Ketimine Mannich reaction

Last updated

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 [1] 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 (which is an enol). 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.; [2] 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.

Contents

Theory

It is well-known that Lewis acids and bases can influence carbonyl activity by either protonation of the oxygen or de-protonation of a beta-site to influence electrophilicity at the carbon. In the former case, there is a stabilization effect and priming for a potential leaving group by providing an equilibrium between a formal positive charge and an alcohol group. In either case, the carbon sees enhanced reactivity as the bonds are strained to equalize charge by seeking a nucleophile to which the building charge will equalize into in the subsequent reaction. Potentiation in such reactions is therefore generally driven by how well a material can produce and stabilize a charge gradient through resonance or inductive effects in substituents. The surfaces of acids can catalyze these effects by stabilizing the distortion of the carbonyl's electron cloud, leading to sensitivity in the reaction to acid-base conditions, with acidic conditions typically favoring reaction by enhancing carbonyl leavability.

Mannich reactions are named after their pioneer, Carl Mannich. After the discovery of Mannich reaction in 1912, he found that combination of a ketone, aldehyde and amine consistently produced an addition of the aldehyde unto the site originally containing an acidic β-proton in the ketone. He theorized the mechanism to be one of mixed-aldol; featuring the dehydration of an alcohol and Michael-Addition of the complex. The reaction suffered from a lack of specificity, and this problem persisted until 1997, in which an asymmetrical method was discovered independently by several researchers. Kobayashi [3] et al. used Brønsted-acid to demonstrate that the organic reaction could occur in an aqueous medium, and also found an enantiomeric excess of nearly 55%. It was hypothesized that one of the reactants must be using the acid surface to change its electronic structure since otherwise the mixture would be immiscible, though the study used chiral Lewis acid as catalyst. Prior, Hajos [4] et al. demonstrated a method by which L-proline could be used in an aldol-cyclization.

Subsequent studies have focused on improving the catalyst or materials through substituent effects, from which controls using sterics or restriction of charge sites are expected to improve catalytic yields. Alternatives to the catalyst are not readily explored due to the ubiquity of L-proline, its low cost, and its high selectivity; cementing its proliferation as a catalytic standard. L-proline restricted reactions through its five-membered ring which favours the addition of another reactant in a fixed direction. List [2] et al. theorize in their study that this restriction is primarily in the carboxylic-acid group next to the enamine attachment, which can stabilize the imine product. List also expands upon the role of the imine by listing its ratio with the aldehyde to be 1 using 1H-NMR methods, indicating that the Michael Addition is rate-determining and not the proline-complexation or, obviously, aldehyde activation. This is further confirmed in a Hammett study on para-substituted aromatic aldehydes, in which List confirmed positive correlation between withdrawing effect and increasing imine reactivity, a sign that Michael addition had to be happening with the former aldehyde as an electrophile. [2]

Substituent studies

Substituent studies have focused on increasing or decreasing steric or electronic influences on material to favor product. In 2012, [5] the first non-aromatic ketiminoester was produced with the capability of reduction into either syn- or anti- lactones with NaBH4 reduction. Kano et al. demonstrated a 20:1 excess of desired product, electrophilic enhancement of the ketone by different flanking esters.

Endocyclic ketimines without electron-withdrawing substituents

BINOL phosphoric acid-catalyzed Mannich reaction (2019)

Reddy and coworkers proposed the method to produce endocyclic N-acyl ketimines from a stable precursor, 3-aryl-3-hydroxyisoindolin-1-ones in 2019. The reaction yields a high stereoselectivity under high temperature as the adjacent quaternary and stereogenic centers creating chirality. [6] Meanwhile, if 3-hydroxy-3-pentylisoindoline-1-one is used, 95% enamide would be generated instead of Mannich product under the same mechanism.

Scheme 11. BINOL phosphoric acid-catalyzed Mannich reaction of endocyclic N-acyl ketimines. Scheme 11 G.png
Scheme 11. BINOL phosphoric acid-catalyzed Mannich reaction of endocyclic N-acyl ketimines.

