Big Chemical Encyclopedia

Chemical substances, components, reactions, process design ...

Articles Figures Tables About

Alkylation reactions of enolates

The alkylation reactions of enolate anions of both ketones and esters have been extensively utilized in synthesis. Both very stable enolates, such as those derived from (i-ketoesters, / -diketones, and malonate esters, as well as less stable enolates of monofunctional ketones, esters, nitriles, etc., are reactive. Many aspects of the relationships between reactivity, stereochemistry, and mechanism have been clarified. A starting point for the discussion of these reactions is the structure of the enolates. Because of the delocalized nature of enolates, an electrophile can attack either at oxygen or at carbon. [Pg.435]

Enolate geometry is of fundamental interest in order to be able to control the steric course of alkylation reactions of enolates. Determination of the configuration of an enolate is not trivial and has been carried out by first coupling it to the preferential transition state of the Claisen rearrangement2 5. [Pg.725]

The work of Myers et al. [6] illustrates the synthetic potential of the use of metal salts (instead of HMPA ) in alkylation reactions of enolates, employing easily accessible amide, enolates of the chiral auxiliary pseudoephed-rine. It is not surprising that the mechanism of chiral induction is not yet fully understood further investigations are necessary. Nonetheless, unanswered questions in enolate chemistry remain even for tailor-made, well-established auxiliaries, whose asymmetric induction can be explained convincingly by working models on monomer enolate structures, considering chelation control and steric factors. [Pg.20]

Cyclization of 47 discussed in Section 9.2.E showed an intramolecular alkylation reaction of enolates to give 49, but the formation of 50 illustrated an important competing reaction in the alkylation of enolates. An enolate anion is a bidentate nucleophile because it contains two nucleophilic centers, the carbanionic center and the oxygen, as illustrated by 90. Reaction of the O nucleophile in 90 with an alkyl halide leads to the vinyl ether 91. Reaction at the C nucleophile of 90 generates the usual alkylation product 92. Since a carb-anion is usually more nucleophilic than an alkoxy anion for common electrophiles, 92 tends to be the major product. There are several factors that can influence the relative proportion of these products, however. When... [Pg.735]

The catalytic enantioselective direct alkylation reaction of enolates is a less developed field. Early research from Evans group demonstrated that preformed titanium enolates derived from chiral Af-acyloxazolidinones reacted with orthoesters to provide the alkylated adducts with high levels of diastereo-control. In2005, the same group reported the enantioselective nickel-catalysed... [Pg.255]

Scheme 4.9 Evans auxiliaries 45-47 and examples for alkylation reactions of enolates 49 and 52. Scheme 4.9 Evans auxiliaries 45-47 and examples for alkylation reactions of enolates 49 and 52.
In contrast to the selectivities observed for the cyclooctanones, alkylation of the nine- or ten-membered ring ketones affords nearly 1 1 mixtures of cis and trans stereoisomers [3]. The alkylation reactions of enolates derived from medium-sized lactones, however, are notable. The series of methyl-substituted lactones from the nine-membered (27) through to the 1.3-membered (30) rings all show a strong preference for cis substitution (Equation 1) [3]. [Pg.7]

Interactive to use a web-based palette to predict products in halogenation and alkylation reactions of carbonyl enolates. [Pg.855]

Perhaps the single most important reaction of enolate ions is their alkylation by treatment with an alkyl halide or tosylate, thereby forming a new C-C bond and joining two smaller pieces into one larger molecule. Alkylation occurs when the nucleophilic enolate ion reacts with the electrophilic alkyl halide in an SN2 reaction and displaces the leaving group by backside attack. [Pg.855]

Figure 22.7 The biosynthesis of indolmycin from indolylpyruvate occurs through a pathway that includes an alkylation reaction of a short-lived enolate ion intermediate. Figure 22.7 The biosynthesis of indolmycin from indolylpyruvate occurs through a pathway that includes an alkylation reaction of a short-lived enolate ion intermediate.
Alpha hydrogen atoms of carbonyl compounds are weakly acidic and can be removed by strong bases, such as lithium diisopropylamide (LDA), to yield nucleophilic enolate ions. The most important reaction of enolate ions is their Sn2 alkylation with alkyl halides. The malonic ester synthesis converts an alkyl halide into a carboxylic acid with the addition of two carbon atoms. Similarly, the acetoacetic ester synthesis converts an alkyl halide into a methyl ketone. In addition, many carbonyl compounds, including ketones, esters, and nitriles, can be directly alkylated by treatment with LDA and an alkyl halide. [Pg.866]

