Petasis reaction

Last updated

Contents

Petasis reaction
Named afterNicos A. Petasis
Reaction type Coupling reaction
Identifiers
Organic Chemistry Portal petasis-reaction
RSC ontology ID RXNO:0000232

The Petasis reaction (alternatively called the Petasis borono–Mannich (PBM) reaction) is the multi-component reaction of an amine, a carbonyl, and a vinyl- or aryl-boronic acid to form substituted amines.

The Petasis reaction Petasis Reaction Scheme.png
The Petasis reaction

Reported in 1993 by Nicos Petasis as a practical method towards the synthesis of a geometrically pure antifungal agent, naftifine. [1] [2] [3] In the Petasis reaction, the vinyl group of the organoboronic acid serves as the nucleophile. In comparison to other methods of generating allyl amines, the Petasis reaction tolerates a multifunctional scaffold, with a variety of amines and organoboronic acids as potential starting materials. Additionally, the reaction does not require anhydrous or inert conditions. As a mild, selective synthesis, the Petasis reaction is useful in generating α-amino acids, and is utilized in combinatorial chemistry and drug discovery. [4] [5] [6] [7]

Reaction scope and synthetic applications

The amine is condensed with the carbonyl followed by addition of the boronic acid . [1]

Alpha amino acid synthesis Alphaaminoacidsynthesis.png
Alpha amino acid synthesis

One of the most attractive features of the Petasis reaction is the stability of the vinyl boronic acids. With the advent of the Suzuki coupling, many are commercially available.

organoboronic acid synthesis Organoboronic acid synthesis.png
organoboronic acid synthesis

Other methods of generating boronic acids were also reported. [8] [9]

A wide variety of functional groups including alcohols, carboxylic acids, and amines are tolerated in the Petasis Reaction. Known substrates that are compatible with reaction conditions include vinylboronate esters, arylboronate esters, and potassium organotrifluoroborates. [10] [11] [12] Additionally, a variety of substituted amines can be used other than secondary amines. Tertiary aromatic amines, hydrazines, hydroxylamines, sulfonamides, and indoles have all been reported. [13] [14] [15] [16]

Synthesis of allyl amines

Vinyl boronic acids react with the adducts of secondary amines and paraformaldehyde to give tertiary allylamines. The geometry of the double bond of the starting vinyl boronic acid is retained in the final product: [1]

geometrically pure allylamines E-Allylamines.png
geometrically pure allylamines

This reaction was used to synthesize naftifine [1]

synthesis of naftifine Petasis-naftifine.png
synthesis of naftifine

Synthesis of amino acids

β,γ-unsaturated, N-substituted amino acids are prepared through the condensation of organoboronic acids, boronates, or boronic esters with amines and glyoxylic acids. The highly polar protic solvents Hexafluoroisopropanol (HFIP) can shorten reaction time and improve yield. [17]

PBM coupling to synthesize amino acid with HFIP solvent Generic rxn 1 HFIP.png
PBM coupling to synthesize amino acid with HFIP solvent

Apart from vinyl boronic acids, aryl boronic acids and other heterocyclic derivatives can also be used in Petasis multicomponent coupling. Possible substrate scope includes thienyl, pyridyl, furyl, and benzofuranyl, 1-naphthyl, and aryl groups with either electron-donating or electron-withdrawing substituent. [10]

PBM coupling to synthesize aryl glycine Aryl glycine rxn scheme.png
PBM coupling to synthesize aryl glycine

Racemic Clopidogrel, an antiplatelet agent, was synthesized in two steps using Petasis reaction. [18]

synthesis of clopidogrel via PBM coupling Clopidogrel synthesis.png
synthesis of clopidogrel via PBM coupling

The Petasis reaction exhibits high degrees of stereocontrol when a chiral amine or aldehyde is used as a substrate. When certain chiral amines, such as (S)-2-phenylglycinol, are mixed with an α-keto acid and vinyl boronic acid at room temperature, the corresponding allylamine is formed as a single diastereomer. Furthermore, enantiomeric purity can be achieved by hydrogenation of the diastereoselective product.

stereoselective alpha amino acids Diastereoselectivealphaaminoacids.png
stereoselective alpha amino acids

Unconventional synthesis of carboxylic acids

Apart from amino-acids, PBM reaction can also be used to prepare carboxylic acids, albeit with unconventional mechanisms. In the case of N-substituted indoles as amine equivalent, the reaction begins with the nucleophilic attack of the 3-position of the "N"-substituted indole to electrophilic aldehyde, followed by formation of "ate complex" 1 via the reaction of boronic acid with the carboxylic acid. The intermediate then undergoes dehydration, followed by migration of boronate-alkyl group to furnish the final carboxylic acid product. A wide range of aryl boronic acids is tolerated, while the usage of vinyl boronic acids is not reported. "N"-unsubstituted indoles react very sluggishly under normal reaction conditions, thus confirming the mechanism below. [16]

