Pyridyne

Last updated
Pyridynes Pyridynes.svg
Pyridynes

Pyridyne in chemistry is the pyridine analogue of benzyne. [1] Pyridynes are the class of reactive intermediates derived from pyridine. Two isomers exist, the 2,3-pyridine (2,3-didehydropyridine) and the 3,4-pyridyne (3,4-didehydropyridine). The reaction of 3-bromo-4-chloropyridine with furan and lithium amalgam gives 1,4-epoxy-dihydroquinoline through the 2,3-pyridyne intermediate. The reaction of 4-bromopyridine with sodium in liquid ammonia gives both 3-aminopyridine and 4-aminopyridine through the 3,4-pyridyne intermediate and an E1cB-elimination reaction. [2]

Contents

History

Pyridynes were first postulated by Levine and Leake in 1955. [3] In 1969 Zoltewicz and Nisi trapped 3,4-pyridyne in a reaction of 3-bromopyridine with methylmercaptan and sodium amide in ammonia. The methylthio and amino pyridines were found to be formed in the same ratio. [4]

Pyridyne intermediate.svg

In 1972 Kramer and Berry inferred the formation of 3,4-pyridyne in gas-phase photolysis of pyridine-3-diazonium-4-carboxylate via time-of-flight mass spectrometry. The dimer compound diazabiphenylene was detected. [5] In 1988 Nam and Leroy reported the matrix isolation (13K, Ar) of 3,4-pyridyne by photolysis of 3,4-pyridinedicarboxylic anhydride with the IR-spectrum revealing an acetylenic bond in the same way as ortho-benzyne.

Scope

Strategies involving pyridynes have been employed in the total syntheses of ellipticine [6] [7] and (S)-Macrostomine. [8]

Related Research Articles

<span class="mw-page-title-main">Pyridine</span> Heterocyclic aromatic organic compound

Pyridine is a basic heterocyclic organic compound with the chemical formula C5H5N. It is structurally related to benzene, with one methine group (=CH−) replaced by a nitrogen atom. It is a highly flammable, weakly alkaline, water-miscible liquid with a distinctive, unpleasant fish-like smell. Pyridine is colorless, but older or impure samples can appear yellow, due to the formation of extended, unsaturated polymeric chains, which show significant electrical conductivity. The pyridine ring occurs in many important compounds, including agrochemicals, pharmaceuticals, and vitamins. Historically, pyridine was produced from coal tar. As of 2016, it is synthesized on the scale of about 20,000 tons per year worldwide.

<span class="mw-page-title-main">Sodium amide</span> Chemical compound

Sodium amide, commonly called sodamide, is the inorganic compound with the formula NaNH2. It is a salt composed of the sodium cation and the azanide anion. This solid, which is dangerously reactive toward water, is white, but commercial samples are typically gray due to the presence of small quantities of metallic iron from the manufacturing process. Such impurities do not usually affect the utility of the reagent. NaNH2 conducts electricity in the fused state, its conductance being similar to that of NaOH in a similar state. NaNH2 has been widely employed as a strong base in organic synthesis.

<span class="mw-page-title-main">Oxazole</span> Chemical compound

Oxazole is the parent compound for a vast class of heterocyclic aromatic organic compounds. These are azoles with an oxygen and a nitrogen separated by one carbon. Oxazoles are aromatic compounds but less so than the thiazoles. Oxazole is a weak base; its conjugate acid has a pKa of 0.8, compared to 7 for imidazole.

Arynes and benzynes are highly reactive species derived from an aromatic ring by removal of two substituents. Arynes are examples of didehydroarenes, although 1,3- and 1,4-didehydroarenes are also known. Arynes are examples of strained alkynes.

Thiazole, or 1,3-thiazole, is a heterocyclic compound that contains both sulfur and nitrogen. The term 'thiazole' also refers to a large family of derivatives. Thiazole itself is a pale yellow liquid with a pyridine-like odor and the molecular formula C3H3NS. The thiazole ring is notable as a component of the vitamin thiamine (B1).

