Chemistry:2,6-Lutidine

From HandWiki
2,6-Lutidine[1]
2,6-Lutidine.svg
Names
Preferred IUPAC name
2,6-Dimethylpyridine
Other names
Lutidine
Identifiers
3D model (JSmol)
105690
ChEBI
ChemSpider
EC Number
  • 203-587-3
2863
UNII
UN number 2734
Properties
C7H9N
Molar mass 107.153 g/mol
Appearance colorless oily liquid
Density 0.9252
Melting point −5.8 °C (21.6 °F; 267.3 K)
Boiling point 144 °C (291 °F; 417 K)
27.2% at 45.3 °C
Acidity (pKa) 6.72[2]
−71.72×10−6 cm3/mol
Hazards
NFPA 704 (fire diamond)
Flammability code 3: Liquids and solids that can be ignited under almost all ambient temperature conditions. Flash point between 23 and 38 °C (73 and 100 °F). E.g. gasolineHealth code 2: Intense or continued but not chronic exposure could cause temporary incapacitation or possible residual injury. E.g. chloroformReactivity code 0: Normally stable, even under fire exposure conditions, and is not reactive with water. E.g. liquid nitrogenSpecial hazards (white): no codeNFPA 704 four-colored diamond
3
2
0
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
☒N verify (what is ☑Y☒N ?)
Infobox references
Tracking categories (test):

2,6-Lutidine is a natural heterocyclic aromatic organic compound with the formula (CH3)2C5H3N. It is one of several dimethyl-substituted derivative of pyridine, all of which are referred to as lutidines. It is a colorless liquid with mildly basic properties and a pungent, noxious odor.

Occurrence and production

It was first isolated from the basic fraction of coal tar and from bone oil.[1]

A laboratory route involves condensation of ethyl acetoacetate, formaldehyde, and an ammonia source to give a bis(carboxy ester) of a 2,6-dimethyl-1,4-dihydropyridine, which, after hydrolysis, undergoes decarboxylation.[3]

It is produced industrially by the reaction of formaldehyde, acetone, and ammonia.[2]

Uses

2,6-Lutidine has been evaluated for use as a food additive owing to its nutty aroma when present in solution at very low concentrations.

Due to the steric effects of the two methyl groups, 2,6-lutidine is less nucleophilic than pyridine. Protonation of lutidine gives lutidinium, [(CH3)2C5H3NH]+, salts of which are sometimes used as a weak acid because the conjugate base (2,6-lutidine) is so weakly coordinating. In a similar implementation, 2,6-lutidine is thus sometimes used in organic synthesis as a sterically hindered mild base.[4] One of the most common uses for 2,6-lutidine is as a non-nucleophilic base in organic synthesis. It takes part in the formation of silyl ethers as shown in multiple studies.[5][6]

Oxidation of 2,6-lutidine with air gives 2,6-diformylpyridine:

C5H3N(CH3)2 + 2 O2 → C5H3N(CHO)2 + 2 H2O

Biodegradation

The biodegradation of pyridines proceeds via multiple pathways.[7] Although pyridine is an excellent source of carbon, nitrogen, and energy for certain microorganisms, methylation significantly retards degradation of the pyridine ring. In soil, 2,6-lutidine is significantly more resistant to microbiological degradation than any of the picoline isomers or 2,4-lutidine.[8] Estimated time for complete degradation was over 30 days.[9]

See also

Toxicity

Like most alkylpyridines, the LD50 of 2,6-dimethylpyridine is modest, being 400 mg/kg (oral, rat).[2]

References

  1. 1.0 1.1 An Encyclopedia of Chemicals, Drugs, and Biologicals (11th ed.), Merck, 1989, ISBN 091191028X , 5485
  2. 2.0 2.1 2.2 Shimizu, Shinkichi; Watanabe, Nanao; Kataoka, Toshiaki; Shoji, Takayuki; Abe, Nobuyuki; Morishita, Sinji; Ichimura, Hisao (2007). "Ullmann's Encyclopedia of Industrial Chemistry". Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH. doi:10.1002/14356007.a22_399. 
  3. Singer, Alvin; McElvain, S. M. (1934). "2,6-Dimethylpyridine". Organic Syntheses 14: 30. doi:10.15227/orgsyn.014.0030. 
  4. Prudhomme, Daniel R.; Park, Minnie; Wang, Zhiwei; Buck, Jason R.; Rizzo, Carmelo J. (2000). "Synthesis of 2′-Deoxyribonucleosides: Β-3′,5′-Di-o-benzoylthymidine". Org. Synth. 77: 162. doi:10.15227/orgsyn.077.0162. 
  5. Corey, E. J.; Cho, H.; Rücker, C.; Hua, D. H. (1981). "Studies with trialkylsilyltriflates: new syntheses and applications". Tetrahedron Letters 22 (36): 3455–3458. doi:10.1016/s0040-4039(01)81930-4. 
  6. Franck, Xavier; Figadère, Bruno; Cavé, André (1995). "Mild deprotection of tert-butyl and tert-amyl ethers leading either to alcohols or to trialkylsilyl ethers". Tetrahedron Letters 36 (5): 711–714. doi:10.1016/0040-4039(94)02340-H. ISSN 0040-4039. 
  7. Philipp, Bodo; Hoff, Malte; Germa, Florence; Schink, Bernhard; Beimborn, Dieter; Mersch-Sundermann, Volker (2007). "Biochemical Interpretation of Quantitative Structure-Activity Relationships (QSAR) for Biodegradation of N-Heterocycles: A Complementary Approach to Predict Biodegradability". Environmental Science & Technology 41 (4): 1390–1398. doi:10.1021/es061505d. PMID 17593747. Bibcode2007EnST...41.1390P. https://kops.uni-konstanz.de/bitstreams/6a1d118f-e52a-478f-8f9e-d8cbd9384229/download. 
  8. Sims, G. K.; Sommers, L. E. (1985). "Degradation of pyridine derivatives in soil". Journal of Environmental Quality 14 (4): 580–584. doi:10.2134/jeq1985.00472425001400040022x. Bibcode1985JEnvQ..14..580S. 
  9. Sims, G. K.; Sommers, L. E. (1986). "Biodegradation of Pyridine Derivatives in Soil Suspensions". Environmental Toxicology and Chemistry 5 (6): 503–509. doi:10.1002/etc.5620050601.