Zincke reaction

Last updated
Zincke reaction
Named after Theodor Zincke
Reaction type Coupling reaction

The Zincke reaction is an organic reaction, named after Theodor Zincke, in which a pyridine is transformed into a pyridinium salt by reaction with 2,4-dinitro-chlorobenzene and a primary amine. [1] [2] [3]

Contents

The Zincke reaction Zincke Reaction.svg
The Zincke reaction

The Zincke reaction should not be confused with the Zincke-Suhl reaction or the Zincke nitration. Furthermore, the Zincke reaction has nothing to do with the chemical element zinc.

Reaction mechanism

The first reaction is the formation of the N-2,4-dinitrophenyl-pyridinium salt (2). This salt is typically isolated and purified by recrystallization.

The formation of the DNP-pyridinium salt Zincke Reaction Mechanism1.png
The formation of the DNP-pyridinium salt

Upon heating a primary amine with the N-2,4-dinitrophenyl-pyridinium salt (2), the addition of the amine leads to the opening of the pyridinium ring. A second addition of amine leads to the displacement of 2,4-dinitroaniline (5) and formation of the König salt [4] (6a and 6b). The trans-cis-trans isomer of the König salt (6a) can react by either sigmatropic rearrangement or nucleophilic addition of a zwitterionic intermediate to give cyclized intermediate (7). [5] This has been suggested to be the rate-determining step. [6] [7] After proton transfer and amine elimination, the desired pyridinium ion (9) is formed.

The mechanism of the Zincke reaction Zincke Reaction Mechanism2.png
The mechanism of the Zincke reaction

This mechanism can be referred to as an instance of the ANRORC mechanism: nucleophilic addition (AN), ring opening and ring closing.

Applications

In one solid-phase synthesis application, the amine is covalently attached to Wang resin. [8]

The Zincke reaction ZinckeReactionSolidState.png
The Zincke reaction

Another example is the synthesis of a chiral isoquinolinium salt. [9]

The Zincke reaction Zinckereactionchiral.png
The Zincke reaction

Zincke aldehydes

With secondary amines and not primary amines the Zincke reaction takes on a different shape forming so-called Zincke aldehydes in which the pyridine ring is ring-opened with the terminal iminium group hydrolyzed to an aldehyde: [10]

Zincke aldehydes Zincke-Aldehyde.svg
Zincke aldehydes

This variation has been applied in the synthesis of novel indoles: [11]

Zincke aldehydes Kearney 2006 ZinckeAldehydeIndoleApplication.svg
Zincke aldehydes Kearney 2006

with cyanogen bromide mediated pyridine activation.

2007 rediscovery

In 2006 and again in 2007 the Zincke reaction was rediscovered by a research group from Japan [12] and a group from the USA. [13] Both groups claimed the synthesis of a 12 membered diazaannulene (structure 1) from an N-aryl pyridinium chloride and an amine, an aniline in the case of the Japanese group (depicted below) and an aliphatic amine (anticipating surfactant properties) in the case of the American group.

DiazaAnnulene.svg

In a letter to Angewandte Chemie, the German chemist Manfred Christl [14] pointed out not only that the alleged new chemistry was in fact 100-year-old Zincke chemistry but also that the proposed structure for the reaction product was not the 12 membered ring but the 6 membered pyridinium salt (structure 2). Initially both groups conceded that they had ignored existing literature on Zincke but held on to the annulene structure based on their electrospray ionization (ESI) results which according to them clearly showed dimer. In his letter Christl remarked that in ESI measurements association of molecules is a common phenomenon. In addition, he noted similarities in melting point and NMR spectroscopy.

As of December 2007 the Japanese group retracted its paper in Organic Letters due to uncertainties regarding what products are formed in the reaction described and the US group added a correction to theirs in the Angewandte Chemie stating they wish(ed) to alter the proposed structure of (the) annulene. [15] The issue did receive some media coverage: [16] [17]

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">Diels–Alder reaction</span> Chemical reaction

