Chiral auxiliary

Last updated

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. [1] [2] 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.

Contents

Auxiliary general scheme.png


General scheme for employing a chiral auxiliary in asymmetric synthesis

Most biological molecules and pharmaceutical targets exist as one of two possible enantiomers; consequently, chemical syntheses of natural products and pharmaceutical agents are frequently designed to obtain the target in enantiomerically pure form. [3] Chiral auxiliaries are one of many strategies available to synthetic chemists to selectively produce the desired stereoisomer of a given compound. [4]

Chiral auxiliaries were introduced by Elias James Corey in 1975 [5] with chiral 8-phenylmenthol and by Barry Trost in 1980 with chiral mandelic acid. The menthol compound is difficult to prepare and as an alternative trans-2-phenyl-1-cyclohexanol was introduced by J. K. Whitesell in 1985.

Asymmetric synthesis

Chiral auxiliaries are incorporated into synthetic routes to control the absolute configuration of stereogenic centers. David A. Evans' synthesis of the macrolide cytovaricin, considered a classic, utilizes oxazolidinone chiral auxiliaries for one asymmetric alkylation reaction and four asymmetric aldol reactions, setting the absolute stereochemistry of nine stereocenters. [6]

Cytovaricin analysis.png
Cytovaricin, synthesized by D. A. Evans in 1990. Blue and red bonds indicate stereocenters set with the use of chiral auxiliaries.

A typical auxiliary-guided stereoselective transformation involves three steps: first, the auxiliary is covalently coupled to the substrate; second, the resulting compound undergoes one or more diastereoselective transformations; and finally, the auxiliary is removed under conditions that do not cause racemization of the desired products. [4] The cost of employing stoichiometric auxiliary and the need to spend synthetic steps appending and removing the auxiliary make this approach appear inefficient. However, for many transformations, the only available stereoselective methodology relies on chiral auxiliaries. In addition, transformations with chiral auxiliaries tend to be versatile and very well-studied, allowing the most time-efficient access to enantiomerically pure products. [2]

Furthermore, [7] the products of auxiliary-directed reactions are diastereomers, which enables their facile separation by methods such as column chromatography or crystallization.

8-phenylmenthol

In an early example of the use of a chiral auxiliary in asymmetric synthesis, E. J. Corey and coworkers conducted an asymmetric Diels-Alder reaction between (−)-8-phenylmenthol acrylate ester and 5-benzyloxymethylcyclopentadiene. [5] The cycloaddition product was carried forward to the iodolactone shown below, an intermediate in the classic Corey synthesis of the prostaglandins. It is proposed that the back face of the acrylate is blocked by the auxiliary, so that cycloaddition occurs at the front face of the alkene.

8 phyenylmenthol acrylate ester diels alder.png
Diastereoselective Diels-Alder cycloaddition with the chiral auxiliary (−)-8-phenylmenthol in route to the prostaglandins

(−)-8-phenylmenthol can be prepared from either enantiomer of pulegone, [8] though neither route is very efficient. Because of the widespread utility of the 8-phenylmenthol auxiliary, alternative compounds that are more easily synthesized, such as trans-2-phenyl-1-cyclohexanol [9] and trans-2-(1-pheyl-1-methylethyl)cyclohexanol [10] have been explored.

1,1’-Binaphthyl-2,2’-diol (BINOL)

1,1’-Binaphthyl-2,2’-diol, or BINOL, has been used as chiral auxiliary for the asymmetric synthesis since 1983. [11] [12]

(R)-BINOL (R)-BINOL.svg
(R)-BINOL

Hisashi Yamamoto first utilized (R)-BINOL as a chiral auxiliary in the asymmetric synthesis of limonene, which is an example of cyclic mono-terpenes. (R)-BINOL mononeryl ether was prepared by the monosilylation and alkylation of (R)-BINOL as the chiral auxiliary. Followed with the reduction by organoaluminum reagent, limonene was synthesized with low yields (29% yield) and moderate enantiomeric excesses up to 64% ee. [12]

First utilization of BINOL as a chiral auxiliary BINOL chiral auxiliary 1.jpg
First utilization of BINOL as a chiral auxiliary

The preparation of a variety of enantiomerically pure uncommon R-amino acids can be achieved by the alkylation of chiral glycine derivatives possessing axially chiral BINOL as an auxiliary. It has been depicted by Fuji et al. Based on different electrophile, the diastereomeric excess varied from 69% to 86. [13]

Diastereoselective addition between Grignard and BINOL protected aldehyde BINOL chiral auxiliary.jpg
Diastereoselective addition between Grignard and BINOL protected aldehyde

