower metastable state and not the N (4S) ground state, but it would be important to check this assumption, by either experimental or theoretical work. If this assumption is not correct, it could have important consequences for Titan's atmospheric chemistry, as N (2D) is very reactive and N (4S) is not. With unsaturated hydrocarbons, N (2P) gives chemical reactions, but the products are unknown. The reactions of the N (2D) metastable state with hydrocarbons involve C–N bond formation, but product yields are not well known. The measurement of relevant rate constants at low temperatures still requires much more laboratory work. Let us note that caution should be exercised when extrapolating rate constants to temperatures outside the temperature range over which the temperature dependence was determined. When no experimental temperature dependence was measured, it is reasonable to predict that the rate constant in the 150–200 K range will have a lower value than at 300 K, when an energy barrier is supposed to be present for the reaction. But no precise estimate can be made without theoretical or experimental work. Much more critical is the situation with the products of the reactions of the neutrals of interest, as there are almost no complete measurements of reaction product branching ratios, even at 300 K. It is worth noting that for neutral–neutral reactions, experimental support is necessary for determining branching ratios, especially at low temperatures, as theoretical estimates can only be achieved using RRKM methods, when reactions behave statistically. Rate constants involving neutral species have been measured for a large number of reactions, but it is only in the past 15 years that the experimental and theoretical tools have become available to explore quantitatively the product branching ratios of reactions. The latest promising development is the use of the Advanced Light Source (ALS) at Berkeley which provides tuneable synchrotron radiation to photoionize the products from neutral–neutral reactions. The ionized products are then analyzed by time of flight or quadrupole mass spectrometry (Osborn et al. 2008; Soorkia et al. 2008). In addition, to our knowledge, no quantitative experimental investigations of the branching ratios of reactions involving atomic nitrogen have been made at low temperature.

For ion–molecule reactions, even though the experimental and theoretical work is plentiful for reactions at 300 K, the information is still incomplete at this temperature. However, with some exceptions, most reaction rate constants are equal or close to the Langevin rate, which is independent of temperature and is the maximum value of the rate constant for target molecules with no permanent dipole. The reactivity of the long-lived N+ (1D) metastable state and of the relevant doubly charged ions is still very poorly known. This lack of information can explain why these reactions have not yet been included in models of planetary atmospheres. In summary, the N+2 reactions with hydrocarbons produce mainly charge transfer (dissociative or non dissociative) and minor N2H+ products via H transfer, whereas the N+ (3P) reactions involve significant chemical reactions, in particular the formation of C-N bonds. The few N+ (1D) metastable state reactions which have been investigated give the same ion products as the N+ (3P) ground state reactions, but their branching ratios are different, with a higher yield for products arising from charge transfer. Let us note that the data concerning the products of ion–molecule reactions is much more complete than for neutral reactions, due to the fact that they can be easily detected by mass spectrometry. However, information concerning the neutral products associated with product ions is sometimes lacking and requires more theoretical or experimental work. More information is also desirable about the isomeric structure of products, as their further reactions can be affected. For example, the production of HCN and HNC in Titan's upper atmosphere has been recently modeled (Hébrard et al. 2012). This study reveals the need for more experimental measurements to improve the chemistry of these two isomers. It is also the case for the corresponding ions, as a very different reactivity is expected for HCN+ and HNC+ isomeric ions, because of their peculiar thermo-chemistry, as HNC+ is more stable than HCN+ by 0.98 eV and HNC is less stable than HCN by 0.58 eV (Hansel et al. 1998a, 1998b). For doubly charged ions, there are experimental studies for only very few reactions, which all show that their rate constant is equal to the Langevin rate. The dominant process for the N++2 reactivity is single charge transfer, both non-dissociative and dissociative, but surprisingly minor products coming from chemical reactions involving bond rearrangement are also observed. With hydrocarbons, there are only preliminary experiments for reactions of N++2 with methane and ethylene, so a huge effort should be made, in particular to measure the ion products with coincidence methods, as those developed by Price et al. (Lockyear et al. 2011; Price 2003, 2007).

The N+2 + e dissociative recombination reaction is still not well understood, as discrepancies remain between the different experimental and theoretical studies: this reaction certainly deserves further work to determine more precise N (2D)/N (2P)/N (4S) yields. The N++2 + e reaction also needs further investigation, as some questions are still unresolved, such as the electronic state of the N+ and N products formed by electron dissociative recombination or the branching ratio between the dissociative electron recombination (N+ + N) and the dissociation induced by electrons without electron capture (N+ + N+).

The study of isotope effects is a highlighted topic among planetary scientists for the last few years. Isotope effects can give a lot of information about the evolution of a planet's atmosphere. Fractionation of the 12C/13C and D/H ratios in methane or 15N/14N in molecular nitrogen have been recently investigated for Titan (Mandt et al. 2009, 2012; Nixon et al. 2012). This fractionation, due to atmospheric escape over geological times, requires an excellent knowledge of the isotopically resolved physicochemical processes, in particular the kinetic isotope effect, i.e. the difference of reactivity between isotopic species. For bimolecular reactions, D/H isotope effects can play a significant role and were reviewed in this paper, even though many data are still lacking as it can be seen in Tables 16 and 17. For carbon and nitrogen, the reactivity is the same for the different isotopic species. Indeed species labeled with 13C and/or 15N atoms are often used as an experimental route to better identify the reaction products, as was mentioned several times in this paper. However, it would be important to better address the question of atom exchange and scrambling of atoms involving carbon and/or nitrogen atoms in reactions, in order to better identify the route of isotopic molecular species in the chain of chemical reactions.