In organic chemistry, the Diels–Alder reaction is a chemical reaction between a conjugated diene and a substituted alkene, commonly termed the dienophile, to form a substituted cyclohexene derivative. It is the prototypical example of a pericyclic reaction with a concerted mechanism. More specifically, it is classified as a thermally-allowed [4+2] cycloaddition with Woodward–Hoffmann symbol [π4s + π2s]. It was first described by Otto Diels and Kurt Alder in 1928. For the discovery of this reaction, they were awarded the Nobel Prize in Chemistry in 1950. Through the simultaneous construction of two new carbon–carbon bonds, the Diels–Alder reaction provides a reliable way to form six-membered rings with good control over the regio- and stereochemical outcomes. Consequently, it has served as a powerful and widely applied tool for the introduction of chemical complexity in the synthesis of natural products and new materials. The underlying concept has also been applied to π-systems involving heteroatoms, such as carbonyls and imines, which furnish the corresponding heterocycles; this variant is known as the hetero-Diels–Alder reaction. The reaction has also been generalized to other ring sizes, although none of these generalizations have matched the formation of six-membered rings in terms of scope or versatility. Because of the negative values of ΔH° and ΔS° for a typical Diels–Alder reaction, the microscopic reverse of a Diels–Alder reaction becomes favorable at high temperatures, although this is of synthetic importance for only a limited range of Diels-Alder adducts, generally with some special structural features; this reverse reaction is known as the retro-Diels–Alder reaction.

<span class="mw-page-title-main">Benzoin condensation</span> Reaction between two aromatic aldehydes

The benzoin addition is an addition reaction involving two aldehydes. The reaction generally occurs between aromatic aldehydes or glyoxals, and results in formation of an acyloin. In the classic example, benzaldehyde is converted to benzoin.

The Fritsch–Buttenberg–Wiechell rearrangement, named for Paul Ernst Moritz Fritsch (1859–1913), Wilhelm Paul Buttenberg, and Heinrich G. Wiechell, is a chemical reaction whereby a 1,1-diaryl-2-bromo-alkene rearranges to a 1,2-diaryl-alkyne by reaction with a strong base such as an alkoxide.

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

Corannulene is a polycyclic aromatic hydrocarbon with chemical formula C20H10. The molecule consists of a cyclopentane ring fused with 5 benzene rings, so another name for it is [5]circulene. It is of scientific interest because it is a geodesic polyarene and can be considered a fragment of buckminsterfullerene. Due to this connection and also its bowl shape, corannulene is also known as a buckybowl. Buckybowls are fragments of buckyballs. Corannulene exhibits a bowl-to-bowl inversion with an inversion barrier of 10.2 kcal/mol (42.7 kJ/mol) at −64 °C.

<span class="mw-page-title-main">Atropisomer</span>

Atropisomers are stereoisomers arising because of hindered rotation about a single bond, where energy differences due to steric strain or other contributors create a barrier to rotation that is high enough to allow for isolation of individual conformers. They occur naturally and are important in pharmaceutical design. When the substituents are achiral, these conformers are enantiomers (atropoenantiomers), showing axial chirality; otherwise they are diastereomers (atropodiastereomers).

<span class="mw-page-title-main">Wilhelm Rudolph Fittig</span> German chemist (1835–1910)

Wilhelm Rudolph Fittig was a German chemist. He discovered the pinacol coupling reaction, mesitylene, diacetyl and biphenyl. Fittig studied the action of sodium on ketones and hydrocarbons. He discovered the Fittig reaction or Wurtz–Fittig reaction for the synthesis of alkylbenzenes, he proposed a diketone structure for benzoquinone and isolated phenanthrene from coal tar. He discovered and synthesized the first lactones and investigated structures of piperine naphthalene and fluorene.

The Lossen rearrangement is the conversion of a hydroxamate ester to an isocyanate. Typically O-acyl, sulfonyl, or phosphoryl O-derivative are employed. The isocyanate can be used further to generate ureas in the presence of amines or generate amines in the presence of H2O.

The Meerwein–Ponndorf–Verley (MPV) reduction in organic chemistry is the reduction of ketones and aldehydes to their corresponding alcohols utilizing aluminium alkoxide catalysis in the presence of a sacrificial alcohol. The advantages of the MPV reduction lie in its high chemoselectivity, and its use of a cheap environmentally friendly metal catalyst.

<span class="mw-page-title-main">Organocatalysis</span> Method in organic chemistry

In organic chemistry, organocatalysis is a form of catalysis in which the rate of a chemical reaction is increased by an organic catalyst. This "organocatalyst" consists of carbon, hydrogen, sulfur and other nonmetal elements found in organic compounds. Because of their similarity in composition and description, they are often mistaken as a misnomer for enzymes due to their comparable effects on reaction rates and forms of catalysis involved.

