Hydrogen iodide

Hydrogen iodide
Names
IUPAC name
Hydrogen iodide
Other names
hydroiodic acid
Identifiers
10034-85-2 YesY
RTECS number MW3760000
Properties
HI
Molar mass 127.904 g/mol
Appearance Colorless gas
Density 2.85 g/mL (−47 °C)
Melting point −50.80 °C (−59.44 °F; 222.35 K)
Boiling point −35.36 °C (−31.65 °F; 237.79 K)
approximately 245 g/100 ml
Acidity (pKa) −10 (in water, estimate)[1]

2.8 (in acetonitrile)[2]

Structure
Terminus
0.38 D
Hazards
Main hazards Toxic, corrosive, Harmful and Irritant
Safety data sheet See: data page
hydrogen iodide
hydroiodic acid
R-phrases R20, R21, R22, R35
S-phrases S7, S9, S26, S45
NFPA 704
Flash point Non-flammable
Related compounds
Other anions
Hydrogen fluoride
Hydrogen chloride
Hydrogen bromide
Hydrogen astatide
Supplementary data page
Refractive index (n),
Dielectric constantr), etc.
Thermodynamic
data
Phase behaviour
solidliquidgas
UV, IR, NMR, MS
N verify (what is YesYN ?)
Infobox references

Hydrogen iodide (HI) is a diatomic molecule. Aqueous solutions of HI are known as hydroiodic acid or hydriodic acid, a strong acid. Hydrogen iodide and hydroiodic acid are, however, different in that the former is a gas under standard conditions, whereas the other is an aqueous solution of said gas. They are interconvertible. HI is used in organic and inorganic synthesis as one of the primary sources of iodine and as a reducing agent.

Properties of hydrogen iodide

HI is a colorless gas that reacts with oxygen to give water and iodine. With moist air, HI gives a mist (or fumes) of hydroiodic acid. It is exceptionally soluble in water, giving hydroiodic acid. One liter of water will dissolve 425 liters of HI, the most concentrated solution having only four water molecules per molecule of HI.[3]

Hydroiodic acid

Once again, although chemically related, hydroiodic acid is not pure HI but a mixture containing it. Commercial "concentrated" hydroiodic acid usually contains 48–57% HI by mass. The solution forms an azeotrope boiling at 127 °C with 57% HI, 43% water. The high acidity is caused by the dispersal of the ionic charge over the anion. The iodide ion is much larger than the other common halides, which results in the negative charge being dispersed ove

HBr(g) + H2O(l) → H3O+(aq) + Br(aq)Ka ≈ 109
HCl(g) + H2O(l) → H3O+(aq) + Cl(aq)Ka ≈ 106

Synthesis

The industrial preparation of HI involves the reaction of I2 with hydrazine, which also yields nitrogen gas:[4]

2 I2 + N2H4 → 4 HI + N2

When performed in water, the HI must be distilled.

HI can also be distilled from a solution of NaI or other alkali iodide in concentrated phosphoric acid (note that concentrated sulfuric acid will not work for acidifying iodides, as it will oxidize the iodide to elemental iodine).

Another way HI may be prepared is by bubbling hydrogen sulfide steam through an aqueous solution of iodine, forming hydroiodic acid (which is distilled) and elemental sulfur (this is filtered):

H2S + I2 → 2 HI + S

Additionally, HI can be prepared by simply combining H2 and I2:

H2 + I2 → 2 HI

This method is usually employed to generate high-purity samples.

For many years, this reaction was considered to involve a simple bimolecular reaction between molecules of H2 and I2. However, when a mixture of the gases is irradiated with the wavelength of light equal to the dissociation energy of I2, about 578 nm, the rate increases significantly. This supports a mechanism whereby I2 first dissociates into 2 iodine atoms, which each attach themselves to a side of an H2 molecule and break the H−H bond:[5]

H2 + I2 + 578 nm radiation → H2 + 2 I → I···H···H···I → 2 HI

In the laboratory, another method involves hydrolysis of PI3, the iodine equivalent of PBr3. In this method, I2 reacts with phosphorus to create phosphorus triiodide, which then reacts with water to form HI and phosphorous acid:

3 I2 + 2 P + 6 H2O → 2 PI3 + 6 H2O → 6 HI + 2 H3PO3

Key reactions and applications


4 HI + O2 → 2 H2O + 2 I2
HI + I2 → HI3

HI3 is dark brown in color, which makes aged solutions of HI often appear dark brown.

