Weinreb ketone synthesis

Last updated

Contents

Weinreb ketone synthesis
Named after Steven M. Weinreb
Reaction type Coupling reaction
Identifiers
Organic Chemistry Portal weinreb-ketone-synthesis

The Weinreb ketone synthesis or 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. [1] 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 (see amide reduction).

The Weinreb-Nahm ketone synthesis Weinreb ketone synthesis.svg
The Weinreb–Nahm ketone synthesis

The major advantage of this method over addition of organometallic reagents to more typical acyl compounds is that it avoids the common problem of over-addition. For these latter reactions, two equivalents of the incoming group add to form an alcohol rather than a ketone or aldehyde. This occurs even if the equivalents of nucleophile are closely controlled.

Overaddition of nucleophiles Overaddition of organometallic reagent.svg
Overaddition of nucleophiles

The Weinreb–Nahm amide has since been adopted into regular use by organic chemists as a dependable method for the synthesis of ketones. These functional groups are present in a large number of natural products and can be reliably reacted to form new carbon–carbon bonds or converted into other functional groups. This method has been used in a number of syntheses, including macrosphelides A and B, [2] amphidinolide J, [3] and spirofungins A and B. [4]

Mechanism

Weinreb and Nahm originally proposed the following reaction mechanism to explain the selectivity shown in reactions of the Weinreb–Nahm amide. Their suggestion was that the tetrahedral intermediate (A below) formed as a result of nucleophilic addition by the organometallic reagent is stabilized by chelation from the methoxy group as shown. [1] This intermediate is stable only at low temperatures, requiring a low-temperature quench.

Chelation mechanism Weinreb ketone synthesis mechanism.svg
Chelation mechanism

This chelation is in contrast to the mechanism for formation of the over-addition product wherein collapse of the tetrahedral intermediate allows a second addition. The mechanistic conjecture on the part of Weinreb was immediately accepted by the academic community, but it was not until 2006 that it was confirmed by spectroscopic and kinetic analyses. [5]

Preparation

In addition to the original procedure shown above (which may have compatibility issues for sensitive substrates), Weinreb amides can be synthesized from a variety of acyl compounds. The vast majority of these procedures utilize the commercially available salt N,O-dimethylhydroxylamine hydrochloride [MeO(Me)NH•HCl], which is typically easier to handle than the free amine. [6]

Treatment of an ester or lactone with AlMe3 or AlMe2Cl affords the corresponding Weinreb amide in good yields. Alternatively, non-nucleophilic Grignard reagents such as isopropyl magnesium chloride can be used to activate the amine before addition of the ester. [7]

Example of syntheses from esters and lactones Weinrebester.png
Example of syntheses from esters and lactones

A variety of peptide coupling reagents can also be used to prepare Weinreb–Nahm amides from carboxylic acids. Various carbodiimide-, hydroxybenzotriazole-, and triphenylphosphine-based couplings have been reported specifically for this purpose. [6] [7]

Example of Syntheses from Carboxyllic Acids Weinrebacid.png
Example of Syntheses from Carboxyllic Acids

Finally, an aminocarbonylation reaction reported by Stephen Buchwald allows conversion of aryl halides directly into aryl Weinreb–Nahm amides. [8]

Aminocarbonylation to form Weinreb-Nahm amides Weinrebaminocarb.png
Aminocarbonylation to form Weinreb–Nahm amides

Scope

The standard conditions for the Weinreb–Nahm ketone synthesis are known to tolerate a wide variety of functional groups elsewhere in the molecule, including alpha-halogen substitution, N-protected amino acids, α-β unsaturation, silyl ethers, various lactams and lactones, sulfonates, sulfinates, and phosphonate esters. [6] [7] A wide variety of nucleophiles can be used in conjunction with the amide. Lithiates and Grignard reagents are most commonly employed; examples involving aliphatic, vinyl, aryl, and alkynyl carbon nucleophiles have been reported. However, with highly basic or sterically hindered nucleophiles, elimination of the methoxide moiety to release formaldehyde can occur as a significant side reaction. [9]

