Pictet–Spengler reaction

Last updated

Contents

Pictet-Spengler reaction
Named after Amé Pictet
Theodor Spengler
Reaction typeRing-forming reaction
Identifiers
RSC ontology ID RXNO:0000059

The Pictet–Spengler reaction is a chemical reaction in which a β-arylethylamine undergoes condensation with an aldehyde or ketone followed by ring closure. The reaction was first discovered in 1911 by Amé Pictet and Theodor Spengler (22 February 1886 – 18 August 1965). [1] Traditionally, an acidic catalyst in protic solvent was employed with heating; [2] however, the reaction has been shown to work in aprotic media in superior yields and sometimes without acid catalysis. [3] The Pictet–Spengler reaction can be considered a special case of the Mannich reaction, which follows a similar reaction pathway. The driving force for this reaction is the electrophilicity of the iminium ion generated from the condensation of the aldehyde and amine under acid conditions. This explains the need for an acid catalyst in most cases, as the imine is not electrophilic enough for ring closure but the iminium ion is capable of undergoing the reaction.

The Pictet-Spengler reaction Pictet-Spengler Reaction Scheme.png
The Pictet–Spengler reaction

The Pictet–Spengler reaction is widespread in both industry and biosynthesis. It has remained an important reaction in the fields of alkaloid and organic synthesis since its inception, where it has been employed in the development of many beta-carbolines. Natural Pictet–Spengler reaction typically employ an enzyme, such as strictosidine synthase. Pictet–Spengler products can be isolated from many products initially derived from nature, including foodstuffs such as soy sauce and ketchup. In such cases it is common to find the amino acid tryptophan and various aldoses used as the biological feedstock.

Nucleophilic aromatic rings such as indole or pyrrole give products in high yields and mild conditions, while less nucleophilic aromatic rings such as a phenyl group give poorer yields or require higher temperatures and strong acid. The original Pictet–Spengler reaction was the reaction of phenethylamine and dimethoxymethane, catalysed by hydrochloric acid forming a tetrahydroisoquinoline.

The Pictet–Spengler reaction has been applied to solid-phase combinatorial chemistry with great success. [4] [5]

An analogous reaction with an aryl-β-ethanol is called oxa-Pictet–Spengler reaction. [6]

Reaction mechanism

The reaction mechanism occurs by initial formation of an iminium ion (2) followed by electrophilic addition at the 3-position, in accordance with the expected nucleophilicity of indoles, to give the spirocycle 3. After migration of the best migrating group, deprotonation gives the product (5).

The mechanism of the Pictet-Spengler reaction Pictet Spengler.png
The mechanism of the Pictet–Spengler reaction

Variations

Pictet–Spengler tetrahydroisoquinoline synthesis

Replacing an indole with a 3,4-dimethoxyphenyl group give the reaction named the Pictet–Spengler tetrahydroisoquinoline synthesis. Reaction conditions are generally harsher than the indole variant, and require refluxing conditions with strong acids like hydrochloric acid, trifluoroacetic acid or superacids. [7] [8]

The Pictet-Spengler isoquinoline synthesis Pictet-Spengler Isoquinoline Synthesis Scheme.png
The Pictet–Spengler isoquinoline synthesis

N-acyliminium ion Pictet–Spengler reaction

Instead of catalyzing the Pictet–Spengler cyclization with strong acid, one can acylate the iminium ion forming the intermediate N-acyliminium ion. The N-acyliminium ion is a very powerful electrophile and most aromatic ring systems will cyclize under mild conditions with good yields. [9]

The N-acyliminium Pictet-Spengler reaction N-acyliminium Pictet-Spengler reaction.svg
The N-acyliminium Pictet–Spengler reaction

Tadalafil is synthesized via the N-acyliminium Pictet–Spengler reaction. [10] This reaction can also be catalyzed by AuCl3 and AgOTf. [11]