Protected at the aldehyde function with (R)-BINOL, arylglyoxals reacted diastereoselectively with Grignard reagents to afford protected atrolactaldehyde with moderate to excellent diastereomeric excess and high yields. [14]

Diastereoselective addition between Grignard and BINOL protected aldehyde BINOL chiral auxiliary 3.jpg
Diastereoselective addition between Grignard and BINOL protected aldehyde

BINOL was also used as a chiral auxiliary to control the formation of a P-stereocenter in an asymmetric metal-catalyzed C-P coupling process. Mondal et al. discovered that the Pd-catalysed C-P cross-coupling reaction between axially chiral BINOL-based phosphoramidites and aryl halides or triflates proceeds with excellent stereoselectivity due to the presence of BINOL near the reacting P center. [15]

trans-2-Phenylcyclohexanol

Chemical structure of trans-2-phenylcyclohexanol Trans-2-phenylcyclohexanol.svg
Chemical structure of trans-2-phenylcyclohexanol

One type of chiral auxiliary is based on the trans-2-phenylcyclohexanol motif as introduced by James K. Whitesell and coworkers in 1985. This chiral auxiliary was used in ene reactions of the derived ester of glyoxylic acid. [16]

trans-2-Phenylcyclohexanol was used in ene reaction as a chiral auxiliary. TPC 1.jpg
trans-2-Phenylcyclohexanol was used in ene reaction as a chiral auxiliary.

In the total synthesis of (−)-heptemerone B and (−)-guanacastepene E, attached with trans-2-phenylcyclohexanol, the glyoxylate reacted with 2,4-dimethyl-pent-2-ene, in the presence of tin(IV) chloride, yielding the desired anti adduct as the major product, together with a small amount of its syn isomer with 10:1 diastereomeric ratio. [17]

The glyoxylate reacted with 2,4-dimethyl-2-pentane with trans-2-phenylcyclohexanol as a chiral auxiliary TPC 2.jpg
The glyoxylate reacted with 2,4-dimethyl-2-pentane with trans-2-phenylcyclohexanol as a chiral auxiliary

For even greater conformational control, switching from a phenyl to a trityl group gives trans-2-tritylcyclohexanol (TTC). In 2015, the Brown group published an efficient chiral permanganate-mediated oxidative cyclization with TTC. [18]

trans-2-Tritylcyclohexanol was used in asymmetric permanganate-mediated oxidative cyclization. TPC 3.jpg
trans-2-Tritylcyclohexanol was used in asymmetric permanganate-mediated oxidative cyclization.

Oxazolidinones

Oxazolidinone auxiliaries, popularized by David A. Evans, have been applied to many stereoselective transformations, including aldol reactions, [19] alkylation reactions, [20] and Diels-Alder reactions. [21] [22] The oxazolidinones are substituted at the 4 and 5 positions. Through steric hindrance, the substituents direct the direction of substitution of various groups. The auxiliary is subsequently removed e.g. through hydrolysis.

Preparation

Oxazolidinones can be prepared from amino acids or readily available amino alcohols. A large number of oxazolidinones are commercially available, including the four below.

Evans auxiliaries Evans-Auxiliar1.svg
Evans auxiliaries
Some of the commercially available oxazolidinone chiral auxiliaries

Acylation of the oxazolidinone is achieved by deprotonation with n-butyllithium and quench with an acyl chloride.

Oxazolidinone acylation.png
Acylation of a chiral oxazolidinone with propanoyl chloride

Alkylation reactions

Deprotonation at the α-carbon of an oxazolidinone imide with a strong base such as lithium diisopropylamide selectively furnishes the (Z)-enolate, which can undergo stereoselective alkylation.

Oxazolidinone imide benzylation.png
Alkylation of an oxazolidinone imide with benzyl bromide

Activated electrophiles, such as allylic or benzylic halides, are very good substrates.

Aldol reactions

Chiral oxazolidinones have been employed most widely in stereoselective aldol reactions.

Soft enolization with the Lewis acid dibutylboron triflate and the base diisopropylethylamine gives the (Z)-enolate, which undergoes a diastereoselective aldol reaction with an aldehyde substrate. The transformation is particularly powerful because it establishes two contiguous stereocenters simultaneously.

Evans syn aldol.png
Stereoselective Evans aldol reaction

A model for the observed stereoselectivity can be found below. The syn-stereo relationship between the methyl group and the new secondary alcohol results from a six-membered ring Zimmerman-Traxler transition state, wherein the enolate oxygen and the aldehyde oxygen both coordinate boron. The aldehyde is oriented such that the hydrogen is placed in a pseudo-axial orientation to minimize 1,3-diaxial interactions. The absolute stereochemistry of the two stereocenters is controlled by the chirality in the auxiliary. In the transition structure, the auxiliary carbonyl is oriented away from the enolate oxygen so as to minimize the net dipole of the molecule; one face of the enolate is blocked by the substituent on the chiral auxiliary.