Side reaction Weinrebelim.png
Side reaction

Nonetheless, the Weinreb–Nahm amide figures prominently into many syntheses, serving as an important coupling partner for various fragments. Shown below are key steps involving Weinreb amides in the synthesis of several natural products, including members of the immunosuppressant family of macrosphelides, and the antibiotic family of spirofungins. [2] [3] [4]

Syntheses using Weinreb-Nahm amide Weinrebsyntheses.png
Syntheses using Weinreb–Nahm amide

Variations

Reaction of Weinreb–Nahm amides with Wittig reagents has been performed to avoid the sometimes harsh conditions required for addition of hydride reagents or organometallic compounds. This yields an N-methyl-N-methoxy-enamine that converts to the corresponding ketone or aldehyde upon hydrolytic workup. [10]

Reaction of Weinreb-Nahm amides with Wittig reagents WeinrebWittig.png
Reaction of Weinreb–Nahm amides with Wittig reagents

Additionally, a one-pot magnesium–halogen exchange with subsequent arylation has been developed, showcasing the stability of the Weinreb–Nahm amide and providing an operationally simple method for the synthesis of aryl ketones. [11]

One-pot arylation reaction Weinrebonepot.png
One-pot arylation reaction

More unusual reagents with multiple Weinreb–Nahm amide functional groups have been synthesized, serving as CO2 and α-diketone synthons. [12] [13]

Synthons based on Weinreb-Nahm-amides Weinrebsynthons.png
Synthons based on Weinreb–Nahm-amides

Finally, Stephen G. Davies of Oxford has designed a chiral auxiliary that combines the functionality of the Weinreb amide with that of the Myers' pseudoephedrine auxiliary, allowing diastereoselective enolate alkylation followed by facile cleavage to the corresponding enantioenriched aldehyde or ketone. [14]

Davies' Auxiliary with Weinreb-Nahm-like functionality Weinrebdavies.png
Davies' Auxiliary with Weinreb–Nahm-like functionality

See also

Related Research Articles

<span class="mw-page-title-main">Grignard reaction</span> Organometallic coupling reaction

The Grignard reaction is an organometallic chemical reaction in which, according to the classical definition, carbon alkyl, allyl, vinyl, or aryl magnesium halides are added to the carbonyl groups of either an aldehyde or ketone under anhydrous conditions. This reaction is important for the formation of carbon-carbon bonds.

<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.

A tetrahedral intermediate is a reaction intermediate in which the bond arrangement around an initially double-bonded carbon atom has been transformed from trigonal to tetrahedral. Tetrahedral intermediates result from nucleophilic addition to a carbonyl group. The stability of tetrahedral intermediate depends on the ability of the groups attached to the new tetrahedral carbon atom to leave with the negative charge. Tetrahedral intermediates are very significant in organic syntheses and biological systems as a key intermediate in esterification, transesterification, ester hydrolysis, formation and hydrolysis of amides and peptides, hydride reductions, and other chemical reactions.

<span class="mw-page-title-main">Bamford–Stevens reaction</span> Synthesis of alkenes by base-catalysed decomposition of tosylhydrazones

The Bamford–Stevens reaction is a chemical reaction whereby treatment of tosylhydrazones with strong base gives alkenes. It is named for the British chemist William Randall Bamford and the Scottish chemist Thomas Stevens Stevens (1900–2000). The usage of aprotic solvents gives predominantly Z-alkenes, while protic solvent gives a mixture of E- and Z-alkenes. As an alkene-generating transformation, the Bamford–Stevens reaction has broad utility in synthetic methodology and complex molecule synthesis.

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

1,1'-Carbonyldiimidazole (CDI) is an organic compound with the molecular formula (C3H3N2)2CO. It is a white crystalline solid. It is often used for the coupling of amino acids for peptide synthesis and as a reagent in organic synthesis.

The Reformatsky reaction is an organic reaction which condenses aldehydes or ketones with α-halo esters using metallic zinc to form β-hydroxy-esters:

Nucleophilic acyl substitution describes a class of substitution reactions involving nucleophiles and acyl compounds. In this type of reaction, a nucleophile – such as an alcohol, amine, or enolate – displaces the leaving group of an acyl derivative – such as an acid halide, anhydride, or ester. The resulting product is a carbonyl-containing compound in which the nucleophile has taken the place of the leaving group present in the original acyl derivative. Because acyl derivatives react with a wide variety of nucleophiles, and because the product can depend on the particular type of acyl derivative and nucleophile involved, nucleophilic acyl substitution reactions can be used to synthesize a variety of different products.