Asymmetric Pictet–Spengler reaction

When the Pictet–Spengler reaction is performed with an aldehyde other than formaldehyde, a new chiral center is created. Several substrate- or auxiliary-controlled diastereoselective Pictet–Spengler reactions have been developed. [12] [13] Additionally, List et al. have published a chiral Brønsted acid that catalyzes asymmetric Pictet–Spengler reactions. [14]

Tryptophans: diastereocontrolled reaction
The reaction of enantiopure tryptophan or its short-chain alkylesters leads to 1,2,3,4-tetrahydro-β-carbolines in which a new chiral center at C-1 adopts either a cis or trans configuration towards the C-3 carboxyl group. The cis conduction is kinetically controlled, i.e. it is performed at lower temperatures. At higher temperatures the reaction becomes reversible and usually favours racemisation. 1,3-trans dominated products can be obtained with Nb-benzylated tryptophans, which are accessible by reductive amination. The benzyl group can be removed hydrogenolytically afterwards. As a rough rule, 13C NMR signals for C1 and C3 are downfield shifted in cis products relative to trans products (see steric compression effect). [3] [15]

See also

Related Research Articles

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

Isoquinoline is an individual chemical specimen - a heterocyclic aromatic organic compound - as well as the name of a family of many thousands of natural plant alkaloids, any one of which might be referred to as "an isoquinoline". It is a structural isomer of quinoline. Isoquinoline and quinoline are benzopyridines, which are composed of a benzene ring fused to a pyridine ring. In a broader sense, the term isoquinoline is used to make reference to isoquinoline derivatives. 1-Benzylisoquinoline is the structural backbone in many naturally occurring alkaloids such as papaverine. The isoquinoline ring in these natural compound derives from the aromatic amino acid tyrosine.

In organic chemistry, the Mannich reaction is a three-component organic reaction that involves the amino alkylation of an acidic proton next to a carbonyl functional group by formaldehyde and a primary or secondary amine or ammonia. The final product is a β-amino-carbonyl compound also known as a Mannich base. Reactions between aldimines and α-methylene carbonyls are also considered Mannich reactions because these imines form between amines and aldehydes. The reaction is named after Carl Mannich.

In retrosynthetic analysis, a synthon is a hypothetical unit within a target molecule that represents a potential starting reagent in the retroactive synthesis of that target molecule. The term was coined in 1967 by E. J. Corey. He noted in 1988 that the "word synthon has now come to be used to mean synthetic building block rather than retrosynthetic fragmentation structures". It was noted in 1998 that the phrase did not feature very prominently in Corey's 1981 book The Logic of Chemical Synthesis, as it was not included in the index. Because synthons are charged, when placed into a synthesis an uncharged form is found commercially instead of forming and using the potentially very unstable charged synthons.

<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.

The Bischler–Napieralski reaction is an intramolecular electrophilic aromatic substitution reaction that allows for the cyclization of β-arylethylamides or β-arylethylcarbamates. It was first discovered in 1893 by August Bischler and Bernard Napieralski, in affiliation with Basle Chemical Works and the University of Zurich. The reaction is most notably used in the synthesis of dihydroisoquinolines, which can be subsequently oxidized to isoquinolines.

<span class="mw-page-title-main">Biginelli reaction</span> Multicomponent chemical reaction

The Biginelli reaction is a multiple-component chemical reaction that creates 3,4-dihydropyrimidin-2(1H)-ones 4 from ethyl acetoacetate 1, an aryl aldehyde, and urea 3. It is named for the Italian chemist Pietro Biginelli.

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

The Petasis reaction is the multi-component reaction of an amine, a carbonyl, and a vinyl- or aryl-boronic acid to form substituted amines.

<span class="mw-page-title-main">Indole alkaloid</span> Class of alkaloids

Indole alkaloids are a class of alkaloids containing a structural moiety of indole; many indole alkaloids also include isoprene groups and are thus called terpene indole or secologanin tryptamine alkaloids. Containing more than 4100 known different compounds, it is one of the largest classes of alkaloids. Many of them possess significant physiological activity and some of them are used in medicine. The amino acid tryptophan is the biochemical precursor of indole alkaloids.