PBM coupling with N-substituted indole N-sub indole Petasis-acid formation.png
PBM coupling with N-substituted indole


Tertiary aromatic amines can be used in the Petasis reaction as another equivalent of amine nucleophile. The mechanism is similar to the N-substituted indole case. The reaction is carried out under harsh conditions (24-hr reflux in 1,4-dioxane), but the resultant carboxylic acid is obtained in reasonable yield. Usage of α-ketoacids instead of glyoxylic acid does not diminish yields. 1,3,5-trioxygenated benzene derivatives can also be used in lieu of tertiary aromatic amines. [15]

PBM coupling with trisubstituted aromatic amine Petasis with tri-substituted aromatic amine.png
PBM coupling with trisubstituted aromatic amine

Synthesis of iminodicarboxylic acid derivatives


When used as nitrogen nucleophiles, amino acids can furnish various iminodicarboxylic acid derivatives. High diastereoselectivity is usually observed, and the newly formed stereocenter usually share the same configuration with the starting amino acid. This reaction works well in highly polar solvents (ex. water, ethanol, etc.). Peptides with unprotected nitrogen terminal can also be used as a nitrogen nucleophile equivalent. Petasis and coworkers prepared Enalaprilat, an ACE inhibitor, with this method. [19]

synthesis of Enalaprilat via PBM coupling Enalaprilat scheme.png
synthesis of Enalaprilat via PBM coupling

Synthesis of peptidomimetic heterocycles

When diamines are used in PBM reactions, heterocycles of various structures, such as piperazinones, benzopiperazinones, and benzodiazepinones, are efficiently prepared. Lactamization reactions are commonly employed to form the heterocycles, usually under strongly acidic conditions. [19]

preparation of Piperazinones, benzopiperazinones, and benzodiazepinones via PBM coupling Piperazinones, benzopiperazinones, and benzodiazepinones.png
preparation of Piperazinones, benzopiperazinones, and benzodiazepinones via PBM coupling

Synthesis of amino alcohols

When a α-hydroxy aldehyde is used as a substrate in the synthesis of β-amino alcohols, a single diastereomer is generated. This reaction forms exclusively anti-product, confirmed by 1H NMR spectroscopy. The product does not undergo racemization, and when enantiomerically pure α-hydroxy aldehydes are used, enantiomeric excess can be achieved. It is believed that the boronic acid first reacted with the chiral hydroxyl group, furnishing a nucleophilic alkenyl boronate, followed by face selective, intramolecular migration of the alkenyl group into the electrophilic iminium carbon, forming the desired C-C bond irreversibly. [3]

Stereoselective B amino alcohols StereocontrolofBaminoalcohols.png
Stereoselective B amino alcohols

Diastereoselectivity may arise from the reaction of the more stable (and, in this case, more reactive) conformation of the ate complex, where 1,3 allylic strain is minimized. [20] [21] [22]

diastereoselectivity of amino synthesis_mechanism and transition state Stereocontrol mechanism of amino alcohol.png
diastereoselectivity of amino synthesis_mechanism and transition state

With dihydroxyacetone, a somewhat unconventional aldehyde equivalent, Petasis reaction give the core structure of FTY720, a potent immunosuppressive agent. [23]

synthesis of FTY720 FTY720 synthesis.png
synthesis of FTY720

Synthesis of amino polyols and amino sugars

Petasis and coworkers reported the usage of unprotected carbohydrates as the carbonyl component in PBM reactions. It is used as the equivalent of α-hydroxyl aldehydes with pre-existing chirality, and the aminopolyol product is usually furnished with moderate to good yield, with excellent selectivity. A wide variety of carbohydrates, as well as nitrogen nucleophiles (ex. amino acids), can be used to furnish highly stereochemically enriched products. The aminopolyol products can then undergo further reactions to prepare aminosugars. Petasis used this reaction to prepare Boc-protected mannosamine from D-arabinose. [19]

synthesis of Boc-protected mannosammine Mannosamine synthesis.png
synthesis of Boc-protected mannosammine

Applications in enantioselective synthesis

With chiral amine nucleophile

Generally speaking, when chiral amine is used in Petasis coupling, the stereochemical outcome of Petasis reaction is strongly correlated to the chirality of the amine, and high to excellent diastereoselectivity is observed even without the usage of bulky chiral inducing groups. Chiral benzyl amines, [24] 2-substituted pyrrolidines, [25] and 5-substituted 2-morpholinones [26] [27] have been shown to induce good to excellent diastereomeric excess under different Petasis reaction conditions.