Oxazolidinone transition structure.png


Model for the stereoselectivity of an Evans aldol reaction

Removal

A variety of transformations have been developed to facilitate removal of the oxazolidinone auxiliary to generate different synthetically useful functional groups.

Oxazolidinone derivatives.png
Transformation of an oxazolidinone imide to different functional groups

Camphorsultam

Camphorsultam, or Oppolzer's sultam, is a classic chiral auxiliary.

Camphorsultam Oppolzer sultam.svg
Camphorsultam

In the total synthesis of manzacidin B, Ohfune group utilized camphorsultam to construct the core oxazoline ring asymmetrically. Comparing with oxazolidinone as the chiral auxiliary, camphorsultam had a significant (2S,3R)-selectivity. [23]

The chiral camphorsultam was found to be a superior chiral auxiliary to the oxazolidinone in view of the single asymmetric induction. Camphorsultam 1.jpg
The chiral camphorsultam was found to be a superior chiral auxiliary to the oxazolidinone in view of the single asymmetric induction.

Camphorsultam also acts as a chiral auxiliary in Michael addition. Lithium base promoted stereoselective Michael addition of thiols to N-mcthacryloylcamphorsultam produced the corresponding addition products in high diastereoselectivity. [24]

Camphorsultam was used as a chiral auxiliary in Michael addition. Camphorsultam 2.jpg
Camphorsultam was used as a chiral auxiliary in Michael addition.

Camphorsultam was used as a chiral auxiliary for the asymmetric Claisen rearrangement. In the presence of butylated hydroxytoluene (BHT) used as a radical scavenger, a toluene solution of the adduct between geraniol and camphorsultam was heated in a sealed tube at 140 °C, to provide mainly the (2R,3S)-isomer as the major rearrangement product in 72% yield, securing the two contiguous stereocenters including the quaternary carbon. [25]

Camphorsultam was used as a chiral auxiliary in Claisen rearrangement. Camphorsultam 3.jpg
Camphorsultam was used as a chiral auxiliary in Claisen rearrangement.

Pseudoephedrine and pseudoephenamine

Both (R,R)- and (S,S)-pseudoephedrine can be used as chiral auxiliaries. [26] Pseudoephedrine is reacted with a carboxylic acid, acid anhydride, or acyl chloride to give the corresponding amide.

The α-proton of the carbonyl compound is easily deprotonated by a non-nucleophilic base to give the enolate, which can further react. The configuration of the addition compound, such as with an alkyl halide, is directed by the methyl group. Thus, any addition product will be syn with the methyl and anti to the hydroxyl group. The pseudoephedrine chiral auxiliary is subsequently removed by cleaving the amide bond with an appropriate nucleophile.

Preparation

Both enantiomers of pseudoephedrine are commercially available. Racemic pseudoephedrine has many medical uses. Because pseudoephedrine can be used to illegally make methamphetamine, the purchase of pseudoephedrine for use in academic or industrial research is rather regulated. As an alternative, Myers et al. reported the utility of pseudoephenamine chiral auxiliaries in alkylation reactions. [27] While pseudoephenamine is not readily available from commercial sources, it can be synthesized with relative ease from benzil and cannot be used to make amphetamines. [28]

Pseudoephedrine and pseudoephenamine.png
Pseudoephedrine and pseudoephenamine chiral auxiliaries

Pseudoephedrine amides are typically prepared by acylation with an acyl chloride or anhydride. [29]

Pseudoephedrine acylation.png


Acylation of (S,S)-psueodphedrine with an acid anhydride

Alkylation

Pseudoephedrine amides undergo deprotonation by a strong base such as lithium diisopropylamide (LDA) to give the corresponding (Z)-enolates. Alkylation of these lithium enolates proceeds with high facial selectivity.

Pseudoephedrine amide alkylation.png


Diastereoselective alkylation of a pseudoephedrine amide

The diastereoselectivity is believed to result from a configuration wherein one face of the lithium enolate is blocked by the secondary lithium alkoxide and the solvent molecules associated with that lithium cation. In accordance with this proposal, it has been observed that the diastereoselectivity of the alkylation step is highly dependent on the amount of lithium chloride present and on the solvent, tetrahydrofuran (THF). Typically, 4 to 6 equivalents of lithium chloride are sufficient to saturate a solution of enolate in THF at the reaction molarity.

