Julia olefination

Last updated

Contents

Julia olefination
Named after Marc Julia
Reaction type Coupling reaction
Identifiers
Organic Chemistry Portal julia-olefination
RSC ontology ID RXNO:0000117

The Julia olefination (also known as the Julia–Lythgoe olefination) is the chemical reaction used in organic chemistry of phenyl sulfones (1) with aldehydes (or ketones) to give alkenes (olefins)(3) after alcohol functionalization and reductive elimination using sodium amalgam or SmI2. The reaction is named after the French chemist Marc Julia.

Julia Olefination Revised Scheme Julia Olefination Revised Scheme.png
Julia Olefination Revised Scheme

The utility of this connective olefination reaction arises from its versatility, its wide functional group tolerance, and the mild reaction conditions under which the reaction proceeds.

All four steps can be carried out in a single reaction vessel, and use of R3X is optional. However, purification of the sulfone intermediate 2 leads to higher yield and purity. Most often R3 is acetyl or benzoyl, with acetic anhydride or benzoyl chloride used in the preparation of 2.

History

In 1973, Marc Julia and Jean-Marc Paris reported a novel olefin synthesis in which β-acyloxysulfones were reductively eliminated to the corresponding di-, tri-, or tetrasubstituted alkenes. Basil Lythgoe and Philip J. Kocienski explored the scope and limitation of the reaction, and today this olefination is formally known as the Julia-Lythgoe olefination. The reaction involves the addition of a sulfonyl-stabilized carbanion to a carbonyl compound, followed by elimination to form an alkene. In the initial versions of the reactions, the elimination was done under reductive conditions. More recently, a modified version that avoids this step was developed. The former version is sometimes referred to as the Julia-Lythgoe olefination, whereas the latter is called the Julia-Kocienski olefination. In the reductive variant, the adduct is usually acylated and then treated with a reducing agent, such as sodium amalgam or SmI2. Several reviews of these reactions have been published.

Reaction mechanism

The initial steps are straightforward. The phenyl sulfone anion (2) reacts with an aldehyde to form the alkoxide (3). The alkoxide is functionalized with R3-X to give the stable intermediate (4). The exact mechanism of the sodium amalgam reduction is unknown but has been shown to proceed through a vinylic radical species (5) . Protonation of the vinylic radical gives the desired product (6).

Julia olefination mechanism wiki Julia olefination mechanism wiki.png
Julia olefination mechanism wiki

The stereochemistry of the alkene (6) is independent of the stereochemistry of the sulfone intermediate 4. It is thought that the radical intermediates are able to equilibrate so that the more thermodynamically stable trans-olefin is produced most often. This transformation highly favors formation of the E-alkene.

Variations

Modified Julia olefination

General modified julia scheme General modified julia scheme.png
General modified julia scheme

The modified Julia olefination, also known as the one-pot Julia olefination is a modification of the classical Julia olefination. The replacement of the phenyl sulfones with heteroaryl sulfones greatly alters the reaction pathway. The most popular example is the benzothiazole sulfone. The reaction of the benzothiazole sulfone (1) with lithium diisopropylamide (LDA) gives a metallated benzothiazolyl sulfone, which reacts quickly with aldehydes (or ketones) to give an alkoxide intermediate (2). Unlike the phenyl sulfones, this alkoxide intermediate (2) is more reactive and will undergo a Smiles rearrangement to give the sulfinate salt (4). The sulfinate salt (4) will spontaneously eliminate sulfur dioxide and lithium benzothiazolone (5) producing the desired alkene (6).

The mechanism of the benzothiazole variation of the Julia olefination Benzothiazole-Julia Olefination Mechanism.png
The mechanism of the benzothiazole variation of the Julia olefination

Since the benzothiazole variation of the Julia olefination does not involve equilibrating intermediates, the stereochemical outcome is a result of the stereochemistry of the initial carbonyl addition. As a result, this reaction often generates a mixture of alkene stereoisomers.

