Hydrogen isocyanide

From Infogalactic: the planetary knowledge core
Jump to: navigation, search
Hydrogen isocyanide
Names
IUPAC names
hydrogen isocyanide
azanylidyniummethanide
Other names
isohydrocyanic acid
hydroisocyanic acid
isoprussic acid
Identifiers
ChEBI CHEBI:36856 YesY
ChemSpider 4937885 YesY
Jmol 3D model Interactive image
PubChem 6432654
  • InChI=1S/CHN/c1-2/h2H YesY
    Key: QIUBLANJVAOHHY-UHFFFAOYSA-N YesY
  • InChI=1/CHN/c1-2/h2H
  • [C-]#[NH+]
Properties
HNC
Molar mass 27.03 g/mol
Vapor pressure {{{value}}}
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
YesY verify (what is YesYN ?)
Infobox references

Hydrogen isocyanide is a chemical with the molecular formula HNC. It is a minor tautomer of hydrogen cyanide (HCN). Its importance in the field of astrochemistry is linked to its ubiquity in the interstellar medium.

Nomenclature

Both 'hydrogen isocyanide' and 'azanylidyniummethanide' are correct IUPAC names for HNC. Currently there is no preferred IUPAC name. The second one is according to the substitutive nomenclature rules, derived from the parent hydride azane (NH3) and the anion methanide (C).[1]

Molecular properties

Hydrogen isocyanide (HNC) is a linear triatomic molecule with C∞v point group symmetry. It is a zwitterion and an isomer of hydrogen cyanide (HCN).[2] Both HNC and HCN have large, similar dipole moments, with respectively μHNC = 3.05 Debye and μHCN = 2.98 Debye.[1] These large dipole moments facilitate the easy observation of these species in the interstellar medium.

HNC−HCN tautomerism

As HNC is higher in energy than HCN by 3920 cm−1 (46.9  kJ/mol), one might assume that the two would have an equilibrium ratio at T < 100 K of ([HNC]/[HCN])eq,T < 100 K < 10−25.[2] However, observations show a very different conclusion; ([HNC]/[HCN])observed is much higher than 10−25, and is in fact on the order of unity in cold environments. This is because of the potential energy path of the tautomerization reaction; there is an activation barrier on the order of roughly 12,000 cm−1 for the tautomerization to occur, which corresponds to a temperature at which HNC would already have been destroyed by neutral-neutral reactions.[3]

Spectral properties

In practice, HNC is almost exclusively observed astronomically using the J = 1→0 transition. This transition occurs at ~90.66 GHz, which is a point of good visibility in the atmospheric window, thus making astronomical observations of HNC particularly simple. Many other related species (including HCN) are observed in roughly the same window.[4][5]

Significance in the interstellar medium

HNC is intricately linked to the formation and destruction of numerous other molecules of importance in the interstellar medium—aside from the obvious partners HCN, HCNH+, and CN, HNC is linked to the abundances of many other compounds, either directly or through a few degrees of separation. As such, an understanding of the chemistry of HNC leads to an understanding of countless other species—HNC is an integral piece in the complex puzzle representing interstellar chemistry.

Furthermore, HNC (alongside HCN) is a commonly used tracer of dense gas in molecular clouds, as referenced in this paper.[6] Aside from the potential to use HNC to investigate gravitational collapse as the means of star formation, HNC abundance (relative to the abundance of other nitrogenous molecules) can be used to determine the evolutionary stage of protostellar cores. This is demonstrated in the aforementioned paper by Tennekes et al. In the same paper, the authors also elaborate on the HNC/HCN abundance ratio as a means of determining the temperature of the environment.