diastereoselective PBM coupling with chiral amine Diastereoselective synthesis w- chiral amine.png
diastereoselective PBM coupling with chiral amine

With chiral N-acyliminium ions

Chiral N-acyliminium ion "starting materials" are generally prepared by in-situ dehydration of cyclic hemiaminal. They also carry a chiral hydroxyl group that is in proximity with the iminium carbon; boronic acids react with such chiral hydroxyl groups to form a chiral and electron-rich boronate species, followed by side-selective and intramolecular boronate vinyl/aryl transfer into the iminium carbon. Hence, the reaction is highly diastereoselective, with cis- boronate aryl/vinyl transfer being the predominant pathway. Hydroxypyrrolidines [28] and Hydroxy-γ- and δ-lactams [29] have been shown to react very diastereoselectively, with good to excellent yield. However, such procedures are limited to the usage of vinyl- or electron-rich aryl- boronic acids only.

diastereoselective PBM coupling with chiral N-acyliminium ion Chiral N-acyliminium Petasis reaction.png
diastereoselective PBM coupling with chiral N-acyliminium ion

(±)-6-Deoxycastanospermine was prepared in 7 steps from the vinyl boronic ester. The key acyclic precursor to deoxycastanospermine (A) is formed first by condensing vinyl boronic ester 1 with Cbz-protected hydroxy-pyrrolidine 2 with a PBM coupling, followed by dihydroxylation and TBS protection. A then undergo intramolecular cyclization via a one-pot imine formation and reduction sequel, followed by TBS deprotection, to afford (±)-6-deoxycastanospermine. [30]

Deoxycastanospermine synthesis. Deoxycastanospermine synthesis.png
Deoxycastanospermine synthesis.

With thiourea catalyst

Enantioselective Petasis-type reaction transform quinolines into respective chiral 1,2-dihydroquinolines (product) using alkenyl boronic acids and chiral thiourea catalyst: [31]

Takemoto et al. Takemoto reaction scheme.png
Takemoto et al.

Chloroformates are required as electrophilic activating agents. Also, a 1,2-amino alcohol functionality is required on the catalyst for the reaction to proceed stereoselectively.

Takemoto et al. transition state Takemoto TS1.png
Takemoto et al. transition state

With chiral biphenols

Chiral α-amino acids with various functionalities are conveniently furnished by mixing alkenyl diethyl boronates, secondary amines, glyoxylates, and chiral biphenol catalyst in toluene in one-pot: [32]

Schaus reaction Schaus reaction scheme.png
Schaus reaction

This reaction tolerates a wide range of functionalities, both on the sides of alkenyl boronates and the secondary amine: the electron-richness of the substrates does not affect the yield and enantioselectivity, and sterically demanding substrates (dialkylsubstituted alkenyl boronates and amines with α-stereocenter) only compromise enantioselectivity slightly. Reaction rates do vary on a case-by-case basis. [32]

Under the reported condition, boronic acids substrates failed to give any enantioselectivity. Also, 3Å molecular sieve is used in the reaction system. While the authors did not provide the reason for such usage in the paper, it was speculated that 3Å molecular sieves act as water scavenger and prevent the decomposition of alkenyl diethyl boronates into their respective boronic acids. The catalyst could be recycled from the reaction and reused without compromising yield or enantioselectivity. [32]

More recently, Yuan with coworkers from Chengdu Institute of Organic Chemistry, Chinese Academy of Science combined both approaches (chiral thiourea catalyst and chiral biphenol) in a single catalyst, reporting for the first time the catalytic system that is capable of performing enantioselective Petasis reaction between salicylaldehydes, cyclic secondary amines and aryl- or alkenylboronic acids: [33]

Yuan reaction Yuan reaction scheme.png
Yuan reaction

In one application the Petasis reaction is used for quick access to a multifunctional scaffold for divergent synthesis. The reactants are the lactol derived from L-phenyl-lactic acid and acetone, l-phenylalanine methyl ester and a boronic acid. The reaction takes place in ethanol at room temperature to give the product, an anti-1,2-amino alcohol with a high diastereomeric excess. [34]

Petasis reaction example (Kumagai et al.) PetasisReaction.png
Petasis reaction example (Kumagai et al.)