Like HBr and HCl, HI add to alkenes:[6]

HI + H2C=CH2 → H3CCH2I

HI is also used in organic chemistry to convert primary alcohols into alkyl halides.[7] This reaction is an SN2 substitution, in which the iodide ion replaces the "activated" hydroxyl group (water):

HI is preferred over other hydrogen halides because the iodide ion is a much better nucleophile than bromide or chloride, so the reaction can take place at a reasonable rate without much heating. This reaction also occurs for secondary and tertiary alcohols, but substitution occurs via the SN1 pathway.

HI (or HBr) can also be used to cleave ethers into alkyl iodides and alcohols, in a reaction similar to the substitution of alcohols. This type of cleavage is significant because it can be used to convert a chemically stable[7] and inert ether into more reactive species. In this example diethyl ether is split into ethanol and iodoethane:

The reaction is regioselective, as iodide tends to attack the less sterically hindered ether carbon.

HI is subject to the same Markovnikov and anti-Markovnikov guidelines as HCl and HBr.

HI reduces certain α-substituted ketones and alcohols replacing the α substituent with a hydrogen atom.[6]

Illicit use of hydroiodic acid

Hydroiodic acid is currently listed as a Federal DEA List I Chemical. Owing to its usefulness as a reducing agent, reduction with HI and red phosphorus has become the most popular method to produce methamphetamine in the United States. Clandestine chemists react pseudoephedrine (recovered from nasal decongestant pills) with hydroiodic acid and red phosphorus under heat. HI reacts with pseudoephedrine to form iodoephedrine, an intermediate which is reduced primarily to methamphetamine.[8] This reaction is stereospecific, producing only (d)-methamphetamine.

Lab using the HI/P method

Due to its listed status and closely monitored sales, clandestine chemists now use red phosphorus and iodine to generate hydroiodic acid in situ.[9]

References

  1. Bell, R.P. The Proton in Chemistry. 2nd ed., Cornell University Press, Ithaca, NY, 1973.
  2. Raamat, E.; Kaupmees, K.; Ovsjannikov, G.; Trummal, A.; Kütt, A.; Saame, J.; Koppel, I.; Kaljurand, I.; Lipping, L.; Rodima, T.; Pihl, V.; Koppel, I. A.; Leito, I. "Acidities of strong neutral Brønsted acids in different media." J. Phys. Org. Chem. 2013, 26, 162-170. doi:10.1002/poc.2946
  3. Holleman, A. F.; Wiberg, E. "Inorganic Chemistry" Academic Press: San Diego, 2001. ISBN 0-12-352651-5.
  4. Greenwood, N. N. and A. Earnshaw. The Chemistry of the Elements. 2nd ed. Oxford: Butterworth-Heineman. p 809–815. 1997.
  5. Holleman, A. F. Wiberg, E. Inorganic Chemistry. San Diego: Academic Press. p. 371, 432–433. 2001.
  6. 1 2 Breton, G. W., P. J. Kropp, P. J.; Harvey, R. G. “Hydrogen Iodide” in Encyclopedia of Reagents for Organic Synthesis (Ed: L. Paquette) 2004, J. Wiley & Sons, New York. DOI: 10.1002/047084289.
  7. 1 2 Bruice, Paula Yurkanis. Organic Chemistry 4th ed. Prentice Hall: Upper Saddle River, N. J, 2003 p. 438–439, 452.
  8. Skinner, Harry F. "Methamphetamine Synthesis via HI/Red Phosphorous Reduction of Ephedrine". Forensic Science International, 48 128-134 (1990)
  9. Skinner HF. "Identification and quantitation of hydriodic acid manufactured from iodine, red phosphorus and water" . Journal of the Clandestine Laboratory Investigation Chemists Association 1995;5(4):12; Microgram 1995;28(11):349

See also: Nishikata, E., T.; Ishii, and T. Ohta. “Viscosities of Aqueous Hydrochloric Acid Solutions, and Densities and Viscosities of Aqueous Hydroiodic Acid Solutions”. J. Chem. Eng. Data. 26. 254-256. 1981.

External links

This article is issued from Wikipedia - version of the Monday, May 02, 2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.