Julia–Kocienski olefination

Julia–Kocienski olefination
Named after Marc Julia
Philip Joseph Kocienski
Reaction type Coupling reaction
Identifiers
Organic Chemistry Portal modified-julia-kocienski-olefination
RSC ontology ID RXNO:0000304
General julia kocienski scheme General julia kocienski scheme.png
General julia kocienski scheme

The Julia–Kocienski Olefination, a further refinement of the Modified Julia olefination, offers very good E-selectivity. In the Julia–Kocienski olefination the alkylating agent is a tetrazole. It proceeds with the same mechanism as the benzothiazole sulfone above. The high E-selectivity of the Julia–Kocienski olefination is the result of kinetically controlled diastereoselective addition of metalated 1-phenyl-1H-tetrazol-5-yl (PT) sulfones to nonconjugated aldehydes. This yields anti-β-alkoxysulfones which stereospecifically decompose to the E-alkenes. In one adaptation, with t-butyltetrazoylmethyl sulfone the reaction conditions are either sodium bis(trimethylsilyl)amide at −70 °C in tetrahydrofuran or caesium carbonate at +70 °C. This reaction is named after Philip J. Kocienski for his modification to the Julia olefination.

Julia-Kocienski olefination wiki Julia-Kocienski olefination wiki.png
Julia-Kocienski olefination wiki

Synthetic Applications

The Julia or modified Julia olefination reaction is a powerful and versatile synthetic transformation, widely utilized in the construction of complex natural products with excellent control of geometrical isomerism.

Pterostilbene

Pterostilbene is a stilbenoid chemically related to resveratrol. It belongs to the group of phytoalexins, agents produced by plants to fight infections. Pterostilbene is a naturally occurring dimethyl ether analog of resveratrol. It is believed that the compound also has anti-diabetic properties, but so far very little has been studied on this issue.

Compared to the Wittig, Wittig-Horner, Perkin, or transition-metal-catalyzed reactions to synthesize pterostilebene, the Julia olefination offers a simple, economical alternative method for preparation of pterostilbene.

Synthesis of pterostillbene through Julia Olefination Synthesis of pterostillbene through Julia Olefination.png
Synthesis of pterostillbene through Julia Olefination

Resveratrol

One adaptation of the Julia-Kocienski olefination gives the synthesis of the stilbenoid resveratrol, a natural compound found in common foods like grapes, wines and nuts. Resveratrol is a biologically important stilbenoid which has been suggested to have many health benefits. The Julia-Kocienski olefination serves as a powerful reaction in the synthesis of resveratrol analogues with 3,5-bis(trifluoromethyl)phenyl sulfones. The following schematic displays the general scheme for synthesizing resveratrol analogues, where R2 is an aryl group.

General Resveratrol Analogue Scheme General Resveratrol Analogue Scheme.png
General Resveratrol Analogue Scheme

(−)-Callystatin A

In the asymmetric total synthesis of (−)-callystatin A by Amos Smith, two separate Julia olefinations were used to append two E-alkene moieties. (−)-callystatin A is a member of the leptomycin family of antibiotics. The following schematic displays the Julia-Kocienski olefination used to achieve the precursor to the natural product, as indicated by use of the PT-sulfone.

Julia olefination for callystatin A Julia olefination for callystatin A.png
Julia olefination for callystatin A

See also

Related Research Articles

<span class="mw-page-title-main">Alkene</span> Hydrocarbon compound containing one or more C=C bonds

In organic chemistry, an alkene is a hydrocarbon containing a carbon–carbon double bond.

<span class="mw-page-title-main">Elias James Corey</span> American chemist (born 1928)

Elias James Corey is an American organic chemist. In 1990, he won the Nobel Prize in Chemistry "for his development of the theory and methodology of organic synthesis", specifically retrosynthetic analysis. Regarded by many as one of the greatest living chemists, he has developed numerous synthetic reagents, methodologies and total syntheses and has advanced the science of organic synthesis considerably.