Notice that the authors cannot assess syn-1,2-amino alcohol with this method due to intrinsic mechanistic selectivity, and the authors argue that such intrinsic selectivity hampers their ability to access the full matrix of stereoisomeric products for the usage of small molecule screening. In a recent report, Schaus and co-workers reported that syn amino alcohol can be obtained with the following reaction condition, using a chiral dibromo-biphenol catalyst their group developed: [35]

schaus_ACIE_reaction Schaus ACIE reaction scheme.png
schaus_ACIE_reaction

Although the syn vs. anti diastereomeric ratio ranges from mediocre to good (1.5:1 to 7.5:1), the substrate scope for such reactions remain rather limited, and the diastereoselectivity is found to be dependent on the stereogenic center on the amine starting material. [35]

Petasis reaction and total synthesis

Beau and coworkers assembled the core dihydropyran framework of zanamivir congeners via a combination of PBM reaction and Iron(III)-promoted deprotection-cyclization sequence. A stereochemically defined α-hydroxyaldehyde 2, diallylamine and a dimethylketal-protected boronic acid 1 is coupled to form the acyclic, stereochemically defined amino-alcohol 3, which then undergoes an Iron(III)-promoted cyclization to form a bicyclic dihydropyran 4. Selective opening of the oxazoline portion of the dihydropyran intermediate 4 with water or timethylsilyl azide then furnish downstream products that have structures resembling the Zanamivir family members. [36]

zanamivir core_Beau et al. Beau et al. schemes3.png
zanamivir core_Beau et al.

Wong and coworkers prepared N-acetylneuraminic acid with a PBM coupling, followed by nitrone-[3+2] cycloaddition. Vinylboronic acid is first coupled with L-arabinose 1 and Bis(4-methoxyphenyl)methanamine 2 to form an stereochemically defined allyl amine 3. Afterwards, the sequence of dipolar cycloaddition, base-mediated N–O bond breakage and hydrolysis then complete the synthesis of N-acetylneuraminic acid. [37]

N-acetylneuraminic acid_Wong et al. Wong et al. scheme.png
N-acetylneuraminic acid_Wong et al.

See also

Related Research Articles

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

The aldol reaction is a means of forming carbon–carbon bonds in organic chemistry. Discovered independently by the Russian chemist Alexander Borodin in 1869 and by the French chemist Charles-Adolphe Wurtz in 1872, the reaction combines two carbonyl compounds to form a new β-hydroxy carbonyl compound. These products are known as aldols, from the aldehyde + alcohol, a structural motif seen in many of the products. Aldol structural units are found in many important molecules, whether naturally occurring or synthetic. For example, the aldol reaction has been used in the large-scale production of the commodity chemical pentaerythritol and the synthesis of the heart disease drug Lipitor.

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.

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

The Henry reaction is a classic carbon–carbon bond formation reaction in organic chemistry. Discovered in 1895 by the Belgian chemist Louis Henry (1834–1913), it is the combination of a nitroalkane and an aldehyde or ketone in the presence of a base to form β-nitro alcohols. This type of reaction is also referred to as a nitroaldol reaction. It is nearly analogous to the aldol reaction that had been discovered 23 years prior that couples two carbonyl compounds to form β-hydroxy carbonyl compounds known as "aldols". The Henry reaction is a useful technique in the area of organic chemistry due to the synthetic utility of its corresponding products, as they can be easily converted to other useful synthetic intermediates. These conversions include subsequent dehydration to yield nitroalkenes, oxidation of the secondary alcohol to yield α-nitro ketones, or reduction of the nitro group to yield β-amino alcohols.

<span class="mw-page-title-main">Chiral auxiliary</span> Stereogenic group placed on a molecule to encourage stereoselectivity in reactions

In stereochemistry, a chiral auxiliary is a stereogenic group or unit that is temporarily incorporated into an organic compound in order to control the stereochemical outcome of the synthesis. The chirality present in the auxiliary can bias the stereoselectivity of one or more subsequent reactions. The auxiliary can then be typically recovered for future use.

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

The Betti reaction is a chemical addition reaction of aldehydes, primary aromatic amines and phenols producing α-aminobenzylphenols.

<span class="mw-page-title-main">Weinreb ketone synthesis</span> Chemical reaction

The Weinreb–Nahm ketone synthesis is a chemical reaction used in organic chemistry to make carbon–carbon bonds. It was discovered in 1981 by Steven M. Weinreb and Steven Nahm as a method to synthesize ketones. The original reaction involved two subsequent nucleophilic acyl substitutions: the conversion of an acid chloride with N,O-Dimethylhydroxylamine, to form a Weinreb–Nahm amide, and subsequent treatment of this species with an organometallic reagent such as a Grignard reagent or organolithium reagent. Nahm and Weinreb also reported the synthesis of aldehydes by reduction of the amide with an excess of lithium aluminum hydride.

<span class="mw-page-title-main">Boronic acid</span> Organic compound of the form R–B(OH)2

A boronic acid is an organic compound related to boric acid in which one of the three hydroxyl groups is replaced by an alkyl or aryl group. As a compound containing a carbon–boron bond, members of this class thus belong to the larger class of organoboranes.