<span class="mw-page-title-main">Zincke aldehyde</span>

Zincke aldehydes, or 5-aminopenta-2,4-dienals, are the product of the reaction of a pyridinium salt with two equivalents of any secondary amine, followed by basic hydrolysis. Using secondary amines the Zincke reaction takes on a different shape forming Zincke aldehydes in which the pyridine ring is ring-opened with the terminal iminium group hydrolyzed to an aldehyde. The use of the dinitrophenyl group for pyridine activation was first reported by Theodor Zincke. The use of cyanogen bromide for pyridine activation was independently reported by W. König:

The total synthesis of quinine, a naturally-occurring antimalarial drug, was developed over a 150-year period. The development of synthetic quinine is considered a milestone in organic chemistry although it has never been produced industrially as a substitute for natural occurring quinine. The subject has also been attended with some controversy: Gilbert Stork published the first stereoselective total synthesis of quinine in 2001, meanwhile shedding doubt on the earlier claim by Robert Burns Woodward and William Doering in 1944, claiming that the final steps required to convert their last synthetic intermediate, quinotoxine, into quinine would not have worked had Woodward and Doering attempted to perform the experiment. A 2001 editorial published in Chemical & Engineering News sided with Stork, but the controversy was eventually laid to rest once and for all when Williams and coworkers successfully repeated Woodward's proposed conversion of quinotoxine to quinine in 2007.

<span class="mw-page-title-main">Diels–Reese reaction</span>

The Diels–Reese Reaction is a reaction between hydrazobenzene and dimethyl acetylenedicarboxylate first reported in 1934 by Otto Diels and Johannes Reese. Later work by others extended the reaction scope to include substituted hydrazobenzenes. The exact mechanism is not known. By changing the acidic or basic nature of the solvent, the reaction gives different products. With acetic acid as solvent (acidic), the reaction gives an diphenylpyrazolone. With xylene as solvent (neutral), the reaction gives an indole. With pyridine as solvent (basic), the reaction gives a carbomethoxyquinoline which can be degraded to a dihydroquinoline.

The Glaser coupling is a type of coupling reaction. It is by far the oldest acetylenic coupling and is based on cuprous salts like copper(I) chloride or copper(I) bromide and an additional oxidant like oxygen. The base in its original scope is ammonia. The solvent is water or an alcohol. The reaction was first reported by Carl Andreas Glaser in 1869. He suggested the following process for his way to diphenylbutadiyne:

Asymmetric hydrogenation is a chemical reaction that adds two atoms of hydrogen to a target (substrate) molecule with three-dimensional spatial selectivity. Critically, this selectivity does not come from the target molecule itself, but from other reagents or catalysts present in the reaction. This allows spatial information to transfer from one molecule to the target, forming the product as a single enantiomer. The chiral information is most commonly contained in a catalyst and, in this case, the information in a single molecule of catalyst may be transferred to many substrate molecules, amplifying the amount of chiral information present. Similar processes occur in nature, where a chiral molecule like an enzyme can catalyse the introduction of a chiral centre to give a product as a single enantiomer, such as amino acids, that a cell needs to function. By imitating this process, chemists can generate many novel synthetic molecules that interact with biological systems in specific ways, leading to new pharmaceutical agents and agrochemicals. The importance of asymmetric hydrogenation in both academia and industry contributed to two of its pioneers — William Standish Knowles and Ryōji Noyori — being awarded one half of the 2001 Nobel Prize in Chemistry.

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

HATU is a reagent used in peptide coupling chemistry to generate an active ester from a carboxylic acid. HATU is used along with Hünig's base, or triethylamine to form amide bonds. Typically DMF is used as solvent, although other polar aprotic solvents can also be used.

The Parikh–Doering oxidation is an oxidation reaction that transforms primary and secondary alcohols into aldehydes and ketones, respectively. The procedure uses dimethyl sulfoxide (DMSO) as the oxidant and the solvent, activated by the sulfur trioxide pyridine complex (SO3•C5H5N) in the presence of triethylamine or diisopropylethylamine as base. Dichloromethane is frequently used as a cosolvent for the reaction.

<span class="mw-page-title-main">Strychnine total synthesis</span>

Strychnine total synthesis in chemistry describes the total synthesis of the complex biomolecule strychnine. The first reported method by the group of Robert Burns Woodward in 1954 is considered a classic in this research field.

Carbene dimerization is a type of organic reaction in which two carbene or carbenoid precursors react in a formal dimerization to an alkene. This reaction is often considered an unwanted side-reaction but it is also investigated as a synthetic tool. In this reaction type either the two carbenic intermediates react or a carbenic intermediate reacts with a carbene precursor. An early pioneer was Christoph Grundmann reporting on a carbene dimerisation in 1938. In the domain of persistent carbenes the Wanzlick equilibrium describes an equilibrium between a carbene and its alkene.

The Kröhnke pyridine synthesis is reaction in organic synthesis between α-pyridinium methyl ketone salts and α, β-unsaturated carbonyl compounds used to generate highly functionalized pyridines. Pyridines occur widely in natural and synthetic products, so there is wide interest in routes for their synthesis. The method is named after Fritz Kröhnke.