<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">Dakin oxidation</span> Organic redox reaction that converts hydroxyphenyl aldehydes or ketones into benzenediols

The Dakin oxidation (or Dakin reaction) is an organic redox reaction in which an ortho- or para-hydroxylated phenyl aldehyde (2-hydroxybenzaldehyde or 4-hydroxybenzaldehyde) or ketone reacts with hydrogen peroxide (H2O2) in base to form a benzenediol and a carboxylate. Overall, the carbonyl group is oxidised, whereas the H2O2 is reduced.

<span class="mw-page-title-main">Grignard reagent</span> Organometallic compounds used in organic synthesis

In organic chemistry, a Grignard reagent or Grignard compound is a chemical compound with the general formula R−Mg−X, where X is a halogen and R is an organic group, normally an alkyl or aryl. Two typical examples are methylmagnesium chloride Cl−Mg−CH3 and phenylmagnesium bromide (C6H5)−Mg−Br. They are a subclass of the organomagnesium compounds.

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

Schwartz's reagent is the common name for the organozirconium compound with the formula (C5H5)2ZrHCl, sometimes called zirconocene hydrochloride or zirconocene chloride hydride, and is named after Jeffrey Schwartz, a chemistry professor at Princeton University. This metallocene is used in organic synthesis for various transformations of alkenes and alkynes.

<span class="mw-page-title-main">Organozinc chemistry</span>

Organozinc chemistry is the study of the physical properties, synthesis, and reactions of organozinc compounds, which are organometallic compounds that contain carbon (C) to zinc (Zn) chemical bonds.

The Kulinkovich reaction describes the organic synthesis of substituted cyclopropanols through reaction of esters with dialkyl­dialkoxy­titanium reagents, which are generated in situ from Grignard reagents containing a hydrogen in beta-position and titanium(IV) alkoxides such as titanium isopropoxide. This reaction was first reported by Oleg Kulinkovich and coworkers in 1989.

Organomanganese chemistry is the chemistry of organometallic compounds containing a carbon to manganese chemical bond. In a 2009 review, Cahiez et al. argued that as manganese is cheap and benign, organomanganese compounds have potential as chemical reagents, although currently they are not widely used as such despite extensive research.

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.

<span class="mw-page-title-main">Carbonyl reduction</span> Organic reduction of any carbonyl group by a reducing agent

In organic chemistry, carbonyl reduction is the conversion of any carbonyl group, usually to an alcohol. It is a common transformation that is practiced in many ways. Ketones, aldehydes, carboxylic acids, esters, amides, and acid halides - some of the most pervasive functional groups, -comprise carbonyl compounds. Carboxylic acids, esters, and acid halides can be reduced to either aldehydes or a step further to primary alcohols, depending on the strength of the reducing agent. Aldehydes and ketones can be reduced respectively to primary and secondary alcohols. In deoxygenation, the alcohol group can be further reduced and removed altogether by replacement with H.

Steven M. Weinreb is an American chemist and is a professor of chemistry at Pennsylvania State University in United States. Together with Steven Nahm, he developed the Weinreb ketone synthesis, which allows for mono-addition of an organometallic reagent such as a Grignard reagent or organolithium reagent to an amide.

<i>N</i>-<i>tert</i>-Butylbenzenesulfinimidoyl chloride Chemical compound

N-tert-Butylbenzenesulfinimidoyl chloride is a useful oxidant for organic synthesis reactions. It is a good electrophile, and the sulfimide S=N bond can be attacked by nucleophiles, such as alkoxides, enolates, and amide ions. The nitrogen atom in the resulting intermediate is basic, and can abstract an α-hydrogen to create a new double bond.

<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 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.