Chiral Lewis acids (CLAs) are a type of Lewis acid catalyst. These acids affect the chirality of the substrate as they react with it. In such reactions, synthesis favors the formation of a specific enantiomer or diastereomer. The method is an enantioselective asymmetric synthesis reaction. Since they affect chirality, they produce optically active products from optically inactive or mixed starting materials. This type of preferential formation of one enantiomer or diastereomer over the other is formally known as asymmetric induction. In this kind of Lewis acid, the electron-accepting atom is typically a metal, such as indium, zinc, lithium, aluminium, titanium, or boron. The chiral-altering ligands employed for synthesizing these acids often have multiple Lewis basic sites that allow the formation of a ring structure involving the metal atom.

Electrophilic amination is a chemical process involving the formation of a carbon–nitrogen bond through the reaction of a nucleophilic carbanion with an electrophilic source of nitrogen.

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

Benzylic activation and stereocontrol in tricarbonyl(arene)chromium complexes refers to the enhanced rates and stereoselectivities of reactions at the benzylic position of aromatic rings complexed to chromium(0) relative to uncomplexed arenes. Complexation of an aromatic ring to chromium stabilizes both anions and cations at the benzylic position and provides a steric blocking element for diastereoselective functionalization of the benzylic position. A large number of stereoselective methods for benzylic and homobenzylic functionalization have been developed based on this property.

The Baylis–Hillman reaction is a carbon-carbon bond forming reaction between the α-position of an activated alkene and a carbon electrophile such as an aldehyde. Employing a nucleophilic catalyst, such as a tertiary amine and phosphine, this reaction provides a densely functionalized product. It is named for Anthony B. Baylis and Melville E. D. Hillman, two of the chemists who developed this reaction while working at Celanese. This reaction is also known as the Morita–Baylis–Hillman reaction or MBH reaction, as K. Morita had published earlier work on it.

<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.

Metal-catalyzed C–H borylation reactions are transition metal catalyzed organic reactions that produce an organoboron compound through functionalization of aliphatic and aromatic C–H bonds and are therefore useful reactions for carbon–hydrogen bond activation. Metal-catalyzed C–H borylation reactions utilize transition metals to directly convert a C–H bond into a C–B bond. This route can be advantageous compared to traditional borylation reactions by making use of cheap and abundant hydrocarbon starting material, limiting prefunctionalized organic compounds, reducing toxic byproducts, and streamlining the synthesis of biologically important molecules. Boronic acids, and boronic esters are common boryl groups incorporated into organic molecules through borylation reactions. Boronic acids are trivalent boron-containing organic compounds that possess one alkyl substituent and two hydroxyl groups. Similarly, boronic esters possess one alkyl substituent and two ester groups. Boronic acids and esters are classified depending on the type of carbon group (R) directly bonded to boron, for example alkyl-, alkenyl-, alkynyl-, and aryl-boronic esters. The most common type of starting materials that incorporate boronic esters into organic compounds for transition metal catalyzed borylation reactions have the general formula (RO)2B-B(OR)2. For example, bis(pinacolato)diboron (B2Pin2), and bis(catecholato)diborane (B2Cat2) are common boron sources of this general formula.

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">Protodeboronation</span>

Protodeboronation, or protodeborylation is a chemical reaction involving the protonolysis of a boronic acid in which a carbon-boron bond is broken and replaced with a carbon-hydrogen bond. Protodeboronation is a well-known undesired side reaction, and frequently associated with metal-catalysed coupling reactions that utilise boronic acids. For a given boronic acid, the propensity to undergo protodeboronation is highly variable and dependent on various factors, such as the reaction conditions employed and the organic substituent of the boronic acid.

In organophosphorus chemistry, the Pudovik reaction is a method for preparing α-aminomethylphosphonates. Under basic conditions, the phosphorus–hydrogen bond of a dialkylphosphite, (RO)2P(O)H, adds across the carbon–nitrogen double bond of an imine (a hydrophosphonylation reaction). The reaction is closely related to the three-component Kabachnik–Fields reaction, where an amine, phosphite, and an organic carbonyl compound are condensed, which was reported independently by Martin Kabachnik and Ellis Fields in 1952. In the Pudovik reaction, a generic imine, RCH=NR', would react with a phosphorous reagent like diethylphosphite as follows:

Miyaura borylation, also known as the Miyaura borylation reaction, is a named reaction in organic chemistry that allows for the generation of boronates from vinyl or aryl halides with the cross-coupling of bis(pinacolato)diboron in basic conditions with a catalyst such as PdCl2(dppf). The resulting borylated products can be used as coupling partners for the Suzuki reaction.