The fundamental aspects of the structure and stability of carbanions were discussed in Chapter 6 of Part A. In the present chapter we relate the properties and reactivity of carbanions stabilized by carbonyl and other EWG substituents to their application as nucleophiles in synthesis. As discussed in Section 6.3 of Part A, there is a fundamental relationship between the stabilizing functional group and the acidity of the C-H groups, as illustrated by the pK data summarized in Table 6.7 in Part A. These pK data provide a basis for assessing the stability and reactivity of carbanions. The acidity of the reactant determines which bases can be used for generation of the anion. Another crucial factor is the distinction between kinetic or thermodynamic control of enolate formation by deprotonation (Part A, Section 6.3), which determines the enolate composition. Fundamental mechanisms of Sw2 alkylation reactions of carbanions are discussed in Section 6.5 of Part A. A review of this material may prove helpful. [Pg.2]

Six-Membered Ring (exo-Cyclic). The diastereoselective alkylation reactions of exo-cyclic enolates involving 1,2-asymmetric inductions are anti-inductions. In Scheme 2-2, there are two possible enolate chair conformations in which the two possible transition-state geometries lead to the major diaster-eomer 9e (where the substituent takes the equatorial orientation). However, for the case in which R = methyl and X = alkoxyl or alkyl, one would expect the... [Pg.74]

The diastereoselective alkylation reaction of endo-cyclic five-membered ring enolates exhibits good potential for both 1,3- and 1,2-asymmetric induction. In Scheme 2-6, the factor controlling the alkylation transition state is steric rather than stereoelectronic, leading to an auh-induction.13... [Pg.76]

TABLE 2-5. Diastereoselective Alkylation Reaction of the Lithium Enolates Derived from Imides 22 and 23... [Pg.85]

A further attempt has been made to develop a predictive model for chirality transfer achieved through alkylation reactions of ester enolates which feature chiral auxiliaries. " Hippurate esters (30) derived from (lI , 25 )-trani-2-(p-substituted phenyl)cyclohexanols were found, on reaction with benzyl bromide, to give (31) with predominantly the S configuration at the alkylation centre but with no correlation between the degree of stereoselectivity (20-98%) and the electron density on the aromatic ring. [Pg.357]

In light of these significant challenges, Evans and Leahy reexamined the rhodium-catalyzed allylic alkylation using copper(I) enolates, which should be softer and less basic nucleophiles [23]. The copper(I) enolates were expected to circumvent the problems typically associated with enolate nucleophiles in metal-allyl chemistry, namely ehmina-tion of the metal-aUyl intermediate and polyalkylation as well as poor regio- and stereocontrol. Hence, the transmetallation of the lithium enolate derived from acetophenone with a copper(I) hahde salt affords the requisite copper] I) enolate, which permits the efficient regio- and enantiospecific rhodium-catalyzed allylic alkylation reaction of a variety of unsymmetrical acychc alcohol derivatives (Tab. 10.3). [Pg.197]

This section deals with the alkylation reactions of such enolates. In the presence of strong bases, amides carrying at least one a-hydrogen 1 can be deprotonated to form enolate ions which, on subsequent alkylation, give alkylated amides. Further reaction, e g., hydrolysis or reduction, furnishes the corresponding acids or primary alcohols, respectively. The pKa values for deprotonation are typically around 35 (extrapolated value DMSO3 7) unless electron-withdrawing substituents are present in the a-position. Thus, deprotonation usually requires non-nucleophilic bases such as lithium diisopropylamide (extrapolated 8 pKa for the amine in DMSO is around 44) or sodium hexamethyldisilazanide. [Pg.791]

Rotation is hindered in the enolate. Thus, if the a-substituent R1 4= R2, the enolate can exist in two forms, the syn- and anti-forms (enolates 2 and 3, respectively, if R2 has higher priority than R1). Attack of an electrophile on either face of the enolates, 2 or 3, leads to a mixture of the alkylated amides, 4 and 5. If R1 and R2 and the A-substituents R3 and R4 are all achiral, the two alkylated amides will be mirror images and thus a racemate results. If, however, any of the R substituents are chiral, enolate 2 will give a certain ratio of alkylated amide 4/5, whereas enolate 3 will give a different, usually inverted, ratio. Thus, for the successful design of stereoselective alkylation reactions of chiral amide enolates it is of prime importance to control the formation of the enolate so that one of the possible syn- or anti-isomers is produced in large excess over the other,... [Pg.791]