Proline-catalyzed direct asymmetric Mannich reaction of 3-substituted-2H-1,4-benzoxazines (2013)

In 2013, Wang and coworkers studied using 3,4-dihydro-2H-1,4-benzoxazines to produce N-heterocyclic products. [7] This reaction was the first catalytic asymmetric Mannich reaction of 3,4-dihydro-2H-1,4-benzoxazines. The aromatic ring strain in 3,4-dihydro-2H-1,4-benzoxazines helps increase reactivity of the C=N double bond and gives the considerable yield with no electron withdrawing substituents. Similar mechanism also occurred in the reaction catalyzed by wheat germ lipase in 2016 provided by Guan, He and coworkers. [8]

Scheme 12. Proline-catalyzed Mannich reaction of 3-substituted-2H-1,4-benzoxazines. Scheme 12 G.png
Scheme 12. Proline-catalyzed Mannich reaction of 3-substituted-2H-1,4-benzoxazines.

Copper-catalyzed asymmetric Mannich reaction of 2H-azirines with β-keto amides (2018)

In 2018, Lin, Feng, and coworkers developed the first copper-catalyzed catalytic enantioselective Mannich reaction using 2H-azirines and β-ketoamides. [9] The racemic 2H-azirines was first applied and one enantiomer of the 2H-azirine would react with chiral Cu-enolated complex to help overcome its three neighbouring stereogenic centers and gave proper products.

Scheme 13. Copper-catalyzed asymmetric Mannich reaction of 2H-azirines with b-ketoamides. Scheme 13 G.png
Scheme 13. Copper-catalyzed asymmetric Mannich reaction of 2H-azirines with β-ketoamides.

Copper(I)-catalyzed asymmetric decarboxylative Mannich reaction of 2H-azirines (2019)

In 2019, Yin and coworkers used 2H-azirines as an electrophile and a base at the same time to deprotonate cyanoacetic acid and thus a further decarboxylation could proceed. Proton transfer strategy was also discussed since the protonated 2H-azirine obtains high electrophilicity while nonprotonated ones does not even react with exact same environment. Various substituents on both nucleophile and electrophile gave high enantiomer-enriched yields of azirines. [10]

Scheme 14. Copper(I)-catalyzed asymmetric decarboxylative Mannich reaction of 2H-azirines. Scheme 14 G.png
Scheme 14. Copper(I)-catalyzed asymmetric decarboxylative Mannich reaction of 2H-azirines.

Zn-ProPhenol catalyzed asymmetric Mannich reaction of 2H-azirines with alkynyl cycloalkyl ketones (2020)

In 2020, Trost and coworker proposed a Mannich reaction of 2H-aziridines with alkylnyl cycloalkyl ketones under a bimetallic Zn-ProPhenol complex as the catalyst. The bimetallic Zn-ProPhenol complex employed in the reaction to activate both nucleophile and electrophile since the compound obtains a Bronsted basic site and a Lewis acidic group at chiral center at the same time. The hydrogen bonding between N from aziridine and H from carbonyl group might also contribute chirality centred at nitrogen atom. In this case, N-H bond acetylation product was also generated from the reaction. [11]

Scheme 15. Zn-ProPhenol catalyzed asymmetric Mannich reaction of 2H-azirines with alkynyl cycloalkyl ketones. Scheme 15.png
Scheme 15. Zn-ProPhenol catalyzed asymmetric Mannich reaction of 2H-azirines with alkynyl cycloalkyl ketones.

Chiral phosphoric acid-catalyzed Mukaiyama–Mannich reaction of endocyclic N-acyl ketimines (2020)

In 2020, Zhang, Ma, and coworkers used 3-hydroxyisoindolin-1-ones to produce endocyclic N-acyl ketimines by chiral phosphoric acid-catalyzed Mukaiyama–Mannich reaction. Difluorinated silyl enol ethers are involved to produce enantioenriched fluoroalkyl-functionalized isoindolones first. Through the reaction under different conditions, two features were discovered. Firstly, hexafluoroisopropyl alcohol could enhance the reactivity and enantioselectivity. Secondly, reaction is better catalyzed by catalyst with trifluoromethyl group chiral barriers. [12]

Scheme 16. Chiral phosphoric acid-catalyzed Mukaiyama-Mannich reaction of endocyclic N-acyl ketimines. Scheme 16.png
Scheme 16. Chiral phosphoric acid-catalyzed Mukaiyama–Mannich reaction of endocyclic N-acyl ketimines.