Pseudoephedrine model.png


Model for the diastereoslective alkylation of a pseudoephedrine amide enolate

One primary advantage of asymmetric alkylation with pseudoephedrine amides is that the amide enolates are typically nucleophilic enough to react with primary and even secondary halides at temperatures ranging from –78 °C to 0 °C. Construction of quaternary carbon centers by alkylation of α-branched amide enolates is also possible, though the addition of DMPU is necessary for less reactive electrophiles. [30]

Removal

Conditions have been developed for the transformation of pseudoephedrine amides into enantiomerically enriched carboxylic acids, alcohols, aldehydes, and ketones - after cleavage, the auxiliary can be recovered and reused.

Pseudoephedrine derivatives.png


Transformation of pseudoephedrine amides into synthetically useful functional groups

tert-Butanesulfinamide

This specific sulfinamide chiral auxiliary was initially developed by Jonathan A. Ellman, and its use has been explored extensively by his group. [31] [32] Thus, it is often referred to as Ellman's auxiliary or Ellman's sulfinamide.

Sulfinamide enantiomers.png
The two possible enantiomers of tert-butanesulfinamide

Preparation

Either enantiomer of tert-butanesulfinamide can be reached from tert-butyl disulfide in two steps: a catalytic asymmetric oxidation reaction gives the disulfide oxidation product (thiosulfinate) in high yield and enantiomeric excess. Treatment of this compound with lithium amide in ammonia affords optically pure inverted product.

Sulfinamide synthesis.png


Synthesis of tert-butanesulfinamide from readily available chemicals

Condensation of tert-butanesulfinamide with an aldehyde or ketone proceeds in high yield and affords only the (E)-isomer of the corresponding N-sulfinyl imines.

Sulfinamide condensations.png


Condensation of tert-butanesulfinamide with aldehydes and ketones

Synthesis of chiral amines

Addition of a Grignard reagent to a tert-butanesulfinyl aldimine or ketimine results in asymmetric addition to give the branched sulfinamide. The observed stereoselectivity can be rationalized by a six-membered ring transition structure, wherein both oxygen and nitrogen of the sulfinyl imine coordinate magnesium.

Sulfinyl aldimine grignard addition.png


Addition of a Grignard reagent to a tert-butanesulfinyl aldimine

Removal

The auxiliary can be removed from the desired amine by treatment with hydrochloric acid in protic solvents.

Sulfinamide removal.png


Acidic cleavage of the sulfinamide auxiliary

SAMP/RAMP

Alkylation reactions of chiral (S)-1-amino-2-methoxymethylpyrrolidine (SAMP) and (R)-1-amino-2-methoxymethylpyrrolidine (RAMP) hydrazones were developed by Dieter Enders and E.J. Corey. [33] [34]

Preparation

SAMP can be prepared in six steps from (S)-proline, and RAMP can be prepared in six steps from (R)-glutamic acid.

EndersSamp.png


Preparation of SAMP and RAMP from commercially available materials.

Alkylation reactions

Condensation of SAMP or RAMP with an aldehyde or ketone affords the (E)-hydrazine. Deprotonation with lithium diisopropylamide and addition of an alkyl halide affords the alkylated product. The auxiliary can be removed by ozonolysis or hydrolysis.

EndersReaction.png


Condensation, alkylation, and cleavage with SAMP/RAMP chiral auxiliaries

Chiral auxiliaries in industry

Chiral auxiliaries are generally reliable and versatile, enabling the synthesis of a large number of enantiomerically pure compounds in a time-efficient manner. Consequently, chiral auxiliaries are often the method of choice in the early phases of drug development. [2]

Tipranavir

The HIV protease inhibitor Tipranavir is marketed for the treatment of AIDS. The first enantioselective medicinal chemistry route to Tipranavir included the conjugate addition of an organocuprate reagent to a chiral Michael acceptor. [35] The chiral oxazolidinone in the Michael acceptor controlled the stereochemistry of one of two stereocenters in the molecule. The final, commercial route to Tipranavir does not feature a chiral auxiliary; instead, this stereocenter is set by an asymmetric hydrogenation reaction. [36]

Tipranavir routes.png
Synthetic strategies for setting a key stereocenter in Tipranavir

Atorvastatin

The calcium salt of atorvastatin is marketed under the trade name Lipitor for the lowering of blood cholesterol. The first enantioselective medicinal chemistry route to atorvastatin relied on a diastereoselective aldol reaction with a chiral ester to set one of the two alcohol stereocenters. [37] In the commercial route to atorvastatin, this stereocenter is carried forward from the readily available food additive isoascorbic acid. [38]

Atorvastatin routes.png
Synthetic strategies for setting a key stereocenter in atorvastatin

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.