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

The nitro-Mannich reaction is the nucleophilic addition of a nitroalkane to an imine, resulting in the formation of a beta-nitroamine. With the reaction involving the addition of an acidic carbon nucleophile to a carbon-heteroatom double bond, the nitro-Mannich reaction is related to some of the most fundamental carbon-carbon bond forming reactions in organic chemistry, including the aldol reaction, Henry reaction and Mannich reaction.

The ketimine Mannich reaction is an asymmetric synthetic technique using differences in starting material to push a Mannich reaction to create an enantiomeric product with steric and electronic effects, through the creation of a ketimine group. Typically, this is done with a reaction with proline or another nitrogen-containing heterocycle, which control chirality with that of the catalyst. This has been theorized to be caused by the restriction of undesired (E)-isomer by preventing the ketone from accessing non-reactive tautomers. Generally, a Mannich reaction is the combination of an amine, a ketone with a β-acidic proton and aldehyde to create a condensed product in a β-addition to the ketone. This occurs through an attack on the ketone with a suitable catalytic-amine unto its electron-starved carbon, from which an imine is created. This then undergoes electrophilic addition with a compound containing an acidic proton. It is theoretically possible for either of the carbonyl-containing molecules to create diastereomers, but with the addition of catalysts which restrict addition as of the enamine creation, it is possible to extract a single product with limited purification steps and in some cases as reported by List et al.; practical one-pot syntheses are possible. The process of selecting a carbonyl-group gives the reaction a direct versus indirect distinction, wherein the latter case represents pre-formed products restricting the reaction's pathway and the other does not. Ketimines selects a reaction group, and circumvent a requirement for indirect pathways.