Isatin-derived ketimines

Isatin-derived ketimine commonly undergoes catalytic asymmetric Mannich reaction to generate 3-substituted-3-aminooxindoles. The structure could be found in various natural or biological molecules, and such reaction could lead to enantio-enriched products. Moreover, the isatin-derived ketimines are comparatively reactive as an electrophile and thus was commonly used in various methods or reactions.

Morimoto, Ohshima, and coworkers applied decarboxylative Mannich reaction on N-unprotected isatin-derived ketimines in 2018. [13] The researchers applied an enantio-enriched chiral oxindoles on the 3-positions with primary amines. In this case, substrates with less substituted groups are more likely to undergo Mannich reactions with higher yield. Also, this routine allows a great range of ketoacids to react with ketimines before decarboxylation and generate Mannich products.

Scheme 17. Copper-catalyzed Mannich reaction of N-unprotected isatin-derived ketimines. Scheme 17.png
Scheme 17. Copper-catalyzed Mannich reaction of N-unprotected isatin-derived ketimines.

Wolf and coworkers further developed a copper-catalyzed stereo-divergent Mannich reaction in 2020. [14] Isatin-derived ketimines first react with α-fluoro-α-arylnitrile and result in a chiral cuprous keteniminate complex. Then, BTMG would act as a nucleophile and the stereoselectivity would be determined by the chosen of chiral ligands. From experiment, Segphos–copper complex generated anti diastereomers while Taniaphos complex results in syn diastereomers.

Scheme 18. Stereodivergent asymmetric Mannich reaction of a-fluoronitriles with isatin-derived ketimines. Scheme 18.png
Scheme 18. Stereodivergent asymmetric Mannich reaction of α-fluoronitriles with isatin-derived ketimines.

Feng and coworkers used chiral N,N' -dioxide/Zn(II) complexes to catalyze silyl enol ethers and ketimines in 2019. [15] The reaction started with an α-alkenyl addition of silyl enol ethers. Dioxide (compound 21 and 22 in scheme 20) was proved to the best reaction condition for a wide range of isatin-derived ketimines to generate great yield of products with high stereoselectivity. Pyrazolinone-derived ketimines are also found to be suitable substates. The mechanism demonstrates that Mukaiyama–Mannich addition intermediates produced β-amino silyl enol ethers as product.

Scheme 20. Tandem a-alkenyl addition/proton shift reaction of silyl enol ethers with ketimines. Scheme 20.png
Scheme 20. Tandem α-alkenyl addition/proton shift reaction of silyl enol ethers with ketimines.

Acyclic ketimines with electron-withdrawing groups and/or alkynes

Esters, perfluorinated alkyl, or alkyne groups are good electrophiles, that can be used to form ketimine enolate with different nucleophiles. However, subtle differences in ketimine structures can significantly affect stereoselectivities and yields. Nevertheless, many medicinally and synthetically significant chiral structure can be created by creative design on modified ketimines.

In 2016, a novel direct asymmetric Mannich reaction of ketiminoesters with thionolactones were proposed by Terada et al. by using bis(guanidino)iminophosphorane as a chiral organosuperbase catalyst. (scheme 21) [16] Comparing to thionolactones as an appropriate nucleophile, lactone is not acidic enough to be enolized by base catalyst, which makes lactone unable to act as a nucleophile. For electrophiles, having proper substituents on benzoyl protecting group can improve yields and diastereoselectivity. For instance, comparing to trifluromethyl group on ketiminoester, methoxy substituents gives much lower yield and selectivity.

Scheme 21. Chiral organosuperbase-catalyzed direct Mannich reaction of ketiminoesters. Scheme 21 Z.png
Scheme 21. Chiral organosuperbase-catalyzed direct Mannich reaction of ketiminoesters.