<span class="mw-page-title-main">Organolithium reagent</span> Chemical compounds containing C–Li bonds

In organometallic chemistry, organolithium reagents are chemical compounds that contain carbon–lithium (C–Li) bonds. These reagents are important in organic synthesis, and are frequently used to transfer the organic group or the lithium atom to the substrates in synthetic steps, through nucleophilic addition or simple deprotonation. Organolithium reagents are used in industry as an initiator for anionic polymerization, which leads to the production of various elastomers. They have also been applied in asymmetric synthesis in the pharmaceutical industry. Due to the large difference in electronegativity between the carbon atom and the lithium atom, the C−Li bond is highly ionic. Owing to the polar nature of the C−Li bond, organolithium reagents are good nucleophiles and strong bases. For laboratory organic synthesis, many organolithium reagents are commercially available in solution form. These reagents are highly reactive, and are sometimes pyrophoric.

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

In organic chemistry, the ene reaction is a chemical reaction between an alkene with an allylic hydrogen and a compound containing a multiple bond, in order to form a new σ-bond with migration of the ene double bond and 1,5 hydrogen shift. The product is a substituted alkene with the double bond shifted to the allylic position.

In chemistry, stereoselectivity is the property of a chemical reaction in which a single reactant forms an unequal mixture of stereoisomers during a non-stereospecific creation of a new stereocenter or during a non-stereospecific transformation of a pre-existing one. The selectivity arises from differences in steric and electronic effects in the mechanistic pathways leading to the different products. Stereoselectivity can vary in degree but it can never be total since the activation energy difference between the two pathways is finite. Both products are at least possible and merely differ in amount. However, in favorable cases, the minor stereoisomer may not be detectable by the analytic methods used.

<span class="mw-page-title-main">Enolate</span> Organic anion formed by deprotonating a carbonyl (>C=O) compound

In organic chemistry, enolates are organic anions derived from the deprotonation of carbonyl compounds. Rarely isolated, they are widely used as reagents in the synthesis of organic compounds.

<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">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">Asymmetric induction</span> Preferential formation of one chiral isomer over another in a chemical reaction

In stereochemistry, asymmetric induction describes the preferential formation in a chemical reaction of one enantiomer or diastereoisomer over the other as a result of the influence of a chiral feature present in the substrate, reagent, catalyst or environment. Asymmetric induction is a key element in asymmetric synthesis.

The Nazarov cyclization reaction is a chemical reaction used in organic chemistry for the synthesis of cyclopentenones. The reaction is typically divided into classical and modern variants, depending on the reagents and substrates employed. It was originally discovered by Ivan Nikolaevich Nazarov (1906–1957) in 1941 while studying the rearrangements of allyl vinyl ketones.

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.

The [2,3]-Wittig rearrangement is the transformation of an allylic ether into a homoallylic alcohol via a concerted, pericyclic process. Because the reaction is concerted, it exhibits a high degree of stereocontrol, and can be employed early in a synthetic route to establish stereochemistry. The Wittig rearrangement requires strongly basic conditions, however, as a carbanion intermediate is essential. [1,2]-Wittig rearrangement is a competitive process.

<i>tert</i>-Butanesulfinamide Chemical compound

tert-Butanesulfinamide is an organosulfur compound and a member of the class of sulfinamides. Both enantiomeric forms are commercially available and are used in asymmetric synthesis as chiral auxiliaries, often as chiral ammonia equivalents for the synthesis of amines. tert-Butanesulfinamide and the associated synthetic methodology was introduced in 1997 by Jonathan A. Ellman et al.

Nucleophilic epoxidation is the formation of epoxides from electron-deficient double bonds through the action of nucleophilic oxidants. Nucleophilic epoxidation methods represent a viable alternative to electrophilic methods, many of which do not epoxidize electron-poor double bonds efficiently.

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">Enders SAMP/RAMP hydrazone-alkylation reaction</span>

The Enders SAMP/RAMP hydrazone alkylation reaction is an asymmetric carbon-carbon bond formation reaction facilitated by pyrrolidine chiral auxiliaries. It was pioneered by E. J. Corey and D. Enders in 1976, and was further developed by D. Enders and his group. This method is usually a three-step sequence. The first step is to form the hydrazone between (S)-1-amino-2-methoxymethylpyrrolidine (SAMP) or (R)-1-amino-2-methoxymethylpyrrolidine (RAMP) and a ketone or aldehyde. Afterwards, the hydrazone is deprotonated by lithium diisopropylamide (LDA) to form an azaenolate, which reacts with alkyl halides or other suitable electrophiles to give alkylated hydrazone species with the simultaneous generation of a new chiral center. Finally, the alkylated ketone or aldehyde can be regenerated by ozonolysis or hydrolysis.