References

  1. 1 2 3 4 Petasis, N. A.; Akritopoulou, I. (1993). "The boronic acid mannich reaction: A new method for the synthesis of geometrically pure allylamines". Tetrahedron Lett. 34 (4): 583–586. doi:10.1016/S0040-4039(00)61625-8.
  2. Petasis, N. A.; Zavialov, I. A. (1997). "A New and Practical Synthesis of -Amino Acids from Alkenyl Boronic Acids". J. Am. Chem. Soc. 119 (2): 445–446. doi:10.1021/ja963178n.
  3. 1 2 Petasis, N. A.; Zavialov, I. A. (1998). "Highly Stereocontrolled One-Step Synthesis of anti-β-Amino Alcohols from Organoboronic Acids, Amines, and α-Hydroxy Aldehydes". J. Am. Chem. Soc. 120 (45): 11798–11799. doi:10.1021/ja981075u.
  4. Candeias, N. R.; Montalbano, F.; Cal, P.M.S.D., Gois, P.M.P. (2010). "Boronic Acids and Esters in the Petasis-Borono Mannich Multicomponent Reaction". Chem. Rev. 110 (10): 6169–6193. doi:10.1021/cr100108k. PMID   20677749.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  5. Batey, R.A. (2005). "Nucleophilic Addition Reactions of Aryl and Alkenylboronic Acids and Their Derivatives to Imines and Iminium Ions". In Hall, D.G. (ed.). Boronic Acids: Preparation and Applications in Organic Synthesis and Medicine . Wiley-VCH. pp.  279–304. doi:10.1002/3527606548.ch7. ISBN   9783527606542.
  6. Ramadhar, T.R.; Batey, R.A. (2011). "Recent Advances in Nucleophilic Addition Reactions of Organoboronic Acids and Their Derivatives to Unsaturated CN Functionalities". In Hall, D.G. (ed.). Boronic Acids: Preparation and Applications in Organic Synthesis, Medicine and Materials, Second Edition. Wiley-VCH. pp. 427–477. doi:10.1002/9783527639328.ch9. ISBN   9783527639328.
  7. Wu, Peng; Givskov, Michael; Nielsen, Thomas E. (27 August 2019). "Reactivity and Synthetic Applications of Multicomponent Petasis Reactions". Chemical Reviews. 119 (20): 11245–11290. doi: 10.1021/acs.chemrev.9b00214 . PMC   6813545 . PMID   31454230.
  8. Hoffmann. R.W.; Dresely, S. (1988), "Preparation of 3-substituted (E)-1-alkenylboronic esters", Synthesis , 1988 (2): 103–106, doi:10.1055/s-1988-27480
  9. Brown, H.C.; Bhat, N.G.; Iyer, R.R. (1991), "A novel route to 1,3-dienyl-2-boronic esters providing simple syntheses of conjugated (E,E)-dienes and conjugated (E)-alkenones", Tetrahedron Lett. , 32 (30): 3655–3658, doi:10.1016/s0040-4039(00)79758-9
  10. 1 2 Petasis, N.A.; Goodman, A., Zavialov, I.A. (1997), "A new synthesis of α-arylglycines from aryl boronic acids", Tetrahedron , 53 (48): 16463–16470, doi:10.1016/S0040-4020(97)01028-4 {{citation}}: CS1 maint: multiple names: authors list (link)
  11. Kabalka, G.W.; Venkataiah, B.; Dong, G. (2004), "The use of potassium alkynyltrifluoroborates in Mannich reactions", Tetrahedron Lett., 45 (4): 729–731, doi:10.1016/j.tetlet.2003.11.049
  12. Tremblay-Morin, J.-P.; Raeppel, S.; Gaudette, F. (2004), "Lewis acid-catalyzed Mannich type reactions with potassium organotrifluoroborate", Tetrahedron Lett., 45 (17): 3471–3474, doi:10.1016/j.tetlet.2004.03.014
  13. Portlock, D.E.; Naskar, D.; West, L.; Li, M. (2002), "Petasis boronic acid-Mannich Reactions of substituted hydrazines: synthesis of α-hydrazino carboxylic acids", Tetrahedron Lett., 43 (38): 6845–6847, doi:10.1016/S0040-4039(02)01511-3
  14. Naskar, D., Roy, A., Seibel, W.L., Portlock, D.E. (2003), "Hydroxylamines and sulfinamide as amine components in the Petasis boronic acid–Mannich reaction: synthesis of N-hydroxy or alkoxy-α-aminocarboxylicacids and N-(tert-butyl sulfinyl)-α-amino carboxylicacids", Tetrahedron Lett., 44 (49): 8865–8868, doi:10.1016/j.tetlet.2003.09.179 {{citation}}: CS1 maint: multiple names: authors list (link)
  15. 1 2 Naskar, D.; Roy, A.; Seibel, W.L., Portlock, D.E. (2003), "Novel Petasis Boronic Acid—Mannich Reactions with Tertiary Aromatic Amines", Tetrahedron Lett. , 44 (31): 5819–5821, doi:10.1016/S0040-4039(03)01405-9 {{citation}}: CS1 maint: multiple names: authors list (link)
  16. 1 2 Naskar, D.; Neogi, S.; Roy, A.; Mandal, A.B. (2008), "Novel Petasis boronic acid reactions with indoles: synthesis of indol-3-yl-aryl-acetic acids", Tetrahedron Lett., 49 (48): 6762–6764, doi:10.1016/j.tetlet.2008.08.029
  17. Jourdan, H.; Gouhier, G.; Van Hijfte, L.; Angibaud, P.; Piettre, S. R. (2005). "On the use of boronates in the Petasis reaction". Tetrahedron Lett. 46 (46): 8027–8031. doi:10.1016/j.tetlet.2005.09.060.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  18. Kalinski, C.; Lemoine, H.; Schmidt, J.; Burdack, C.; Kolb, J.; Umkehrer, M.; Ross, G. (2008). "Multicomponent Reactions as a Powerful Tool for Generic Drug Synthesis". Synthesis. 2008 (24): 4007–4011. doi:10.1055/s-0028-1083239.