In 2016, direct asymmetric catalytic Mannich reaction of an alpha, beta-unsaturated gamma-butyrolactam with ketiminoesters was proposed by Kumagai el al. using soft Lewis acid/hard Brønsted base cooperative catalysis. (scheme 22) [17] As a result, alpha-addition products were synthesized from alpha,beta-unsaturated gamma-butyrolactams onto ketimines through asymmetric catalysis, which is a novel catalytic asymmetric reaction. The corresponding phosphinoyl protected ketiminoester did not react without the cooperative catalysis by the soft Lewis acid / hard Bronsted base, which means the interaction between copper catalyst (as soft Lewis acid) and thiophosphinoyl protecting group of ketimines (as soft Lewis base) is significant for this successful Mannich reaction.

Scheme 22. Direct enantioselective Mannich reaction of a,b-unsaturated g-butyrolactam with ketiminoesters. Scheme 22.png
Scheme 22. Direct enantioselective Mannich reaction of α,β-unsaturated γ-butyrolactam with ketiminoesters.

In 2017, a direct enantioselective Mannich reaction with an N-unprotected trifluoromethyl ketiminoester was proposed by Morimoto et al. (scheme 23) [18] The unprotected N-H ketimines alleviate problems caused by E/Z isomerism in the asymmetric catalytic reactions, although they generally have lower stability. Morimoto et al. showed that the unprotected N-H ketimines are able to react with malonates, esters, oxindoles, and cyclic beta-keto-nitiles under different temperatures to give products whose diastereoselectivities can be controlled under asymmetric catalysis.

Scheme 23. Direct catalytic Mannich reaction with an N-unprotected trifluoromethyl ketiminoester. Scheme 23.png
Scheme 23. Direct catalytic Mannich reaction with an N-unprotected trifluoromethyl ketiminoester.

In 2018, Zn-ProPhenol was used to catalyze asymmetric Mannich reaction of butenolides with perfluoroalkyl alkynyl ketimines, which is proposed by Trost et al. (scheme 24) [19] Several features about this reaction were reported: 1. Ketimines are more electrophilic, resulted by fluorinated groups 2. Alkyne group has less steric effect, and its high s character makes ketimines more electrophilic 3. Steric difference between alkynyl and perfluoroalkyl groups would give ketimines in a single diastereomeric form, which is critical for high stereoselectivities. The reaction is broad in scope, which allows further contribution in medicinal field and synthetic chemistry.

Scheme 24. Direct catalytic asymmetric vinylogous Mannich reaction of polyfluorinated alkynyl ketimines. Scheme 24.png
Scheme 24. Direct catalytic asymmetric vinylogous Mannich reaction of polyfluorinated alkynyl ketimines.

Chiral amines like amino acid derivatives can catalyze many enantioselective transformations, although enantiomers of amino acids are hard to access sometimes. In 2019, Lan et al. proposed a strategy to apply their chiral amine catalyst with minimal modification to the Mannich reaction of alkynyl ketiminoesters. (scheme 25) [20] Their application shows that catalyst 26 and MeCN can be replaced by 28 and dichloroethane for many alkynyl trifluoromethyl-, pyrazolinone-derived-, and isatin-derived ketimines to give the same stereoselectivities and yields.

Scheme 25. Enantiodivergent Mannich reaction of alkynyl ketiminoesters catalyzed by chiral amines. Scheme 25.png
Scheme 25. Enantiodivergent Mannich reaction of alkynyl ketiminoesters catalyzed by chiral amines.

It is hard to control E/Z selectivity for ketimines having two structurally similar substituents, which makes such kind of ketimines merely used in asymmetric catalysis field comparing to ketimines that have single diastereomeric form. In 2021, Kano et al. proposed method to use alkynyl ketimines as synthetic equivalents of dialkyl ketimines, and dialkyl ketimes are a typical example of isomeric ketimine mixtures. The alkynyl groups can be reduced by hydrogenation back to corresponding alkyl groups. Moreover, a single diastereomeric alkynyl alkyl N-protected ketimines can be obtained by using steric difference between alkyl and alkynyl group, which makes them be stereoselectively controlled by chiral catalyst to give asymmetric Mannich product. Additionally, small steric hindrance and high s character of alkynyl group result in its high nucleophilicity, making alkynyl-substituted ketimines more reactive than alkyl counterparts. Even though additional step of hydrogenation is needed, the advantages provided by alkynyl group are still impressive. To solve the problem that a single diastereomeric form of N-Boc alkyl alkynyl ketimines are difficult to access, Kano et al. proposed a synthesis to form these ketimines (reaction 1 and 2), followed by operating chiral amine-catalyzed Mannich reactions (reaction 3). (scheme 26) [21] The chiral amine catalyzed the asymmetric Mannich reaction of aldehyde and (Z)-alkyl alkynyl ketimines. The method they used is stereodivergent, in which the phenylcyclopropane amine 29 gives syn products, and the proline catalyst gives anti product. After the hydrogenation step, they successfully obtained chiral amines substituted with two similar alkyl groups, which is hard to obtained in a single enantiomeric compound.