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.

<i>N</i>-Sulfinyl imine

N-Sulfinyl imines are a class of imines bearing a sulfinyl group attached to nitrogen. These imines display useful stereoselectivity reactivity and due to the presence of the chiral electron withdrawing N-sulfinyl group. They allow 1,2-addition of organometallic reagents to imines. The N-sulfinyl group exerts powerful and predictable stereodirecting effects resulting in high levels of asymmetric induction. Racemization of the newly created carbon-nitrogen stereo center is prevented because anions are stabilized at nitrogen. The sulfinyl chiral auxiliary is readily removed by simple acid hydrolysis. The addition of organometallic reagents to N-sulfinyl imines is the most reliable and versatile method for the asymmetric synthesis of amine derivatives. These building blocks have been employed in the asymmetric synthesis of numerous biologically active compounds.

In organic chemistry, the Keck asymmetric allylation is a chemical reaction that involves the nucleophilic addition of an allyl group to an aldehyde. The catalyst is a chiral complex that contains titanium as a Lewis acid. The chirality of the catalyst induces a stereoselective addition, so the secondary alcohol of the product has a predictable absolute stereochemistry based on the choice of catalyst. This name reaction is named for Gary Keck.

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. Key Chiral Auxiliary Applications (Second Edition)(ed.: Roos, G.), Academic Press, Boston, 2014 ISBN   978-0-12-417034-6
  2. 1 2 3 Glorius, F.; Gnas, Y. (2006). "Chiral Auxiliaries — Principles and Recent Applications". Synthesis . 2006 (12): 1899–1930. doi:10.1055/s-2006-942399.
  3. Jamali, Fakhreddin (1993). "Chapter 14: Stereochemically Pure Drugs: An Overview". In Wainer, Irving W. (ed.). Drug Stereochemistry: Analytical Methods and Pharmacology . Marcel Dekker, Inc. pp.  375–382. ISBN   978-0-8247-8819-3.
  4. 1 2 Evans, D. A.; Helmchen, G.; Rüping, M. (2007). "Chiral Auxiliaries in Asymmetric Synthesis". In Christmann, M (ed.). Asymmetric Synthesis — The Essentials. Wiley-VCH Verlag GmbH & Co. pp. 3–9. ISBN   978-3-527-31399-0.
  5. 1 2 Corey, E. J.; Ensley, H. E. (1975). "Preparation of an Optically Active Prostaglandin Intermediate via Asymmetric Induction". J. Am. Chem. Soc. 97 (23): 6908–6909. doi:10.1021/ja00856a074. PMID   1184891.
  6. Nicolau, K. C. (2008). Classics in Total Synthesis (5th ed.). New York, New York: Wiley-VCH. pp. 485–508. ISBN   978-3-527-29231-8.
  7. Miller, J. P. (2013). "ChemInform Abstract: Recent Advances in Asymmetric Diels-Alder Reactions". ChemInform . 44 (48): no. doi:10.1002/chin.201348243.
  8. Corey, E. J.; Ensley, H. E.; Parnell, C. A. (1978). "Convenient Synthesis of a Highly Efficient and Recyclable Chiral Director for Asymmetric Induction". J. Org. Chem. 43 (8): 1610–1611. doi:10.1021/jo00402a037.
  9. Whitesell, J. K.; Chen, H. H.; Lawrence, R. M. (1985). "trans-2-Phenylcyclohexanol. A powerful and readily available chiral auxiliary". J. Org. Chem. 50 (23): 4663–4664. doi:10.1021/jo00223a055.
  10. Comins, D. L.; Salvador, J. D. (1993). "Efficient Synthesis and Resolution of trans-2-(1-Aryl-1-methylethyl)cyclohexanols: Practical Alternatives to 8-P henylmenthol". J. Org. Chem. 58 (17): 4656–4661. doi:10.1021/jo00069a031.
  11. Brunel, Jean Michel (2005). "BINOL: A Versatile Chiral Reagent". Chemical Reviews. 105 (3): 857–898. doi:10.1021/cr040079g. PMID   15755079.