In a similar manner to that described for bicyclic lactams (Section 1.1.1.3.3.4.1.5.I.). alkylation reactions of tricyclic lactams, which contain a fused benzene ring adjacent to the carbon undergoing alkylation, have been exploited14. The first alkylation of the benzo-annulated bicyclic lactam 1 gives a mixture of diastereomers, which is then further alkylated. In the second alkylation step, the counterion on the alkoxide, which is formed prior to enolate formation, proved to be crucial for the diastereoselectivity of the subsequent alkylation reaction. The best diastcrcoselectivity was obtained when either dichlorobis(ij5-cyclopentadienyl)zirconium or triisopropoxytitanium chloride was added to the preformed alkoxide, followed by enolization and alkylation. Using this method the second alkylation step gives a satisfactory diastereoselectivity. Hydride reduction of the purified major diastereomer 2, followed by acid treatment of the product, furnishes chiral naphthalenones 414. [Pg.881]

In the alkylation reactions of the chiral 3-acyl-2-oxazolidinones, deprotonation to the lithium or sodium enolate is by treatment with lithium diisopropylamide or lithium or sodium hexamethyldisilazanide in tetrahydrofuran at low temperature (usually — 78 °C). The haloalka-ne, usually in excess, is then added to the enolate solution at low temperature (usually — 78 °C) for the sodium enolates and at higher temperatures (between —78 and 0CC) for the lithium enolates. When small, less sterically demanding halides, such as iodomethane, are used the sodium enolate is usually preferred 2 24 and in these cases up to five equivalents2,6- 24,26,27 of the halide are necessary in order to obtain good yields of the alkylation products. Conventional extractive workup provides the crude product as a diastereomeric mixture (d.r. usually > 90 10) which is relatively easy to separate by silica gel chromatography and/or by recrystallization (for crystalline products). Thus, it is possible to obtain one diastereomer in very high diastereomeric purity. [Pg.888]

Deprotonation of either the (4S.5R)- or (4/ ,5S)-enantiomer of 3-acyl-1,5-dimelhyl-4-phenyl-2-imidazolidinones 4 by lithium cyclohexylisopropylamide (LICA)1 or diisopropylamide2 furnishes chiral, supposedly chelated enolates, very similar to those enolates obtained from 2-oxazolidi-nones (see Section 1.1.1.3.3.4.2.1.). With LICA the. yyn-enolate is formed exclusively, as shown by O-silylation of the enolate with /ert-butylchlorodimethylsilane1. Attack of an electrophile, such as a haloalkane, from the less hindered side furnishes products (usually crystalline) with a moderate to high degree of diastereoselectivity (see Tabled)1 2. The diastereoselectivities observed in comparable alkylation reactions of the 3-acyl-4-cyclohexyl-l,5-dimethylimidazo-lidinone 3b are superior to those obtained with the 4-phenyl derivative 3a2,7. Thus, as also observed in similar alkylations with oxazolidinones10 (see Section 1.1.1.3.3.4.2.1), a phenyl substituent on the chiral auxiliary seems to be relatively inefficient as a steric control element. [Pg.901]

Reaction of enolate 5 with methyl trifluoromethanesulfonate diastereoselectively provides 6 as may be expected, alkylation occurs from the less hindered face of the enolate opposite the phosphane ligand87. [Pg.959]