Scheme 26. Chiral amine-catalyzed Mannich reactions of (Z)-alkynyl alkyl ketimines. Scheme 26.png
Scheme 26. Chiral amine-catalyzed Mannich reactions of (Z)-alkynyl alkyl ketimines.

Applications

Ketimine Mannich catalysis has been identified as a synthetic tool for primary alcohols. [2] A useful path to dihydroquinazolinone, a precursor to drugs with anti-viral and cardiovascular benefits, was also found by Liu [22] et Al. using N-Heterocyclic carbene-catalyzation [4+2] annulation of β-methyl enals.

Related Research Articles

The aldol reaction is a reaction that combines two carbonyl compounds to form a new β-hydroxy carbonyl compound.

<span class="mw-page-title-main">Enamine</span> Class of chemical compounds

An enamine is an unsaturated compound derived by the condensation of an aldehyde or ketone with a secondary amine. Enamines are versatile intermediates.

<span class="mw-page-title-main">Michael addition reaction</span> Reaction in organic chemistry

In organic chemistry, the Michael reaction or Michael 1,4 addition is a reaction between a Michael donor and a Michael acceptor to produce a Michael adduct by creating a carbon-carbon bond at the acceptor's β-carbon. It belongs to the larger class of conjugate additions and is widely used for the mild formation of carbon-carbon bonds.

The Robinson annulation is a chemical reaction used in organic chemistry for ring formation. It was discovered by Robert Robinson in 1935 as a method to create a six membered ring by forming three new carbon–carbon bonds. The method uses a ketone and a methyl vinyl ketone to form an α,β-unsaturated ketone in a cyclohexane ring by a Michael addition followed by an aldol condensation. This procedure is one of the key methods to form fused ring systems.

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">Corey–Itsuno reduction</span>

The Corey–Itsuno reduction, also known as the Corey–Bakshi–Shibata (CBS) reduction, is a chemical reaction in which a prochiral ketone is enantioselectively reduced to produce the corresponding chiral, non-racemic alcohol. The oxazaborolidine reagent which mediates the enantioselective reduction of ketones was previously developed by the laboratory of Itsuno and thus this transformation may more properly be called the Itsuno-Corey oxazaborolidine reduction.

<span class="mw-page-title-main">Nucleophilic conjugate addition</span> Organic reaction

Nucleophilic conjugate addition is a type of organic reaction. Ordinary nucleophilic additions or 1,2-nucleophilic additions deal mostly with additions to carbonyl compounds. Simple alkene compounds do not show 1,2 reactivity due to lack of polarity, unless the alkene is activated with special substituents. With α,β-unsaturated carbonyl compounds such as cyclohexenone it can be deduced from resonance structures that the β position is an electrophilic site which can react with a nucleophile. The negative charge in these structures is stored as an alkoxide anion. Such a nucleophilic addition is called a nucleophilic conjugate addition or 1,4-nucleophilic addition. The most important active alkenes are the aforementioned conjugated carbonyls and acrylonitriles.

<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">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 Hajos–Parrish–Eder–Sauer–Wiechert reaction in organic chemistry is a proline catalysed asymmetric aldol reaction. The reaction is named after the principal investigators of the two groups who reported it simultaneously: Zoltan Hajos and David Parrish from Hoffmann-La Roche and Rudolf Wiechert and co-workers from Schering AG. Discovered in the 1970s the original Hajos-Parrish catalytic procedure – shown in the reaction equation, leading to the optically active bicyclic ketol – paved the way of asymmetric organocatalysis. The Eder-Sauer-Wiechert modification lead directly to the optically active enedione, through the loss of water from the bicyclic ketol shown in figure.