  19. 1 2 3 Petasis, N.A. (2005). "Multicomponent Reactions with Organoboron Compounds". In Zhu, J.; Bienayme, H. (eds.). Multicomponent Reactions . Wiley-VCH. pp.  199–223. doi:10.1002/3527605118.ch7. ISBN   9783527605118.
  20. Davis, A.S.; Pyne, S. G.; Skelton, B. W.; White, A. H. (2004). "Synthesis of putative uniflorine A". J. Org. Chem. 69 (9): 3139–43. doi:10.1021/jo049806y. PMID   15104453.
  21. Au, C. W. G.; Pyne, S.G. (2006). "Asymmetric synthesis of anti-1,2-amino alcohols via the Borono-Mannich reaction: A formal synthesis of (−)-swainsonine". J. Org. Chem. 71 (18): 7097–9. doi:10.1021/jo0610661. PMID   16930074.
  22. Pyne, S.G.; Au, C. W. G.; Davis, A. S.; Morgan, I. R.; Ritthiwigrom, T.; Yazici, A. (2008). "Exploiting the borono-Mannich reaction in bioactive alkaloid synthesis". Pure Appl. Chem. 80 (4): 751–762. doi: 10.1351/pac200880040751 .
  23. Sugiyama, S.; Arai, S.; Kiriyama, M.; Ishii, K. (2005). "A convenient synthesis of immunosuppressive agent FTY720 using the petasis reaction". Chem. Pharm. Bull. 53 (1): 100–2. doi: 10.1248/cpb.53.100 . PMID   15635240.
  24. Jiang, B.; Yang, C.-G.; Gu, X.-H. (2001). "A highly stereoselective synthesis of indolyl N-substituted glycines". Tetrahedron Lett. 42 (13): 2545–2547. doi:10.1016/s0040-4039(01)00229-5.
  25. Nanda, K.K.; Trotter, B.W. (2005). "Diastereoselective Petasis Mannich reactions accelerated by hexafluoroisopropanol: a pyrrolidine-derived arylglycine synthesis". Tetrahedron Lett. 46 (12): 2025–8. doi:10.1016/j.tetlet.2005.01.151.
  26. Harwood, L.M.; Currie, G. S.; Drew, M. G. B.; Luke, R. W. A. (1996). "Asymmetry in the boronic acid Mannich reaction: diastereocontrolled addition to chiral iminium species derived from aldehydes and (S)-5-phenylmorpholin-2-one". Chem. Commun. (16): 1953. doi:10.1039/cc9960001953.
  27. Currie, G.S.; Drew, M. G. B.; Harwood, L. M.; Hughes, D. J.; Luke, R. W. A.; Vickers, R. J. (2000). "Chirally templated boronic acid Mannich reaction in the synthesis of optically active α-amino acids". J. Chem. Soc., Perkin Trans. 1 (17): 2982–2990. doi:10.1039/B003067H.
  28. Batey, R.A.; MacKay, D. B.; Santhakumar, V. (1999). "Alkenyl and Aryl BoronatesMild Nucleophiles for the Stereoselective Formation of Functionalized N -Heterocycles". J. Am. Chem. Soc. 121 (21): 5075–5076. doi:10.1021/ja983801z.
  29. Morgan, I.R.; Yazici, A.; Pyne, S. G. (2008). "Diastereoselective borono-Mannich reactions on cyclic N-acyliminium ions". Tetrahedron. 64 (7): 1409–1419. doi:10.1016/j.tet.2007.11.046.
  30. 1 2 Batey, R.A.; MacKay, D.B. (2000). "Total synthesis of (±)-6-deoxycastanospermine: an application of the addition of organoboronates to N-acyliminium ions". Tetrahedron Lett. 41 (51): 9935–9938. doi:10.1016/s0040-4039(00)01790-1.
  31. Yamaoka, Y.; Miyabe, H.; Takemoto, Y. (2007). "Catalytic enantioselective petasis-type reaction of quinolines catalyzed by a newly designed thiourea catalyst". J. Am. Chem. Soc. 129 (21): 6686–7. doi:10.1021/ja071470x. PMID   17488015.
  32. 1 2 3 Lou, S.; Schaus, S.E. (2008). "Asymmetric petasis reactions catalyzed by chiral biphenols". J. Am. Chem. Soc. 130 (22): 6922–6923. doi:10.1021/ja8018934. PMC   2440570 . PMID   18459782.
  33. Han, W.-Y.; Wu, Z.-J.; Zhang, X.-M.; Yuan, W.-C. (2012), "Enantioselective Organocatalytic Three-Component Petasis Reaction among Salicylaldehydes, Amines, and Organoboronic Acids", Org. Lett. , ASAP (4): 976–979, doi:10.1021/ol203109a, PMID   22292670, S2CID   9416125
  34. Naoya Kumagai, Giovanni Muncipinto, Stuart L. Schreiber; Muncipinto; Schreiber (2006). "Short Synthesis of Skeletally and Stereochemically Diverse Small Molecules by Coupling Petasis Condensation Reactions to Cyclization Reactions". Angewandte Chemie International Edition . 45 (22): 3635–3638. doi:10.1002/anie.200600497. PMID   16646101.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  35. 1 2 Muncipinto, G.; Moquist, P.N.; Schreiber, S.L.; Scahus, S.E. (2011). "Catalytic Diastereoselective Petasis Reactions". Angew. Chem. Int. Ed. 50 (35): 8172–8175. doi:10.1002/anie.201103271. PMC   4673970 . PMID   21751322.
  36. Soule, J.-F.; Mathieu, A.; Norsikian, S.; Beau, J.-M. (2010). "Coupling the Petasis condensation to an iron(III) chloride-promoted cascade provides a short synthesis of Relenza congeners". Org. Lett. 12 (22): 5322–5325. doi:10.1021/ol102326b. PMID   20945892.
  37. Hong, Z.; Liu, L.; Hsu, C.-C.; Wong, C,-H. (2006). "Three-Step Synthesis of Sialic Acids and Derivatives" (PDF). Angew. Chem. Int. Ed. 45 (44): 7417–7421. doi:10.1002/anie.200601555. PMID   17031889.{{cite journal}}: CS1 maint: multiple names: authors list (link)