This paper[7] demonstrates a myriad of uses for knowledge of the abundance of HNC. In it, the HCO+/HNC line ratio is used to good effect as a measure of density of gas. This information provides great insight into the mechanisms of the formation of (Ultra-)Luminous Infrared Galaxies ((U)LIRGs), as it provides data on the nuclear environment, star formation, and even black hole fueling. Furthermore, the HNC/HCN line ratio is used to distinguish between photon-dissociation regions (PDRs) and X-ray-dissociation regions (XDRs) on the basis that [HNC]/[HCN] is roughly unity in PDR sources, but greater than unity in XDR sources.

The study of HNC is a relatively simple pursuit, and this is one of the greatest motivations for its study. Aside from having its J = 1→0 transition in a clear portion of the atmospheric window, as well as having numerous isotopomers also available for easy study, and in addition to having a large dipole moment that makes observations particularly simple, HNC is, in its molecular nature, a quite simple molecule. This makes the study of the reaction pathways that lead to its formation and destruction a good means of obtaining insight to the workings of these reactions in space. Furthermore, the study of the tautomerization of HNC to HCN (and vice versa), which has been studied extensively, has been suggested as a model by which more complicated isomerization reactions can be studied.[8][9][10]

Chemistry in the interstellar medium

HNC is found primarily in dense molecular clouds, though it is ubiquitous in the interstellar medium. Its abundance is closely linked to the abundances of other nitrogen containing compounds in a complex relationship partially demonstrated in the chart available on page 256 of this article.[11] HNC is formed primarily through the dissociative recombination of HNCH+ and H2NC+, and it is destroyed primarily through ion-neutral reactions with H+
3
and C+. These facts are corroborated in both this article[12] and this article.[13] Rate constants are taken from udfa.net, and data on fractional abundances is taken from this article.[14] Rate calculations were done at 3.16 × 105 years, which is considered early time, and at 20 K, which is a typical temperature for dense molecular clouds.

Formation Reactions
Reactant 1 Reactant 2 Product 1 Product 2 Rate constant Rate/[H2]2 Relative Rate
HCNH+ e HNC H 9.50×10−8 4.76×10−25 3.4
H2NC+ e HNC H 1.80×10−7 1.39×10−25 1.0
Destruction Reactions
Reactant 1 Reactant 2 Product 1 Product 2 Rate constant Rate/[H2]2 Relative Rate
H+
3
HNC HCNH+ H2 8.10×10−9 1.26×10−24 1.7
C+ HNC C2N+ H 3.10×10−9 7.48×10−25 1.0

These four reactions are merely the four most dominant, and thus the most significant in the formation of the HNC abundances in dense molecular clouds; there are dozens more reactions for the formation and destruction of HNC. Though these reactions primarily lead to various protonated species, HNC is linked closely to the abundances of many other nitrogen containing molecules, for example, NH3 and CN. The pathways leading between these species can be found in the paper by Turner et al. that is linked above. The abundance HNC is also inexorably linked to the abundance of HCN, and the two tend to exist in a specific ratio based on the environment, as noted in the paper by Hiraoka et al. that is linked above. This is because the reactions that form HNC can often also form HCN, and vice versa, depending on the conditions in which the reaction occurs, and also that there exist isomerization reactions for the two species. A simplified pathway showing many of the methods of HNC formation and destruction is available as Fig. 10 from Turner et al.

Astronomical detections

HNC was first detected in June 1970 by L. E. Snyder and D. Buhl[15] using the 36-foot radio telescope of the National Radio Astronomy Observatory (NRAO). The main molecular isotope, H12C14N, was observed via its J = 1→0 transition at 88.6 GHz in six different sources: W3 (OH), Orion A, Sgr A(NH3A), W49, W51, DR 21(OH). A secondary molecular isotope, H13C14N, was observed via its J = 1→0 transition at 86.3 GHz in only two of these sources: Orion A and Sgr A(NH3A). HNC was then later detected extragalactically in 1988 by C. Henkel, R. Mauersberger, and P. Schilke[16] using the IRAM 30-m telescope at the Pico de Veleta in Spain. It was observed via its J = 1→0 transition at 90.7 GHz toward IC 342.