Within the area of organocatalysis, (thio)urea organocatalysis describes the use of ureas and thioureas to accelerate and stereochemically alter organic transformations. The effects arise through hydrogen-bonding interactions between the substrate and the (thio)urea. Unlike classical catalysts, these organocatalysts interact by non-covalent interactions, especially hydrogen bonding. The scope of these small-molecule H-bond donors termed (thio)urea organocatalysis covers both non-stereoselective and stereoselective applications.

<span class="mw-page-title-main">Lectka enantioselective beta-lactam synthesis</span>

Lectka and co-workers developed a catalytic, asymmetric method to synthesize β-lactams.

In organic chemistry, the Baylis–Hillman, Morita–Baylis–Hillman, or MBH reaction is a carbon-carbon bond-forming reaction between an activated alkene and a carbon electrophile in the presence of a nucleophilic catalyst, such as a tertiary amine or phosphine. The product is densely functionalized, joining the alkene at the α-position to a reduced form of the electrophile.

In Lewis acid catalysis of organic reactions, a metal-based Lewis acid acts as an electron pair acceptor to increase the reactivity of a substrate. Common Lewis acid catalysts are based on main group metals such as aluminum, boron, silicon, and tin, as well as many early and late d-block metals. The metal atom forms an adduct with a lone-pair bearing electronegative atom in the substrate, such as oxygen, nitrogen, sulfur, and halogens. The complexation has partial charge-transfer character and makes the lone-pair donor effectively more electronegative, activating the substrate toward nucleophilic attack, heterolytic bond cleavage, or cycloaddition with 1,3-dienes and 1,3-dipoles.

<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.

Proline organocatalysis is the use of proline as an organocatalyst in organic chemistry. This theme is often considered the starting point for the area of organocatalysis, even though early discoveries went unappreciated. Modifications, such as MacMillan’s catalyst and Jorgensen's catalysts, proceed with excellent stereocontrol.

<span class="mw-page-title-main">Ugi's amine</span> Chemical compound

Ugi’s amine is a chemical compound named for the chemist who first reported its synthesis in 1970, Ivar Ugi. It is a ferrocene derivative. Since its first report, Ugi’s amine has found extensive use as the synthetic precursor to a large number of metal ligands that bear planar chirality. These ligands have since found extensive use in a variety of catalytic reactions. The compound may exist in either the 1S or 1R isomer, both of which have synthetic utility and are commercially available. Most notably, it is the synthetic precursor to the Josiphos class of ligands.

In organic chemistry, the Roskamp reaction is a name reaction describing the reaction between α-diazoesters (such as ethyl diazoacetate) and aldehydes to form β-ketoesters, often utilizing various Lewis acids (such as BF3, SnCl2, and GeCl2) as catalysts. The reaction is notable for its mild reaction conditions and selectivity.

<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.

