Zincke aldehyde

Last updated

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 (as opposed to primary 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. [1] [2] [3] The use of cyanogen bromide for pyridine activation was independently reported by W. König: [4]

Zincke aldehydes Zincke-Aldehyde.svg
Zincke aldehydes

The synthesis and utility of Zincke aldehydes has been reviewed. [5] [6] [7]

A variation of the Zincke reaction has been applied in the synthesis of novel indoles: [8]

Zincke aldehydes Kearney 2006 ZinckeAldehydeIndoleApplication.svg
Zincke aldehydes Kearney 2006

with cyanogen bromide mediated pyridine activation (König method). [4]

More recently, an interesting rearrangement of Zincke aldehydes to Z-unsaturated amides was discovered serendipitously while trying to do an intramolecular Diels-Alder reaction. [9] The rearrangement gives the Z-product stereospecifically. In a follow-up paper, allylic amines were used and gave products of a rearrangement / intramolecular Diels-Alder cascade. [10] Mechanistic details were also discussed, however further investigations in collaboration with the Houk group revealed an unusual and unexpected mechanism based on computational studies. The new mechanism involves formation of a vinyl ketene. [11]

Rearrangement of Zincke aldehydes to Z-unsaturated amides Zincke aldehyde rearrangement.gif
Rearrangement of Zincke aldehydes to Z-unsaturated amides
New ketene-based mechanism for rearrangement of Zincke aldehydes Zincke RAR mechanism.gif
New ketene-based mechanism for rearrangement of Zincke aldehydes

The Vanderwal group has also reported the synthesis of 4-stannyldienals from Zincke aldehydes by addition of tributylstannyl anion and quenching with acetyl chloride. [12] The products are useful substrates for Stille cross-coupling reactions to give interesting polyene structures.

Formation of stannyldienals from Zincke aldehydes Stannyldienal synthesis.gif
Formation of stannyldienals from Zincke aldehydes

In 2009, the Vanderwal group reported another interesting rearrangement of Zincke aldehydes. Tryptamine-derived Zincke aldehydes are heated with strong base to give the rearranged enal as shown below. This reaction was the key step in their total synthesis of norfluorocurarine, a Strychnos alkaloid. [13] This strategy was also employed in a short synthesis of strychnine, becoming the shortest synthesis of strychnine reported to date at only six linear steps. [14] This works has been highlighted on the blog Totally Synthetic.

Formal cycloaddition of Zincke aldehydes Synthesis of norfluorocurarine.gif
Formal cycloaddition of Zincke aldehydes
Vanderwal synthesis of strychnine Strychnine graphic.gif
Vanderwal synthesis of strychnine

Also in 2009, the first reports of Zincke aldehydes undergoing a Pictet-Spengler reaction appeared from the group of Christian Marazano. [15] This reaction provided the tetrahydro-β-carboline or tetrahydroisoquinoline core present in many alkaloid natural products, and was applied to the construction of a known intermediate in a previous total synthesis.

N-Acyl Pictet-Spengler Reaction by Treatment of Tryptamine and Homoveratrylamine Derived Aminopentadienals with TFAA Pictet-Spengler reaction.gif
N-Acyl Pictet−Spengler Reaction by Treatment of Tryptamine and Homoveratrylamine Derived Aminopentadienals with TFAA

One drawback of the Zincke Aldehyde synthesis is the need for 2 equivalents of the amine in the initial pyridine ring opening reaction. This is of particular concern for the case of complex secondary amines required for natural product synthesis. The group of Marazano recently found an alternative synthesis by condensation onto a variety of glutaconaldehyde derivatives using TFA. This solution has greatly simplified the production and purification of complex Zincke aldehydes. [16]

Related Research Articles

<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.

An alkyne trimerisation is a [2+2+2] cycloaddition reaction in which three alkyne units react to form a benzene ring. The reaction requires a metal catalyst. The process is of historic interest as well as being applicable to organic synthesis. Being a cycloaddition reaction, it has high atom economy. Many variations have been developed, including cyclisation of mixtures of alkynes and alkenes as well as alkynes and nitriles.