  12. 1 2 Sakane, Soichi; Fujiwara, Junya; Maruoka, Keiji; Yamamoto, Hisashi (1983). "Chiral leaving group. Biogenetic-type asymmetric synthesis of limonene and bisabolenes". Journal of the American Chemical Society. 105 (19): 6154–6155. doi:10.1021/ja00357a033.
  13. Tanaka, Kiyoshi; Ahn, Mija; Watanabe, Yukari; Fuji, Kaoru (1996-06-01). "Asymmetric synthesis of uncommon α-amino acids by diastereoselective alkylations of a chiral glycine equivalent". Tetrahedron: Asymmetry. 7 (6): 1771–1782. doi:10.1016/0957-4166(96)00212-1.
  14. Maglioli, Paola; De Lucchi, Ottorino; Delogu, Giovanna; Valle, Giovanni (1992-01-01). "Highly diastereoselective reduction and addition of nucleophiles to binaphthol-protected arylglyoxals". Tetrahedron: Asymmetry. 3 (3): 365–366. doi:10.1016/S0957-4166(00)80276-1.
  15. Mondal, Anirban; Thiel, Niklas O.; Dorel, Ruth; Feringa, Ben L. (January 2022). "P-chirogenic phosphorus compounds by stereoselective Pd-catalysed arylation of phosphoramidites". Nature Catalysis. 5 (1): 10–19. doi:10.1038/s41929-021-00697-9. S2CID   245426891.
  16. Buchi, George; Vogel, Dennis E. (1985). "A new method for the preparation of γ,δ-unsaturated ketones via Claisen rearrangement". The Journal of Organic Chemistry. 50 (23): 4664–4665. doi:10.1021/jo00223a056.
  17. Miller, Aubry K.; Hughes, Chambers C.; Kennedy-Smith, Joshua J.; Gradl, Stefan N.; Dirk Trauner (2006). "Total Synthesis of (−)-Heptemerone B and (−)-Guanacastepene E". Journal of the American Chemical Society. 128 (51): 17057–17062. doi:10.1021/ja0660507. PMID   17177458.
  18. Al Hazmi, Ali M.; Sheikh, Nadeem S.; Bataille, Carole J. R.; Al-Hadedi, Azzam A. M.; Watkin, Sam V.; Luker, Tim J.; Camp, Nicholas P.; Brown, Richard C. D. (2014). "trans-2-Tritylcyclohexanol as a Chiral Auxiliary in Permanganate-Mediated Oxidative Cyclization of 2-Methylenehept-5-enoates: Application to the Synthesis of trans-(+)-Linalool Oxide". Organic Letters. 16 (19): 5104–5107. doi:10.1021/ol502454r. PMID   25225741.
  19. Evans, D. A.; Bartroli, J.; Shih, T. L. (1981). "Enantioselective aldol condensations. 2. Erythro-selective chiral aldol condensations via boron enolates". J. Am. Chem. Soc. 103 (8): 2127–2129. doi:10.1021/ja00398a058.
  20. Evans, D. A.; Ennis, M. D.; Mathre, D. J. (1982). "Asymmetric Alkylation Reactions of Chiral Imide Enolates. A Practical Approach to the Enantioselective Synthesis of a-Substituted Carboxylic Acid Derivatives". J. Am. Chem. Soc. 104 (6): 1737–1739. doi:10.1021/ja00370a050.
  21. Evans, D. A.; Chapman, K. T.; Bisaha, J. (1984). "New Asymmetric Diels-Alder Cycloaddition Reactions. Chiral α,β-Unsaturated Carboximides as Practical Chiral Acrylate and Crotonate Dienophile Synthons". J. Am. Chem. Soc. 106 (15): 4261–4263. doi:10.1021/ja00327a031.
  22. Evans, D. A.; Chapman, K. T.; Hung, D. T.; Kawaguchi, A. T. (1987). "Transition State π-Solvation by Aromatic Rings: An Electronic Contribution to Diels-Alder Reaction Diastereoselectivity". Angew. Chem. Int. Ed. 26 (11): 1184–1186. doi:10.1002/anie.198711841.
  23. Shinada, Tetsuro; Oe, Kentaro; Ohfune, Yasufumi (2012-06-27). "Efficient total synthesis of manzacidin B". Tetrahedron Letters. 53 (26): 3250–3253. doi:10.1016/j.tetlet.2012.04.042.
  24. Tsai, Wen-Jiuan; Lin, Yi-Tsong; Uang, Biing-Jiun (1994-07-01). "Asymmetric Michael addition of thiols to (1R,2R,4R)-(−)-2,10-N-enoylcamphorsultam". Tetrahedron: Asymmetry. 5 (7): 1195–1198. doi:10.1016/0957-4166(94)80155-X.