References

  1. Bjørgo, Johannes; Boyd, Derek R.; Watson, Christopher G.; Jennings, W. Brian (1974). "Equilibrium distribution of E–Z-ketimine isomers". J. Chem. Soc., Perkin Trans. 2 (7): 757–762. doi:10.1039/p29740000757. ISSN   0300-9580.
  2. 1 2 3 4 List, Benjamin; Pojarliev, Peter; Biller, William T.; Martin, Harry J. (2010-05-20). "ChemInform Abstract: The Proline-Catalyzed Direct Asymmetric Three-Component Mannich Reaction: Scope, Optimization, and Application to the Highly Enantioselective Synthesis of 1,2-Amino Alcohols". ChemInform. 33 (29): no. doi:10.1002/chin.200229039. ISSN   0931-7597.
  3. Ishitani, Haruro; Ueno, Masaharu; Kobayashi, Shū (1997-07-01). "Catalytic Enantioselective Mannich-Type Reactions Using a Novel Chiral Zirconium Catalyst". Journal of the American Chemical Society. 119 (30): 7153–7154. doi:10.1021/ja970498d. ISSN   0002-7863.
  4. Hajos, Zoltan G.; Parrish, David R. (June 1974). "Asymmetric synthesis of bicyclic intermediates of natural product chemistry". The Journal of Organic Chemistry. 39 (12): 1615–1621. doi:10.1021/jo00925a003. ISSN   0022-3263.
  5. Kano, Taichi; Aota, Yusuke; Maruoka, Keiji (2017-11-27). "Asymmetric Synthesis of Less Accessible α-Tertiary Amines from Alkynyl Z-Ketimines". Angewandte Chemie International Edition. 56 (51): 16293–16296. doi:10.1002/anie.201710084. ISSN   1433-7851. PMID   29110376.
  6. Reddy, K. Nagarjuna; Rao, M. V. Krishna; Sridhar, B.; Subba Reddy, B. V. (2019-08-07). "BINOL Phosphoric Acid‐Catalyzed Asymmetric Mannich Reaction of Cyclic N‐Acyl Ketimines with Cyclic Enones". Chemistry: An Asian Journal. 14 (17): 2958–2965. doi:10.1002/asia.201900556. ISSN   1861-4728. PMID   31241858. S2CID   195658849.
  7. Wang, You-Qing; Zhang, Yongna; Pan, Kun; You, Junxiong; Zhao, Jin (2013-11-13). "Direct Organocatalytic Asymmetric Mannich Addition of 3-Substituted-2H-1,4-Benzoxazines: Access to Tetrasubstituted Carbon Stereocenters". Advanced Synthesis & Catalysis. 355 (17): 3381–3386. doi:10.1002/adsc.201300654. ISSN   1615-4150.
  8. Wu, Ling-Ling; Xiang, Yang; Yang, Da-Cheng; Guan, Zhi; He, Yan-Hong (2016). "Biocatalytic asymmetric Mannich reaction of ketimines using wheat germ lipase". Catalysis Science & Technology. 6 (11): 3963–3970. doi:10.1039/c5cy01923k. ISSN   2044-4753.
  9. Hu, Haipeng; Xu, Jinxiu; Liu, Wen; Dong, Shunxi; Lin, Lili; Feng, Xiaoming (2018-09-05). "Copper-Catalyzed Asymmetric Addition of Tertiary Carbon Nucleophiles to 2H-Azirines: Access to Chiral Aziridines with Vicinal Tetrasubstituted Stereocenters". Organic Letters. 20 (18): 5601–5605. doi:10.1021/acs.orglett.8b02274. ISSN   1523-7060. PMID   30183317. S2CID   52159388.
  10. Zhang, Hai-Jun; Xie, Yan-Cheng; Yin, Liang (2019-04-12). "Copper(I)-catalyzed asymmetric decarboxylative Mannich reaction enabled by acidic activation of 2H-azirines". Nature Communications. 10 (1): 1699. Bibcode:2019NatCo..10.1699Z. doi:10.1038/s41467-019-09750-5. ISSN   2041-1723. PMC   6461707 . PMID   30979892.
  11. Trost, Barry M.; Zhu, Chuanle (2020-12-03). "Zn-ProPhenol Catalyzed Enantioselective Mannich Reaction of 2H-Azirines with Alkynyl Ketones". Organic Letters. 22 (24): 9683–9687. doi:10.1021/acs.orglett.0c03737. ISSN   1523-7060. PMID   33269592. S2CID   227258883.
  12. Rong, Meng-Yu; Li, Jin-Shan; Zhou, Yin; Zhang, Fa-Guang; Ma, Jun-An (2020-12-31). "Correction to "Catalytic Enantioselective Synthesis of Difluoromethylated Tetrasubstituted Stereocenters in Isoindolones Enabled by a Multiple-Fluorine System"". Organic Letters. 23 (2): 623. doi: 10.1021/acs.orglett.0c04131 . ISSN   1523-7060. PMID   33382266. S2CID   229930648.