<span class="mw-page-title-main">Hemiaminal</span> Organic compound or group with a hydroxyl and amine attached to the same carbon

In organic chemistry, a hemiaminal is a functional group or type of chemical compound that has a hydroxyl group and an amine attached to the same carbon atom: −C(OH)(NR2)−. R can be hydrogen or an alkyl group. Hemiaminals are intermediates in imine formation from an amine and a carbonyl by alkylimino-de-oxo-bisubstitution. Hemiaminals can be viewed as a blend of aminals and geminal diol. They are a special case of amino alcohols.

In organic chemistry, the Knoevenagel condensation reaction is a type of chemical reaction named after German chemist Emil Knoevenagel. It is a modification of the aldol condensation.

<span class="mw-page-title-main">Claisen rearrangement</span> Chemical reaction

The Claisen rearrangement is a powerful carbon–carbon bond-forming chemical reaction discovered by Rainer Ludwig Claisen. The heating of an allyl vinyl ether will initiate a [3,3]-sigmatropic rearrangement to give a γ,δ-unsaturated carbonyl, driven by exergonically favored carbonyl CO bond formation.

<span class="mw-page-title-main">Iminium</span> Polyatomic ion of the form >C=N< and charge +1

In organic chemistry, an iminium cation is a polyatomic ion with the general structure [R1R2C=NR3R4]+. They are common in synthetic chemistry and biology.

<span class="mw-page-title-main">Wagner–Meerwein rearrangement</span> Organic reaction

A Wagner–Meerwein rearrangement is a class of carbocation 1,2-rearrangement reactions in which a hydrogen, alkyl or aryl group migrates from one carbon to a neighboring carbon. They can be described as cationic [1,2]-sigmatropic rearrangements, proceeding suprafacially and with stereochemical retention. As such, a Wagner–Meerwein shift is a thermally allowed pericyclic process with the Woodward-Hoffmann symbol [ω0s + σ2s]. They are usually facile, and in many cases, they can take place at temperatures as low as –120 °C. The reaction is named after the Russian chemist Yegor Yegorovich Vagner; he had German origin and published in German journals as Georg Wagner; and Hans Meerwein.

<span class="mw-page-title-main">Favorskii rearrangement</span> Chemical reaction

The Favorskii rearrangement is principally a rearrangement of cyclopropanones and α-halo ketones that leads to carboxylic acid derivatives. In the case of cyclic α-halo ketones, the Favorskii rearrangement constitutes a ring contraction. This rearrangement takes place in the presence of a base, sometimes hydroxide, to yield a carboxylic acid but most of the time either an alkoxide base or an amine to yield an ester or an amide, respectively. α,α'-Dihaloketones eliminate HX under the reaction conditions to give α,β-unsaturated carbonyl compounds.

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.

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

The Wolff rearrangement is a reaction in organic chemistry in which an α-diazocarbonyl compound is converted into a ketene by loss of dinitrogen with accompanying 1,2-rearrangement. The Wolff rearrangement yields a ketene as an intermediate product, which can undergo nucleophilic attack with weakly acidic nucleophiles such as water, alcohols, and amines, to generate carboxylic acid derivatives or undergo [2+2] cycloaddition reactions to form four-membered rings. The mechanism of the Wolff rearrangement has been the subject of debate since its first use. No single mechanism sufficiently describes the reaction, and there are often competing concerted and carbene-mediated pathways; for simplicity, only the textbook, concerted mechanism is shown below. The reaction was discovered by Ludwig Wolff in 1902. The Wolff rearrangement has great synthetic utility due to the accessibility of α-diazocarbonyl compounds, variety of reactions from the ketene intermediate, and stereochemical retention of the migrating group. However, the Wolff rearrangement has limitations due to the highly reactive nature of α-diazocarbonyl compounds, which can undergo a variety of competing reactions.