Reviews on stoichiometric asymmetric syntheses M. M. Midland, Reductions with Chiral Boron Reagents, in J. D. Morrison, ed., Asymmetric Synthesis, Vol. 2, Chap. 2, Academic Press, New York, 1983 E. R. Grandbois, S. I. Howard, and J. D. Morrison, Reductions with Chiral Modifications of Lithium Aluminum Hydride, in J. D. Morrison, ed.. Asymmetric Synthesis, Vol. 2, Chap. 3, Academic Press, New York, 1983 Y. Inouye, J. Oda, and N. Baba, Reductions with Chiral Dihydropyridine Reagents, in J. D. Morrison, ed., Asymmetric Synthesis, Vol. 2, Chap. 4, Academic Press, New York, 1983 T. Oishi and T. Nakata, Acc. Chem. Res., 17, 338 (1984) G. Solladie, Addition of Chiral Nucleophiles to Aldehydes and Ketones, in J. D. Morrison, ed., Asymmetric Synthesis, Vol. 2, Chap. 6, Academic Press, New York, 1983 D. A. Evans, Stereoselective Alkylation Reactions of Chiral Metal Enolates, in J. D. Morrison, ed., Asymmetric Synthesis, Vol. 3, Chap. 1, Academic Press, New York, 1984. C. H. Heathcock, The Aldol Addition Reaction, in J. D. Morrison, ed., Asymmetric Synthesis, Vol. 3, Chap. 2, Academic Press, New York, 1984 K. A. Lutomski and A. I. Meyers, Asymmetric Synthesis via Chiral Oxazolines, in J. D. Morrison, ed., Asymmetric Synthesis, Vol. 3, Chap. [Pg.249]

Several important reactions of arenols involve aromatic substitution of arenolate ions with carbon electrophiles. In a sense, these reactions are alkylation and acylation reactions as discussed for arenes (Sections 22-4E and 22-4F). In another sense, they are alkylation and acylation reactions of enolate anions and therefore could give rise to products by C- and O-alkyla-tion, or C- and O-acylation (Section 17-4). Thus ... [Pg.1297]

D. A. Evans in Asymmetric Synthesis, Ed. J. D. Morrison, Academic Press, New York (1984), Vol 3, Chpt 1 (stereoselective alkylation reactions of chiral metal enolates)... [Pg.1477]

The nitrogen lone pair has been shown by both experiment and theoretical calculation to bias facially the alkylation reactions of nitrogen-containing pseudo-planar enolates (40) derived from pyrrolidinone (39). The preferred approach anti to the lone pair has been attributed to a heretofore unappreciated electronic effect.46... [Pg.336]

Tetrasubstituted pyrroles could be obtained by skeletal rearrangement of 1,3-oxazolidines, a reaction that is substantially accelerated by microwave irradiation. Dielectric heating of a 1,3-oxazolidine 7, absorbed on silica gel (1 g silica gel/mmol) for 5 min in a household MW oven (900 W power) cleanly afforded the 1,2,3,4-tetrasubstituted pyrrole 8 in 78% yield, thus reducing the reaction time from hours to minutes (Scheme 5) [24], 1,3-Oxazolidines are accessible in one-pot, two-step, solvent-free domino processes (see also Sect. 2.6). The first domino process, a multi-component reaction (MCR) between 2 equivalents of alkyl propiolate and 1 equivalent of aldehyde furnished enol ethers 9 (Scheme 5). Subsequent microwave-accelerated solvent-free reactions of enol ethers 9 with primary amines on silica support afforded intermediate 1,3-oxazolidines, which in situ rearranged to the tetrasubstituted pyrroles (2nd domino process). Performed in a one-pot format, these... [Pg.56]


See other pages where Alkylation reactions of enolates is mentioned: [Pg.579]    [Pg.4]    [Pg.3]    [Pg.4]    [Pg.70]    [Pg.72]    [Pg.76]    [Pg.579]    [Pg.4]    [Pg.3]    [Pg.4]    [Pg.70]    [Pg.72]    [Pg.76]    [Pg.36]    [Pg.172]    [Pg.82]    [Pg.5]    [Pg.352]    [Pg.19]    [Pg.970]    [Pg.795]    [Pg.805]    [Pg.830]    [Pg.961]    [Pg.110]   
See also in sourсe #XX -- [ Pg.899 , Pg.900 ]

See also in sourсe #XX -- [ Pg.338 , Pg.339 , Pg.340 , Pg.341 ]

See also in sourсe #XX -- [ Pg.897 , Pg.898 ]




SEARCH



Alkylation of enolates

Alkylation of enols

Alkylations of enolates

Enol alkyl

Enolate alkylation

Enolate alkylation reaction

Enolates alkylation

Enols alkylation

Mannich Reaction Alkylation of Enols by Iminium Ions

Of alkylation reactions

Reaction of stabilized carbanions (enolates) with alkyl halides (enolate alkylation)

Reactions of Enolates

© 2024 chempedia.info