  13. Sawa, Masanao; Miyazaki, Shotaro; Yonesaki, Ryohei; Morimoto, Hiroyuki; Ohshima, Takashi (2018-08-14). "Catalytic Enantioselective Decarboxylative Mannich-Type Reaction of N-Unprotected Isatin-Derived Ketimines". Organic Letters. 20 (17): 5393–5397. doi:10.1021/acs.orglett.8b02306. ISSN   1523-7060. PMID   30106593. S2CID   52007498.
  14. Ding, Ransheng; De los Santos, Zeus A.; Wolf, Christian (2019-02-06). "Catalytic Asymmetric Mannich Reaction of α-Fluoronitriles with Ketimines: Enantioselective and Diastereodivergent Construction of Vicinal Tetrasubstituted Stereocenters". ACS Catalysis. 9 (3): 2169–2176. doi:10.1021/acscatal.8b05164. ISSN   2155-5435. PMC   6449049 . PMID   30956891.
  15. Kang, Tengfei; Hou, Liuzhen; Ruan, Sai; Cao, Weidi; Liu, Xiaohua; Feng, Xiaoming (2020-08-03). "Lewis acid-catalyzed asymmetric reactions of β,γ-unsaturated 2-acyl imidazoles". Nature Communications. 11 (1): 3869. Bibcode:2020NatCo..11.3869K. doi:10.1038/s41467-020-17681-9. ISSN   2041-1723. PMC   7398931 . PMID   32747706.
  16. Takeda, Tadahiro; Kondoh, Azusa; Terada, Masahiro (2016-03-08). "Construction of Vicinal Quaternary Stereogenic Centers by Enantioselective Direct Mannich-Type Reaction Using a Chiral Bis(guanidino)iminophosphorane Catalyst". Angewandte Chemie International Edition. 55 (15): 4734–4737. doi:10.1002/anie.201601352. ISSN   1433-7851. PMID   26954063.
  17. Lin, Shaoquan; Kumagai, Naoya; Shibasaki, Masakatsu (2016-02-02). "Direct Catalytic Asymmetric Mannich-type Reaction of α,β-Unsaturated γ-Butyrolactam with Ketimines". Chemistry: A European Journal. 22 (10): 3296–3299. doi:10.1002/chem.201600034. ISSN   0947-6539. PMID   26834086.
  18. Sawa, Masanao; Morisaki, Kazuhiro; Kondo, Yuta; Morimoto, Hiroyuki; Ohshima, Takashi (2017-09-26). "Direct Access to N-Unprotected α- and/or β-Tetrasubstituted Amino Acid Esters via Direct Catalytic Mannich-Type Reactions Using N-Unprotected Trifluoromethyl Ketimines". Chemistry: A European Journal. 23 (67): 17022–17028. doi:10.1002/chem.201703516. ISSN   0947-6539. PMID   28950035.
  19. Trost, Barry M.; Hung, Chao‐I. (Joey); Scharf, Manuel J. (2018-07-26). "Direct Catalytic Asymmetric Vinylogous Additions of α,β‐ and β,γ‐Butenolides to Polyfluorinated Alkynyl Ketimines". Angewandte Chemie International Edition. 57 (35): 11408–11412. doi: 10.1002/anie.201806249 . ISSN   1433-7851. PMID   29969528. S2CID   49677124.
  20. Dai, Jun; Wang, Zhuang; Deng, Yuhua; Zhu, Lei; Peng, Fangzhi; Lan, Yu; Shao, Zhihui (2019-11-15). "Enantiodivergence by minimal modification of an acyclic chiral secondary aminocatalyst". Nature Communications. 10 (1): 5182. Bibcode:2019NatCo..10.5182D. doi:10.1038/s41467-019-13183-5. ISSN   2041-1723. PMC   6858435 . PMID   31729388.
  21. Homma, Chihiro; Takeshima, Aika; Kano, Taichi; Maruoka, Keiji (2021). "Construction of chiral α-tert-amine scaffolds via amine-catalyzed asymmetric Mannich reactions of alkyl-substituted ketimines". Chemical Science. 12 (4): 1445–1450. doi: 10.1039/d0sc05269h . ISSN   2041-6520. PMC   8179053 . PMID   34163907. S2CID   229395873.
  22. Kano, Taichi; Song, Sunhwa; Kubota, Yasushi; Maruoka, Keiji (2011-12-15). "Highly Diastereo- and Enantioselective Mannich Reactions of Synthetically Flexible Ketimines with Secondary Amine Organocatalysts". Angewandte Chemie International Edition. 51 (5): 1191–1194. doi:10.1002/anie.201107375. ISSN   1433-7851. PMID   22173941.