This paper is focused on the principal reactions which are the first steps in the chemistry of Titan's atmosphere. In the future, it would be important to make also a critical review of other reactions which play an important role later in the chemical chain. Studies should also be extended to reactions with other molecules present in Titan's atmosphere, such as HCN, which is more abundant than C2H6 and C3H8, or other unsaturated hydrocarbons such as C3H4, C3H6, or C4H2, aromatics such as C6H6, nitriles such as C2N2, CH3CN, HC3N, species, all of which are present in Titan's atmosphere (Bézard 2009; Magee et al. 2009; Teanby et al. 2008, 2009). Studies should involve reactions of other species, in particular radical or ionic hydrocarbons, as well as negative ions, which play an essential role in molecular growth in Titan's atmosphere. Such a review would help to address the important issue of the formation of complex molecules and aerosols in Titan's atmosphere. We must also note that we did not review in this paper neither the termolecular reactions that are present in Titan's atmosphere at low altitudes, nor reactions on surfaces.

Termolecular reactions of N atoms are not relevant for Titan's atmosphere, as N atoms are mainly formed by photodissociation above 1000 km altitude, where the pressure is too low for triple collisions to occur. However at low altitudes, termolecular chemistry can occur with ions formed by galactic cosmic rays (Gronoff et al. 2009a, 2009b, 2011; Molina-Cuberos et al. 1999). Anicich & McEwan (Anicich & McEwan 2001; Anicich et al. 2000; Milligan et al. 2001) studied experimentally termolecular ion–molecule reactions of the main ions of Titan's atmosphere. Vacher et al. (1997, 2000) and Anicich & McEwan (2001) found in particular a substantial clustering of HCNH+ ions with N2 molecules by triple collisions which could be very effective at altitudes below 90 km, according to Anicich & McEwan (2001).

Heterogeneous reactions of nitrogen species in Titan's atmosphere involve most likely interactions with surfaces of aerosols, i.e., surfaces composed of organic (C–H–N) compounds The different heterogeneous processes have been recently described in the book on Titan to be published by Cambridge University Press (Vuitton et al. 2012). For nitrogen species, very few experimental studies have been performed in conditions relevant to Titan. When N2 adsorbs on a surface at high temperature, it can dissociate (Ertl 1980) and the produced atomic nitrogen can proceed to further chemical reactions. This interaction is well known since the beginning of the 20th century and has been used for the catalytic production of ammonia and nitrogen-rich fertilizers (Haber–Bosch process) with substantial socio-economical consequences. However, it is possible to overcome the requirement of high temperature if the surface interacts with N2 molecular ions (Lancaster & Rabalais 1979). Concerning neutral nitrogen surface reactions, a very effective heterogeneous chemistry has been indirectly observed in experimental simulations at low temperatures of the first steps of Titan's ion chemistry (Thissen et al. 2009). The experimental results could not been interpreted without involving a very effective formation of HCN and NH3 on surfaces, probably induced by N atoms (Thissen et al. 2009). Other studies showed the importance of heterogeneous interaction of hydrogen atoms on the surfaces of Tholins, which can have further implications for the chemical composition of the atmosphere (Lavvas et al. 2008c; Sekine et al. 2008a, 2008b). Information on collisions of N+2 with Tholins (laboratory analogs of Titan's aerosols) has so far not been available. Related information may be presumably obtained from studies of hyperthermal collisions of nitrogen ions with hydrocarbons layers (C–H compounds, assumed to be C7–C8 chains) covering carbon and other surfaces (Herman et al. 2009, 2012; Keim et al. 2012). The main process is surface neutralization, the survival probability Sa(%) of nitrogen molecular ions being only 0.002%–0.02% (Herman et al. 2009, 2012). The nature of neutral products is not known. Mass spectra of products ions show sputtering of the surface material (Keim et al. 2012), no heterogeneous chemical reactions of N+2 have been so far clearly identified. For the other nitrogen ions (N+, N++2, N++), the correlation function between Sa and the ionization energy measured for other ions can be used to predict their survival probability (Herman et al. 2009), indicating a value of the same order of magnitude or smaller. At collision energies close to thermal energies, trapping of the incident particles on the surface may be important. Similar studies for other species need to further investigate the heterogeneous interaction of N2 and its ions with the surface of Tholins.

In conclusion, this work of evaluation by experts of all types of reactions is essential for planetary scientists, as well as for astrophysicists. Therefore, the new KIDA database (Wakelam et al. 2012) which includes a critical evaluation of neutral and ion reactions is a very promising and important tool for modelers.

We thank the FP6 and FP7 European Network Europlanet RI (Research Infrastructure) for travel support of several of us within the personnel exchange action. It initiated this review paper in collaboration between 11 different European laboratories. V.V. is grateful to the European Commission for the Marie Curie International Reintegration Grant No. 231013.

Annex: List of contacts among the authors for the different types of reactions and processes

Reactions of neutrals: Nadia Balucani, Piergiorgio Casavecchia, Sébastien Le Picard, Jean-Christophe Loison

Ion–molecule reactions: Odile Dutuit, Roland Thissen, Christian Alcaraz, Daniela Ascenzi, Pietro Franceschi, Paolo Tosi, Zdenek Herman, Jan Zabka, Stephen Price

Electron recombination reactions: André Canosa

Uncertainty propagation: Nathalie Carrasco, Pascal Pernot

Titan ion and neutral chemistry in models: Véronique Vuitton, Panayotis Lavvas

Please wait… references are loading.
10.1088/0067-0049/204/2/20