The Hofmann rearrangement is the organic reaction of a primary amide to a primary amine with one fewer carbon atom. The reaction involves oxidation of the nitrogen followed by rearrangement of the carbonyl and nitrogen to give an isocyanate intermediate. The reaction can form a wide range of products, including alkyl and aryl amines.

<span class="mw-page-title-main">Nucleophilic aromatic substitution</span> Chemical reaction mechanism

A nucleophilic aromatic substitution is a substitution reaction in organic chemistry in which the nucleophile displaces a good leaving group, such as a halide, on an aromatic ring. Aromatic rings are usually nucleophilic, but some aromatic compounds do undergo nucleophilic substitution. Just as normally nucleophilic alkenes can be made to undergo conjugate substitution if they carry electron-withdrawing substituents, so normally nucleophilic aromatic rings also become electrophilic if they have the right substituents.

Dichlorocarbene is the reactive intermediate with chemical formula CCl2. Although this chemical species has not been isolated, it is a common intermediate in organic chemistry, being generated from chloroform. This bent diamagnetic molecule rapidly inserts into other bonds.

The Dakin–West reaction is a chemical reaction that transforms an amino-acid into a keto-amide using an acid anhydride and a base, typically pyridine. It is named for Henry Drysdale Dakin (1880–1952) and Randolph West (1890–1949). In 2016 Schreiner and coworkers reported the first asymmetric variant of this reaction employing short oligopeptides as catalysts.

The Strecker amino acid synthesis, also known simply as the Strecker synthesis, is a method for the synthesis of amino acids by the reaction of an aldehyde with ammonia in the presence of potassium cyanide. The condensation reaction yields an α-aminonitrile, which is subsequently hydrolyzed to give the desired amino acid. The method is used commercially for the production of racemic methionine from methional.

Ethylamine, also known as ethanamine, is an organic compound with the formula CH3CH2NH2. This colourless gas has a strong ammonia-like odor. It condenses just below room temperature to a liquid miscible with virtually all solvents. It is a nucleophilic base, as is typical for amines. Ethylamine is widely used in chemical industry and organic synthesis.

In organic chemistry, the Paal–Knorr Synthesis is a reaction used to synthesize substituted furans, pyrroles, or thiophenes from 1,4-diketones. It is a synthetically valuable method for obtaining substituted furans and pyrroles, which are common structural components of many natural products. It was initially reported independently by German chemists Carl Paal and Ludwig Knorr in 1884 as a method for the preparation of furans, and has been adapted for pyrroles and thiophenes. Although the Paal–Knorr synthesis has seen widespread use, the mechanism wasn't fully understood until it was elucidated by V. Amarnath et al. in the 1990s.

<span class="mw-page-title-main">ANRORC mechanism</span> Reaction mechanism in ring systems

The ANRORC mechanism in organic chemistry describes a special type of substitution reaction. ANRORC stands for Addition of the Nucleophile, Ring Opening, and Ring Closure in nucleophilic attack on ring systems and it helps to explain product formation and distribution in some nucleophilic substitutions especially in heterocyclic compounds. It is widely used in medicinal chemistry.

<span class="mw-page-title-main">Lithium bis(trimethylsilyl)amide</span> Chemical compound

Lithium bis(trimethylsilyl)amide is a lithiated organosilicon compound with the formula LiN(Si(CH3)3)2. It is commonly abbreviated as LiHMDS or Li(HMDS) (lithium hexamethyldisilazide - a reference to its conjugate acid HMDS) and is primarily used as a strong non-nucleophilic base and as a ligand. Like many lithium reagents, it has a tendency to aggregate and will form a cyclic trimer in the absence of coordinating species.