The Wittig reaction or Wittig olefination is a chemical reaction of an aldehyde or ketone with a triphenyl phosphonium ylide called a Wittig reagent. Wittig reactions are most commonly used to convert aldehydes and ketones to alkenes. Most often, the Wittig reaction is used to introduce a methylene group using methylenetriphenylphosphorane (Ph3P=CH2). Using this reagent, even a sterically hindered ketone such as camphor can be converted to its methylene derivative.

<span class="mw-page-title-main">Peterson olefination</span>

The Peterson olefination is the chemical reaction of α-silyl carbanions with ketones to form a β-hydroxysilane (2) which eliminates to form alkenes (3).

The Barton–Kellogg reaction is a coupling reaction between a diazo compound and a thioketone, giving an alkene by way of an episulfide intermediate. The Barton–Kellogg reaction is also known as Barton–Kellogg olefination and Barton olefin synthesis.

<span class="mw-page-title-main">Horner–Wadsworth–Emmons reaction</span>

The Horner–Wadsworth–Emmons (HWE) reaction is a chemical reaction used in organic chemistry of stabilized phosphonate carbanions with aldehydes to produce predominantly E-alkenes.

<span class="mw-page-title-main">Prins reaction</span> Chemical reaction involving organic compounds

The Prins reaction is an organic reaction consisting of an electrophilic addition of an aldehyde or ketone to an alkene or alkyne followed by capture of a nucleophile or elimination of an H+ ion. The outcome of the reaction depends on reaction conditions. With water and a protic acid such as sulfuric acid as the reaction medium and formaldehyde the reaction product is a 1,3-diol (3). When water is absent, the cationic intermediate loses a proton to give an allylic alcohol (4). With an excess of formaldehyde and a low reaction temperature the reaction product is a dioxane (5). When water is replaced by acetic acid the corresponding esters are formed.

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

Tebbe's reagent is the organometallic compound with the formula (C5H5)2TiCH2ClAl(CH3)2. It is used in the methylenation of carbonyl compounds, that is it converts organic compounds containing the R2C=O group into the related R2C=CH2 derivative. It is a red solid that is pyrophoric in the air, and thus is typically handled with air-free techniques. It was originally synthesized by Fred Tebbe at DuPont Central Research.

<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 Corey–Winter olefin synthesis is a series of chemical reactions for converting 1,2-diols into olefins. It is named for the American chemist and Nobelist Elias James Corey and the American-Estonian chemist Roland Arthur Edwin Winter.

In organic chemistry, the Ei mechanism, also known as a thermal syn elimination or a pericyclic syn elimination, is a special type of elimination reaction in which two vicinal (adjacent) substituents on an alkane framework leave simultaneously via a cyclic transition state to form an alkene in a syn elimination. This type of elimination is unique because it is thermally activated and does not require additional reagents, unlike regular eliminations, which require an acid or base, or would in many cases involve charged intermediates. This reaction mechanism is often found in pyrolysis.

The Abramov reaction is the related conversions of trialkyl to α-hydroxy phosphonates by the addition to carbonyl compounds. In terms of mechanism, the reaction involves attack of the nucleophilic phosphorus atom on the carbonyl carbon. It was named after the Russian chemist Vasilii Semenovich Abramov (1904–1968) in 1957.

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.

Desulfonylation reactions are chemical reactions leading to the removal of a sulfonyl group from organic compounds. As the sulfonyl functional group is electron-withdrawing, methods for cleaving the sulfur–carbon bonds of sulfones are typically reductive in nature. Olefination or replacement with hydrogen may be accomplished using reductive desulfonylation methods.

Metal-catalyzed intermolecular carbenoid cyclopropanations are organic reactions that result in the formation of a cyclopropane ring from a metal carbenoid species and an alkene. In the Simmons–Smith reaction the metal involved is zinc.

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">Vinyl iodide functional group</span>

In organic chemistry, a vinyl iodide functional group is an alkene with one or more iodide substituents. Vinyl iodides are versatile molecules that serve as important building blocks and precursors in organic synthesis. They are commonly used in carbon-carbon forming reactions in transition-metal catalyzed cross-coupling reactions, such as Stille reaction, Heck reaction, Sonogashira coupling, and Suzuki coupling. Synthesis of well-defined geometry or complexity vinyl iodide is important in stereoselective synthesis of natural products and drugs.