References

  1. Zincke, Th.; Heuser, G.; Moller, W. (1904). "Ueber Dinitrophenylpyridiniumchlorid und dessen Umwandlungsproducte". Justus Liebigs Annalen der Chemie . 333 (2–3): 296–345. doi:10.1002/jlac.19043330212.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  2. Zincke, Th.; Heuser, G.; Moller, W. (1904). "Ueber Dinitrophenylpyridiniumchlorid und dessen Umwandlungsproducte". Justus Liebigs Annalen der Chemie . 330 (2): 361–374. doi: 10.1002/jlac.19043300217 .{{cite journal}}: CS1 maint: multiple names: authors list (link)
  3. Zincke, T. H.; Weisspfenning, G. (1913). "Über Dinitrophenylisochinoliniumchlorid und dessen Umwandlungsprodukte". Justus Liebigs Annalen der Chemie . 396 (1): 103–131. doi:10.1002/jlac.19133960107.
  4. König, W. (1904). "Über eine neue, vom Pyridin derivierende Klasse von Farbstoffen". Journal für Praktische Chemie . 69 (1): 105–137. doi:10.1002/prac.19040690107.
  5. Kunugi, S.; Okubo, T.; Ise, N. (1976). "A study on the mechanism of the reaction of N-(2,4-dinitrophenyl)-3-carbamoylpyridinium chloride with amines and amino acids with reference to effect of polyelectrolyte addition". Journal of the American Chemical Society . 98 (1): 2282–2287. doi:10.1021/ja00424a047. PMID   1254864.
  6. Marvell, E. N.; Caple, G.; Shahidi, I. J. Am. Chem. Soc. 1970, 92, 5641-5645. ( doi : 10.1021/ja00722a016)
  7. Marvell, E. N.; Shahidi, I. J. Am. Chem. Soc. 1970, 92, 5646-5649. ( doi : 10.1021/ja00722a017)
  8. "The Solid-Phase Zincke Reaction: Preparation of -Hydroxy Pyridinium Salts in the Search for CFTR Activation" Eda, M.; Kurth, M. J.; Nantz, M. H. J. Org. Chem. 2000, 65(17), 5131 - 5135. ( doi : 10.1021/jo0001636)
  9. New Chiral Isoquinolinium Salt Derivatives from Chiral Primary Amines via Zincke Reaction Denis Barbier, Christian Marazano, Bhupesh C. Das, and Pierre Potier J. Org. Chem.; 1996; 61(26) pp 9596 - 9598; (Note) doi : 10.1021/jo961539b
  10. T. Zincke, W. Wurker, Justus Liebigs Ann. Chem. Ueber Dinitrophenylpyridiniumchlorid und dessen Umwandlungsproducte, 1904, 338, 107 – 141; doi : 10.1002/jlac.19043380107
  11. Synthesis of Nitrogen Heterocycles by the Ring Opening of Pyridinium Salts Aaron M. Kearney, Christopher D. Vanderwal Angew. Chem. Int. Ed. 2006, 45, 7803 –7806 doi : 10.1002/anie.200602996
  12. One-Pot Synthesis of N-Substituted Diaza[12]annulenes Yamaguchi, I.; Gobara, Y.; Sato, M. Org. Lett.; (Letter); 2006; 8(19); 4279-4281. doi : 10.1021/ol061585q
  13. [12]Annulene Gemini Surfactants: Structure and Self-Assembly Lei Shi, Dan Lundberg, Djamaladdin G. Musaev, Fredric M. Menger Angewandte Chemie International Edition Volume 46, 2007 Issue 31 , Pages 5889 - 5891 doi : 10.1002/anie.200702140
  14. (in German) 1,7-Diaza[12]annulene Derivatives? 100-Year-Old Pyridinium Salts! (p 9152-9153)Manfred Christl Published Online: Nov 28 2007 5:47AM doi : 10.1002/anie.200704704
  15. Corrigendum [12]Annulene Gemini Surfactants: Structure and Self-Assembly Lei Shi, Dan Lundberg, Djamaladdin G. Musaev, Fredric M. Menger Angewandte Chemie International Edition Volume 46, 2007, Issue 48, Pages 9152 - 9153 doi : 10.1002/anie.200790248
  16. (in German)Ahnungslose Chemiker entdecken Verbindung zum zweiten Mal. Jens Lubbadeh Der Spiegel 6 December 2007 http://www.spiegel.de/wissenschaft/natur/0,1518,521646,00.html
  17. Sanderson, Katharine (4 December 2007). "Where have I seen that before? 103-year-old chemical reaction pops up again". Nature . doi: 10.1038/news.2007.341 .