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

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.

<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.

The Willgerodt rearrangement or Willgerodt reaction is an organic reaction converting an aryl alkyl ketone, alkyne, or alkene to the corresponding amide by reaction with ammonium polysulfide, named after Conrad Willgerodt. The formation of the corresponding carboxylic acid is a side reaction resulting from hydrolysis of the amide. When the alkyl group is an aliphatic chain, multiple reactions take place with the amide group always ending up at the terminal end. The net effect is thus migration of the carbonyl group to the end of the chain and oxidation.

The Castro–Stephens coupling is a cross coupling reaction between a copper(I) acetylide and an aryl halide in pyridine, forming a disubstituted alkyne and a copper(I) halide.

The Debus–Radziszewski imidazole synthesis is a multi-component reaction used for the synthesis of imidazoles from a 1,2-dicarbonyl, an aldehyde, and ammonia or a primary amine. The method is used commercially to produce several imidazoles. The process is an example of a multicomponent 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.

Rearrangements, especially those that can participate in cascade reactions, such as the aza-Cope rearrangements, are of high practical as well as conceptual importance in organic chemistry, due to their ability to quickly build structural complexity out of simple starting materials. The aza-Cope rearrangements are examples of heteroatom versions of the Cope rearrangement, which is a [3,3]-sigmatropic rearrangement that shifts single and double bonds between two allylic components. In accordance with the Woodward-Hoffman rules, thermal aza-Cope rearrangements proceed suprafacially. Aza-Cope rearrangements are generally classified by the position of the nitrogen in the molecule :

The asymmetric addition of alkynylzinc compounds to aldehydes is an example of a Nef synthesis, a chemical reaction whereby a chiral propargyl alcohol is prepared from a terminal alkyne and an aldehyde. This alkynylation reaction is enantioselective and involves an alkynylzinc reagent rather than the sodium acetylide used by John Ulric Nef in his 1899 report of the synthetic approach. Propargyl alcohols are versatile precursors for the chirally-selective synthesis of natural products and pharmaceutical agents, making this asymmetric addition reaction of alkynylzinc compounds useful. For example, Erick Carreira used this approach in a total synthesis of the marine natural product leucascandrolide A, a bioactive metabolite of the calcareous sponge Leucascandra caveolata with cytotoxic and antifungal properties isolated in 1996.

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

2-Carboxybenzaldehyde is a chemical compound. It consists of a benzene ring, with an aldehyde and a carboxylic acid as substituents that are ortho to each other. The compound exhibits ring–chain tautomerism: the two substituents can react with each other to form 3-hydroxyphthalide, a cyclic lactol. This lactol reacts readily with Grignard reagents, forming alkyl- and aryl-substituted phthalides. Other benzo-fused heterocyclic compounds can be derived from 2-carboxybenzaldehyde, including isoindolinones and phthalazinones, with a variety of pharmacological properties, such as the antihistamine azelastine.