<span class="mw-page-title-main">Erlenmeyer–Plöchl azlactone and amino-acid synthesis</span>

The Erlenmeyer–Plöchl azlactone and amino acid synthesis, named after Friedrich Gustav Carl Emil Erlenmeyer who partly discovered the reaction, is a series of chemical reactions which transform an N-acyl glycine to various other amino acids via an oxazolone.

The Chichibabin reaction is a method for producing 2-aminopyridine derivatives by the reaction of pyridine with sodium amide. It was reported by Aleksei Chichibabin in 1914. The following is the overall form of the general reaction:

The Chichibabin pyridine synthesis is a method for synthesizing pyridine rings. The reaction involves the condensation reaction of aldehydes, ketones, α,β-Unsaturated carbonyl compounds, or any combination of the above, with ammonia. It was reported by Aleksei Chichibabin in 1924. Methyl-substituted pyridines, which show widespread uses among multiple fields of applied chemistry, are prepared by this methodology.

<span class="mw-page-title-main">2-Aminopyridine</span> Chemical compound

2-Aminopyridine is an organic compound with the formula H2NC5H4N. It is one of three isomeric aminopyridines. It is a colourless solid that is used in the production of the drugs piroxicam, sulfapyridine, tenoxicam, and tripelennamine. It is produced by the reaction of sodium amide with pyridine, the Chichibabin reaction.

<span class="mw-page-title-main">3-Aminopyridine</span> Chemical compound

3-Aminopyridine is an aminopyridine. It is a colorless solid.

The Boger pyridine synthesis is a cycloaddition approach to the formation of pyridines named after its inventor Dale L. Boger, who first reported it in 1981. The reaction is a form of inverse-electron demand Diels-Alder reaction in which an enamine reacts with a 1,2,4-triazine to form the pyridine nucleus. The reaction is especially useful for accessing pyridines that would be difficult or impossible to access via other methods and has been used in the total synthesis of several complicated natural products.

References

  1. Handbook of Heterocyclic Chemistry, (2010) y, Alan R. Katritzky,Christopher A. Ramsden,J. Joule,Viktor V. Zhdankin
  2. Heterocyclic Chemistry, (2001) Malcolm Sainsbury
  3. Levine, R.; Leake, W. W. (1955). "Rearrangement in the Reaction of 3-Bromopyridine with Sodium Amide and Sodioacetophenone". Science. 121 (3152): 780. Bibcode:1955Sci...121..780L. doi:10.1126/science.121.3152.780. PMID   17773207.
  4. Zoltewicz, John A.; Nisi, Carlo (1969). "Trapping of 3,4-Pyridyne by Thiomethoxide Ion in Ammonia". The Journal of Organic Chemistry. 34 (3): 765–766. doi:10.1021/jo01255a072.
  5. Kramer, Jerry; Berry, R. Stephen (1972). "Gaseous 3,4-Pyridyne and the Formation of Diazabiphenylene". Journal of the American Chemical Society. 94 (24): 8336–8347. doi:10.1021/ja00779a010.
  6. Synthesis and Diels-Alder reactions of 1,3-dimethyl-4-(phenylsulfonyl)-4H-furo[3,4-b]indole. A new annulation strategy for the construction of ellipticine and isoellipticine Gordon W. Gribble, Mark G. Saulnier, Mukund P. Sibi, and Judy A. Obaza-Nutaitis The Journal of Organic Chemistry 1984 49 (23), 4518-4523 doi : 10.1021/jo00197a039
  7. Total syntheses of ellipticine alkaloids and their amino analogues Original Research Article Tetrahedron, Volume 48, Issue 48, 27 November 1992, Pages 10645-10654 Chin-Kang Sha, Jeng-Fenn Yang doi : 10.1016/S0040-4020(01)88360-5
  8. A Five-Step Synthesis of (S)-Macrostomine from (S)-Nicotine Monica F. Enamorado, Pauline W. Ondachi, and Daniel L. Comins Organic Letters 2010 12 (20), 4513-4515 doi : 10.1021/ol101887b