William Clark Still is an American organic chemist. As a distinguished professor at Columbia University, Clark Still made significant contributions to the field of organic chemistry, particularly in the areas of natural product synthesis, reaction development, conformational analysis, macrocyclic stereocontrol, and computational chemistry. Still and coworkers also developed the purification technique known as flash column chromatography which is widely used for the purification of organic compounds.

Carbonyl olefin metathesis is a type of metathesis reaction that entails, formally, the redistribution of fragments of an alkene and a carbonyl by the scission and regeneration of carbon-carbon and carbon-oxygen double bonds respectively. It is a powerful method in organic synthesis using simple carbonyls and olefins and converting them into less accessible products with higher structural complexity.

References

    1. ^ Julia, M.; Paris, J.-M. Tetrahedron Lett. 1973, 14, 4833–4836. ( doi : 10.1016/S0040-4039(01)87348-2)
    2. ^ Kocienski, P. J.; Lythgoe, B.; Ruston, S. J. Chem. Soc., Perkin Trans. 1 1978, 829.
    3. ^ Keck, G. E.; Savin, K. A.; Weglarz, M. A. J. Org. Chem. 1995, 60, 3194–3204. ( doi : 10.1021/jo00115a041)
    4. ^ Kocienski, P. J. Phosphorus and Sulfur 1985, 24, 97–127. (Review)
    5. ^ Kelly, S. E. Comprehensive Organic Synthesis 1991, 1, 792–806. (Review) ( doi : 10.1016/B978-0-08-052349-1.00020-2)
    6. ^ Blakemore, P. R. J. Chem. Soc., Perkin Trans. 1 2002, 2563–2585. ( doi : 10.1039/b208078h)
    7. ^ Baudin, J. B.; Hareau, G.; Julia, S. A.; Ruel, O. Tetrahedron Lett. 1991, 32, 1175. ( doi : 10.1016/S0040-4039(00)92037-9)
    8. ^ Truce, W. E.; Kreider, E. M.; Brand, W. W. Org. React. 1970, 18, 99. (Review)
    9. ^ Paul R. Blakemore, William J. Cole, Philip J. Kocieński, Andrew Morley Synlett 1998, 26–28. ( doi : 10.1055/s-1998-1570)
    10. ^ Christophe Aïssa J. Org. Chem. 2006, 71, 360–63. ( doi : 10.1021/jo051693a)
    11. ^ Zajc, B., & Kumar, R. (2010). Synthesis of Fluoroolefins via Julia-Kocienski Olefination. Synthesis, 2010(11), 1822–1836.( doi : 10.1055/s-0029-1218789)
    12. ^ Langcake, P.; Pryce, R. J. (1977). "A new class of phytoalexins from grapevines". Experientia 33 (2): 151–2. ( doi : 10.1007/BF02124034) PMID   844529.
    13. ^ Moro, A. V.; Cardoso, F. S. P.; Correia, C. R. D. Heck arylation of styrenes with arenediazonium salts: Short, efficient, and stereoselective synthesis of resveratrol, DMU-212, and analogues. Tetrahedron Lett. 2008, 49(39), 5668–5671.
    14. ^ Prabhakar Peddikotla, Amar G. Chittiboyina, Ikhlas A. Khan, (2014) ChemInform Abstract: Synthesis of Pterostilbene by Julia Olefination. ChemInform 45, doi : 10.1002/chin.201408101.
    15. ^ Alonso DA, Fuensanta M, Nájera C, Varea M. J. Org. Chem. 2005; 70:6404–6416. PMID   16050703.
    16. ^ A. B. Smith, III and B. M. Brandt. Total Synthesis of (–)-Callystatin A. Org. Lett. 2001, 3, 1685–1688.
    17. ^ Robiette, R.; Pospíšil, J. On the Origin of E/Z Selectivity in the Modified Julia Olefination: Importance of the Elimination Step; Eur. J. Org. Chem. 2013, 836–840.