References

  1. 1 2 Nahm, S.; Weinreb, S. M. (1981), "N-methoxy-n-methylamides as effective acylating agents", Tetrahedron Letters , 22 (39): 3815–3818, doi:10.1016/s0040-4039(01)91316-4
  2. 1 2 Paek, S.-M.; Seo, S.-Y.; Kim, S.-H.; Jung, J.-W.; Lee, Y.-S.; Jung, J.-K.; Suh, Y.-G. (2005), "Concise Syntheses of (+)-Macrosphelides A and B", Organic Letters , 7 (15): 3159–3162, doi:10.1021/ol0508429, PMID   16018610
  3. 1 2 Barbazanges, M.; Meyer, C.; Cossy, J. (2008), "Total Synthesis of Amphidinolide J", Organic Letters , 10 (20): 4489–4492, doi:10.1021/ol801708x, PMID   18811171
  4. 1 2 Shimizu, T.; Satoh, T.; Murakoshi, K.; Sodeoka, M. (2005), "Asymmetric Total Synthesis of (−)-Spirofungin A and (+)-Spirofungin B", Organic Letters , 7 (25): 5573–5576, doi:10.1021/ol052039k, PMID   16320994
  5. Qu, B.; Collum, D. B. (2006), "Mechanism of Acylation of Lithium Phenylacetylide with a Weinreb Amide", The Journal of Organic Chemistry , 71 (18): 7117–7119, doi:10.1021/jo061223w, PMID   16930080
  6. 1 2 3 Singh, J.; Satyamurthi, N.; Aidhen, I. S. (2000), "The Growing Synthetic Utility of Weinreb's Amide", Journal für praktische Chemie , 342: 340, doi:10.1002/(sici)1521-3897(200004)342:4<340::aid-prac340>3.0.co;2-1
  7. 1 2 3 Mentzel, M.; Hoffmann, H. M. R. (1997), "N-methoxy-N-methylamides (Weinreb amides) in modern organic synthesis", Journal für Praktische Chemie/Chemiker-Zeitung , 339: 517–524, doi:10.1002/prac.19973390194
  8. Martinelli, J. R.; Freckmann, D. M. M.; Buchwald, S. L. (2006), "Convenient Method for the Preparation of Weinreb Amides via Pd-Catalyzed Aminocarbonylation of Aryl Bromides at Atmospheric Pressure", Organic Letters , 8 (21): 4843–4846, doi:10.1021/ol061902t, PMID   17020317
  9. Graham, S. L.; Scholz, T. H. (1990), "A new mode of reactivity of N-methoxy-N-methylamides with strongly basic reagents", Tetrahedron Letters , 31 (44): 6269–6272, doi:10.1016/s0040-4039(00)97039-4
  10. Hisler, K.; Tripoli, R.; Murphy, J. A. (2006), "Reactions of Weinreb amides: formation of aldehydes by Wittig reactions", Tetrahedron Letters , 47 (35): 6293–6295, doi:10.1016/j.tetlet.2006.06.118
  11. Conrad, K.; Hsiao, Y.; Miller, R. (2005), "A practical one-pot process for α-amino aryl ketone synthesis", Tetrahedron Letters , 46 (49): 8587–8589, doi:10.1016/j.tetlet.2005.09.183
  12. Whipple, W. L.; Reich, H. J. (1991), "Use of N,N'-dimethoxy-N,N'-dimethylurea as a carbonyl dication equivalent in organometallic addition reactions. Synthesis of unsymmetrical ketones", The Journal of Organic Chemistry , 56 (8): 2911–2912, doi:10.1021/jo00008a057
  13. Sibi, M. P.; Sharma, R.; Paulson, K. L. (1992), "N,N′-Dimethoxy-N,N -Dimethylethanediamide: A Useful α-Oxo-N-Methoxy-N-Methylamide and 1,2-Diketone Synthon", Tetrahedron Letters , 33: 1941, doi:10.1016/0040-4039(92)88108-h
  14. Davies, S. G.; Goodwin, C. J.; Hepworth, D.; Roberts, P. M.; Thomson, J. E. (2010), "On the Origins of Diastereoselectivity in the Alkylation of Enolates Derived from N-1-(1'-Naphthyl)ethyl-O-tert-butylhydroxamates: Chiral Weinreb Amide Equivalents", The Journal of Organic Chemistry , 75 (4): 1214–1227, doi:10.1021/jo902499s, PMID   20095549