A number of detections have been made towards the end of confirming the temperature dependence of the abundance ratio of [HNC]/[HCN]. A strong fit between temperature and the abundance ratio would allow observers to spectroscopically detect the ratio and then extrapolate the temperature of the environment, thus gaining great insight into the environment of the species. In 1986, Goldsmith et al.[17] measured the abundances of rare isotopes of HNC and HCN along the OMC-1 and determined that the abundance ratio varies by more than an order of magnitude in warm regions versus cold regions. In 1992, Schilke et al.[18] measured abundances of HNC, HCN, and deuterated analogs along the OMC-1 ridge and core and confirmed the temperature dependence of the abundance ratio. Helmich and van Dishoeck[19] performed a survey of the W 3 Giant Molecular Cloud in 1997 in which they detected over 24 different molecular isotopes, comprising over 14 distinct chemical species, including HNC, HN13C, and H15NC. This survey further confirmed the temperature dependence of the abundance ratio, [HNC]/[HCN], this time ever confirming the dependence of the isotopomers.

These are not the only detections of importance of HNC in the interstellar medium. In 1997, Pratap et al.[20] observed HNC along the TMC-1 ridge and found that its abundance relative to HCO+ to be constant along the ridge—this led credence to the reaction pathway that posits that HNC is derived initially from HCO+. One significant astronomical detection that demonstrated the practical use of observing HNC occurred in 2006 by Tennekes et al.[21], in which the authors detected and then used the abundances of various nitrogenous compounds (including HN13C and H15NC) to determine the stage of evolution of the protostellar core Cha-MMS1 based on the relative magnitudes of the abundances.

On 11 August 2014, astronomers released studies, using the Atacama Large Millimeter/Submillimeter Array (ALMA) for the first time, that detailed the distribution of HCN, HNC, H2CO, and dust inside the comae of comets C/2012 F6 (Lemmon) and C/2012 S1 (ISON).[3][4]

See also

External links

References

  1. The suffix 'ylidyne' refers to the lost of three hydrogen atoms from the nitrogen atom in azanium (NH+
    4
    ) See the IUPAC Red Book 2005 Table III, "Suffixes and endings", p. 257.
  2. Chin Fong Pau, Warren J. Hehre "Heat of formation of hydrogen isocyanide by ion cyclotron double resonance spectroscopy"; J. Phys. Chem., 1982, 86 (3), pp. 321–322; doi:10.1021/j100392a006
  3. Lua error in package.lua at line 80: module 'strict' not found.
  4. Lua error in package.lua at line 80: module 'strict' not found.

Lua error in package.lua at line 80: module 'strict' not found.

  1. ^ Lua error in package.lua at line 80: module 'strict' not found.
  2. ^ Lua error in package.lua at line 80: module 'strict' not found.
  3. ^ Lua error in package.lua at line 80: module 'strict' not found.
  4. ^ Lua error in package.lua at line 80: module 'strict' not found.
  5. ^ Lua error in package.lua at line 80: module 'strict' not found.
  6. ^ Lua error in package.lua at line 80: module 'strict' not found.
  7. ^ Lua error in package.lua at line 80: module 'strict' not found.
  8. ^ Lua error in package.lua at line 80: module 'strict' not found.
  9. ^ Lua error in package.lua at line 80: module 'strict' not found.
  10. ^ Lua error in package.lua at line 80: module 'strict' not found.
  11. ^ Lua error in package.lua at line 80: module 'strict' not found.
  12. ^ Lua error in package.lua at line 80: module 'strict' not found.
  13. ^ Lua error in package.lua at line 80: module 'strict' not found.
  14. ^ Lua error in package.lua at line 80: module 'strict' not found.
  15. ^ Lua error in package.lua at line 80: module 'strict' not found.
  16. ^ Lua error in package.lua at line 80: module 'strict' not found.