<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:

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

Spirotryprostatin B is an indolic alkaloid found in the Aspergillus fumigatus fungus that belongs to a class of naturally occurring 2,5-diketopiperazines. Spirotryprostatin B and several other indolic alkaloids have been found to have anti-mitotic properties, and as such they have become of great interest as anti-cancer drugs. Because of this, the total syntheses of these compounds is a major pursuit of organic chemists, and a number of different syntheses have been published in the chemical literature.

Strictosidine synthase (EC 4.3.3.2) is an enzyme in alkaloid biosynthesis that catalyses the condensation of tryptamine with secologanin to form strictosidine in a formal Pictet–Spengler reaction:

<span class="mw-page-title-main">Morten P. Meldal</span> Danish chemist (born 1954)

Morten Peter Meldal is a Danish chemist and Nobel laureate. He is a professor of chemistry at the University of Copenhagen in Copenhagen, Denmark. He is best known for developing the CuAAC-click reaction, concurrently with but independent of Valery V. Fokin and K. Barry Sharpless.

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

Indole is an aromatic, heterocyclic, organic compound with the formula C8H7N. It has a bicyclic structure, consisting of a six-membered benzene ring fused to a five-membered pyrrole ring. Indole is widely distributed in the natural environment and can be produced by a variety of bacteria. As an intercellular signal molecule, indole regulates various aspects of bacterial physiology, including spore formation, plasmid stability, resistance to drugs, biofilm formation, and virulence. The amino acid tryptophan is an indole derivative and the precursor of the neurotransmitter serotonin.

<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.

The imine Diels–Alder reaction involves the transformation of all-carbon dienes and imine dienophiles into tetrahydropyridines.

Electrophilic aromatic substitution is an organic reaction in which an atom that is attached to an aromatic system is replaced by an electrophile. Some of the most important electrophilic aromatic substitutions are aromatic nitration, aromatic halogenation, aromatic sulfonation, alkylation and acylation Friedel–Crafts reaction.

<span class="mw-page-title-main">Hydrogen-bond catalysis</span>

Hydrogen-bond catalysis is a type of organocatalysis that relies on use of hydrogen bonding interactions to accelerate and control organic reactions. In biological systems, hydrogen bonding plays a key role in many enzymatic reactions, both in orienting the substrate molecules and lowering barriers to reaction. However, chemists have only recently attempted to harness the power of using hydrogen bonds to perform catalysis, and the field is relatively undeveloped compared to research in Lewis acid catalysis.

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 :

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

Fascaplysin is a marine alkaloid based on 12H-pyrido[1–2-a:3,4-b′]diindole ring system. It was first isolated as a red pigment from the marine sponge Fascaplysinopsis bergquist collected in the South Pacific near Fiji in 1988. Fascaplysin possesses a broad range of in vitro biological activities including analgesic, antimicrobial, antifungal, antiviral, antimalarial, anti-angiogenic, and antiproliferative activity against numerous cancer cell lines.