  25. Takao, Ken-ichi; Sakamoto, Shu; Touati, Marianne Ayaka; Kusakawa, Yusuke; Tadano, Kin-ichi (2012-11-08). "Asymmetric Construction of All-Carbon Quaternary Stereocenters by Chiral-Auxiliary-Mediated Claisen Rearrangement and Total Synthesis of (+)-Bakuchiol". Molecules. 17 (11): 13330–13344. doi: 10.3390/molecules171113330 . PMC   6268616 . PMID   23138536.
  26. Myers, A. G.; et al. (1997). "Pseudoephedrine as a Practical Chiral Auxiliary for the Synthesis of Highly Enantiomerically Enriched Carboxylic Acids, Alcohols, Aldehydes, and Ketones". J. Am. Chem. Soc. 119 (28): 6496–6511. doi:10.1021/ja970402f.
  27. Myers, A. G.; Morales, M. R.; Mellem, K. T. (2012). "Pseudoephenamine: A Practical Chiral Auxiliary for Asymmetric Synthesis" (PDF). Angew. Chem. 124 (19): 4646–4649. Bibcode:2012AngCh.124.4646M. doi:10.1002/ange.201200370. PMC   3854953 . PMID   22461381.
  28. Mellem, Kevin T.; Myers, Andrew G. (2013). "A Simple, Scalable Synthetic Route to (+)- and (−)-Pseudoephenamine". Organic Letters. 15 (21): 5594–5597. doi:10.1021/ol402815d. ISSN   1523-7060. PMC   3864801 . PMID   24138164.
  29. Myers, A. G.; Yang, B. H.; McKinstry, L.; Kopecky, D. J.; Gleason, J. L. (1997). "Pseudoephedrine as a Practical Chiral Auxiliary for the Synthesis of Highly Enantiomerically Enriched Carboxylic Acids, Alcohols, Aldehydes, and Ketones". J. Am. Chem. Soc. 119 (28): 6496–6511. doi:10.1021/ja970402f.
  30. Kummer, D. A.; Chain, W. J.; Morales, M. R.; Quiroga, O.; Myers, A. G. (2008). "Stereocontrolled Alkylative Construction of Quaternary Carbon Centers". J. Am. Chem. Soc. 130 (40): 13231–13233. doi:10.1021/ja806021y. PMC   2666470 . PMID   18788739.
  31. Liu, Guangcheng; Cogan, Derek A.; Ellman, Jonathan A. (October 1997). "Catalytic Asymmetric Synthesis of tert -Butanesulfinamide. Application to the Asymmetric Synthesis of Amines". Journal of the American Chemical Society. 119 (41): 9913–9914. doi:10.1021/ja972012z. ISSN   0002-7863.
  32. Ellman, J. A.; Owens, T. D.; Tang, T. P. (2002). "N-tert-Butanesulfinyl Imines: Versatile Intermediates for the Asymmetric Synthesis of Amines". Acc. Chem. Res. 35 (11): 984–995. doi:10.1021/ar020066u. PMID   12437323.
  33. Corey, E. J.; Enders, D. (1976). "Applications of N,N-dimethylhydrazones to synthesis. Use in efficient, positionally and stereochemically selective C-C bond formation; oxidative hydrolysis to carbonyl compounds". Tetrahedron Letters . 17 (1): 3–6. doi:10.1016/s0040-4039(00)71307-4.
  34. Kurti, L.; Czako, B. (2005). Strategic Applications of Named Reactions in Organic Synthesis. Burlington, MA: Elsevier Academic Press. pp. 150–151. ISBN   978-0-12-369483-6.
  35. Turner, S. T.; et al. (1998). "Tipranavir (PNU-140690): A Potent, Orally Bioavailable Nonpeptidic HIV Protease Inhibitor of the 5,6-Dihydro-4-hydroxy-2-pyrone Sulfonamide Class". J. Med. Chem. 41 (18): 3467–3476. doi:10.1021/jm9802158. PMID   9719600.
  36. Caron, Stéphane (2011). "Chapter 15: Synthetic Route Development of Selected Contemporary Pharmaceutical Drugs". In Caron, Stéphane (ed.). Practical Synthetic Organic Chemistry . John Wiley & Sons, Inc. pp.  666–670. ISBN   978-0-470-03733-1.
  37. Roth, B. D.; et al. (1991). "Inhibitors of Cholesterol Biosynthesis. 3. Tetrahydro-4-hydroxy-6-[2-(lH-pyrrol-l-yl)ethyl]-2H-pyran-2-one Inhibitors of HMG-CoA Reductase. 2. Effects of Introducing Substituents at Positions Three and Four of the Pyrrole Nucleus". J. Med. Chem. 34 (1): 357–366. doi:10.1021/jm00105a056. PMID   1992137.
  38. Jie Jack Li; Douglas S. Johnson; Drago R. Sliskovic; Bruce D. Roth (2004). "Chapter 9. Atorvastatin Calcium (Lipitor)". Contemporary Drug Synthesis. John Wiley & Sons, Inc. pp. 113–125. ISBN   978-0-471-21480-9.