References

  1. Zincke, Th.; Heuser, G.; Moller, W. (1904). "Ueber Dinitrophenylpyridiniumchlorid und dessen Umwandlungsproducte". Justus Liebigs Annalen der Chemie (in German). 333 (2–3): 296–345. doi:10.1002/jlac.19043330212.
  2. Zincke, Th.; Heuser, G.; Moller, W. (1904). "Ueber Dinitrophenylpyridiniumchlorid und dessen Umwandlungsproducte". Justus Liebigs Annalen der Chemie (in German). 330 (2): 361–374. doi: 10.1002/jlac.19043300217 .
  3. Zincke, Th.; Weisspfenning, G. (1913). "Über Dinitrophenylisochinoliniumchlorid und dessen Umwandlungsprodukte". Justus Liebigs Annalen der Chemie (in German). 396 (1): 103–131. doi:10.1002/jlac.19133960107.
  4. 1 2 König, W. (1904). "Über eine neue, vom Pyridin derivierende Klasse von Farbstoffen". J. Prakt. Chem. (in German). 69 (1): 105–137. doi:10.1002/prac.19040690107.
  5. Becher, J. (1980). "Synthesis and Reactions of Glutaconaldehyde and 5-Amino-2,4-pentadienals". Synthesis . 1980 (8): 589–612. doi:10.1055/s-1980-29134.
  6. Becher, J.; Finsen, L.; Winckelmann, I. (1981). "Derivatives and reactions of glutaconaldehyde—XIII Regiospecific ring opening of 3-substituted pyridines". Tetrahedron . 37 (13): 2375–2378. doi: 10.1016/S0040-4020(01)88892-X .
  7. Cheng, W.-C.; Kurth, M. J. (2002). "The Zincke Reaction: A Review". Org. Prep. Proced. Int. 34 (6): 587–608. doi:10.1080/00304940209355784.
  8. Kearney, Aaron M.; Vanderwal, Christopher D. (2006). "Synthesis of Nitrogen Heterocycles by the Ring Opening of Pyridinium Salts". Angew. Chem. Int. Ed. 45 (46): 7803–7806. doi:10.1002/anie.200602996.
  9. Steinhardt, S. E.; Silverston, J. S.; Vanderwal, C. D. (2008). "Stereocontrolled Synthesis of Z-Dienes via an Unexpected Pericyclic Cascade Rearrangement of 5-Amino-2,4-pentadienals". J. Am. Chem. Soc. 130 (24): 7560–7561. doi:10.1021/ja8028125. PMID   18505251.
  10. Steinhardt, S. E.; Vanderwal, C. D. (2009). "Complex Polycyclic Lactams from Pericyclic Cascade Reactions of Zincke Aldehydes". J. Am. Chem. Soc. 131 (22): 7546–7547. doi:10.1021/ja902439f. PMID   19449870.
  11. Paton, R. S.; Steinhardt, S. E.; Vanderwal, C. D.; Houk, K. N. (2011). "Stereocontrolled Synthesis of Z-Dienes via an Unexpected Pericyclic Cascade Rearrangement of 5-Amino-2,4-pentadienals". J. Am. Chem. Soc. 133 (11): 3895–3905. doi:10.1021/ja107988b. PMID   21351736.
  12. Michels, T.; Rhee, J. U.; Vanderwal, C. D. (2008). "Synthesis of δ-Tributylstannyl Unsaturated Aldehydes from Pyridines". Org. Lett. 10 (21): 4787–4790. doi:10.1021/ol8020435. PMID   18817407.
  13. Martin, D. B. C.; Vanderwal, C. D. (2009). "Efficient Access to the Core of the Strychnos, Aspidosperma and Iboga Alkaloids. A Short Synthesis of Norfluorocurarine". J. Am. Chem. Soc. 131 (10): 3472–3473. doi:10.1021/ja900640v. PMID   19236094.
  14. Martin, D. B. C.; Vanderwal, C. D. (2011). "A Synthesis of Strychnine by a Longest Linear Sequence of Six Steps". Chem. Sci. 2 (4): 649. doi:10.1039/C1SC00009H.
  15. Nuhant, P.; Raikar, S. B.; Wypych, J.-C.; Delpech, B.; Marazano, C. (2009). "Enhancement of 5-Aminopenta-2,4-dienals Electrophilicity via Activation by O,N-Bistrifluoroacetylation. Application to an N-Acyl Pictet−Spengler Reaction". J. Org. Chem. 74 (24): 9413–21. doi:10.1021/jo9019545. PMID   19924881.
  16. Nguyen, T.; Peixoto, S.; Ouairy, C.; Nguyen, T.; Bénéchie, M.; Marazano, C.; Michel, P. (2010). "Simple and Convenient Method for the Synthesis of 2-Substituted Glutaconaldehyde Salts and 2-Substituted Glutaconaldehyde Derivatives". Synthesis . 2010 (1): 103–109. doi:10.1055/s-0029-1217105.