References

  1. Pictet, A.; Spengler, T. (1911). "Über die Bildung von Isochinolin-derivaten durch Einwirkung von Methylal auf Phenyl-äthylamin, Phenyl-alanin und Tyrosin". Berichte der Deutschen Chemischen Gesellschaft. 44 (3): 2030–2036. doi:10.1002/cber.19110440309.
  2. Whaley, W. M.; Govindachari, T. R. (1951). "The Pictet-Spengler synthesis of tetrahydroisoquinolines and related compounds". Org. React. 6: 74.
  3. 1 2 Cox, E. D.; Cook, J. M. (1995). "The Pictet-Spengler condensation: a new direction for an old reaction". Chemical Reviews. 95 (6): 1797–1842. doi:10.1021/cr00038a004.
  4. Nielsen, T. E.; Diness, F.; Meldal, M. (2003). "Solid-Phase Synthesis of Pyrroloisoquinolines via the Intramolecular N-Acyliminium Pictet-Spengler Reaction". Curr. Opin. Drug Discov. Dev. 6 (6): 801–814. PMID   14758752.
  5. Nielsen, T. E.; Meldal, M. (2005). "Solid-Phase Synthesis of Pyrroloisoquinolines via the Intramolecular N-Acyliminium Pictet-Spengler Reaction". J. Comb. Chem. 7 (4): 599–610. doi:10.1021/cc050008a. PMID   16004504.
  6. Larghi, E. L.; Kaufman, T. S. (2006). "The oxa-Pictet-Spengler Cyclization. Synthesis of Isochromanes and Related Pyran-Type Heterocycles". Synthesis (2): 187–210. doi:10.1055/s-2005-918502.
  7. Yokoyama, Akihiro; Ohwada, Tomohiko; Shudo, Koichi (1999). "Prototype Pictet−Spengler Reactions Catalyzed by Superacids. Involvement of Dicationic Superelectrophiles". J. Org. Chem. 64 (2): 611–617. doi:10.1021/jo982019e.
  8. Quevedo, R.; Baquero, E.; Rodriguez, M. (2010). "Regioselectivity in isoquinoline alkaloid Synthesis". Tetrahedron Letters. 51 (13): 1774–1778. doi:10.1016/j.tetlet.2010.01.115.
  9. Maryanoff, B. E.; Zhang, H.-C.; Cohen, J. H.; Turchi, I. J.; Maryanoff, C. A. (2004). "Cyclizations of N-acyliminium ions". Chem. Rev. 104 (3): 1431–1628. doi:10.1021/cr0306182. PMID   15008627.
  10. Bonnet, D.; Ganesan, A. (2002). "Solid-Phase Synthesis of Tetrahydro-β-carbolinehydantoins via the N-Acyliminium Pictet-Spengler Reaction and Cyclative Cleavage". J. Comb. Chem. 4 (6): 546–548. doi:10.1021/cc020026h. PMID   12425597.
  11. Youn, S. W. (2006). "Development of the Pictet-Spengler Reaction Catalyzed by AuCl3/AgOTf". J. Org. Chem. 71 (6): 2521–2523. doi:10.1021/jo0524775. PMID   16526809.
  12. Gremmen, C.; Willemse, B.; Wanner, M. J.; Koomen, G.-J. (2000). "Enantiopure Tetrahydro-β-carbolines via Pictet-Spengler Reactions with N-Sulfinyl Tryptamines". Org. Lett. 2 (13): 1955–1958. doi:10.1021/ol006034t. PMID   10891200.
  13. a) The intermolecular Pictet-Spengler condensation with chiral carbonyl derivatives in the stereoselective syntheses of optically-active isoquinoline and indole alkaloids Enrique L. Larghi, Marcela Amongero, Andrea B. J. Bracca, and Teodoro S. Kaufman Arkivoc (RL-1554K) pp 98–153 2005 (Online Review [ permanent dead link ]); b) Teodoro S. Kaufman "Synthesis of Optically-Active Isoquinoline and Indole Alkaloids Employing the Pictet-Spengler Condensation with Removable Chiral Auxiliaries Bound to Nitrogen". in "New Methods for the Asymmetric Synthesis of Nitrogen Heterocycles"; Ed.: J. L. Vicario. ISBN   81-7736-278-X. Research SignPost, Trivandrum, India. 2005. Chapter 4, pp. 99–147.
  14. Seayad, J.; Seayad, A. M.; List, B. (2006). "Catalytic Asymmetric Pictet-Spengler Reaction". J. Am. Chem. Soc. 128 (4): 1086–1087. doi:10.1021/ja057444l. PMID   16433519.
  15. Ungemach, F.; Soerens, D.; Weber, R.; Dipierro, M.; Campos, O.; Mokry, P.; Cook, J. M.; Silverton, J. V. (1980). "General method for the assignment of stereochemistry of 1,3-disubstituted 1,2,3,4-tetrahydro-β-carbolines by carbon-13 spectroscopy". J. Am. Chem. Soc. 102 (23): 6976–6984. doi:10.1021/ja00543a012.