Big Chemical Encyclopedia

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

Articles Figures Tables About

Alkenes chiral auxiliaries

Efficient acetalization of alkenes bearing various EWG with an optically active 1.3-diol 72 proceeds smoothly utilizing PdCN, CuCI. and O2 in DME to give the 1,3-dioxane 73[113], Methacrylamide bearing 4-t-butyloxazolidin-2-one 74 as a chiral auxiliary reacts with MeOH in the presence of PdCE catalyst... [Pg.31]

The most valuable characteristic of the Patemo-Buchi reaction is the ability to set multiple stereocenters in one reaction and the development of diastereocontrolled reactions has been a major theme of research concerning this reaction. Stereocontrol can be envisioned to spring from either the carbonyl or the alkene and be controlled by either the substrate directly or by a chiral auxiliary. Little success has been achieved in substrate-induced selection by the carbonyl the most successful results were produced by... [Pg.46]

To control the stereochemistry of 1,3-dipolar cycloaddidon reacdons, chiral auxiliaries are introduced into either the dipole-part or dipolarophile A recent monograph covers this topic extensively ° therefore, only typical examples are presented here. Alkenes employed in asymmetric 1,3-cycloaddidon can be divided into three main groups (1) chiral allyhc alcohols, f2 chiral amines, and Hi chiral vinyl sulfoxides or vinylphosphine oxides. [Pg.251]

The syntheses in Schemes 13.45 and 13.46 illustrate the use of oxazolidinone chiral auxiliaries in enantioselective synthesis. Step A in Scheme 13.45 established the configuration at the carbon that becomes C(4) in the product. This is an enolate alkylation in which the steric effect of the oxazolidinone chiral auxiliary directs the approach of the alkylating group. Step C also used the oxazolidinone structure. In this case, the enol borinate is formed and condensed with an aldehyde intermediate. This stereoselective aldol addition established the configuration at C(2) and C(3). The configuration at the final stereocenter at C(6) was established by the hydroboration in Step D. The selectivity for the desired stereoisomer was 85 15. Stereoselectivity in the same sense has been observed for a number of other 2-methylalkenes in which the remainder of the alkene constitutes a relatively bulky group.28 A TS such as 45-A can rationalize this result. [Pg.1205]

Asymmetric 1,3-dipolar cycloaddition of cyclic nitrones to crotonic acid derivatives bearing chiral auxiliaries in the presence of zinc iodide gives bicyclic isoxazolidines with high stereoselectivity (Eq. 8.51). The products are good precursors of (3-amino acids such as (+)sedridine.73 Many papers concerning 1,3-dipolar cycloaddition of nitrones to chiral alkenes have been reported, and they are well documented (see Ref. 63). [Pg.252]

Cycloaddition of nitrile oxides to alkenes with various chiral auxiliaries are summarized in Table 8.1, which shows chiral alkenes and differential excess (de). [Pg.266]

However, most asymmetric 1,3-dipolar cycloaddition reactions of nitrile oxides with alkenes are carried out without Lewis acids as catalysts using either chiral alkenes or chiral auxiliary compounds (with achiral alkenes). Diverse chiral alkenes are in use, such as camphor-derived chiral N-acryloylhydrazide (195), C2-symmetric l,3-diacryloyl-2,2-dimethyl-4,5-diphenylimidazolidine, chiral 3-acryloyl-2,2-dimethyl-4-phenyloxazolidine (196, 197), sugar-based ethenyl ethers (198), acrylic esters (199, 200), C-bonded vinyl-substituted sugar (201), chirally modified vinylboronic ester derived from D-( + )-mannitol (202), (l/ )-menthyl vinyl ether (203), chiral derivatives of vinylacetic acid (204), ( )-l-ethoxy-3-fluoroalkyl-3-hydroxy-4-(4-methylphenylsulfinyl)but-1 -enes (205), enantiopure Y-oxygenated-a,P-unsaturated phenyl sulfones (206), chiral (a-oxyallyl)silanes (207), and (S )-but-3-ene-1,2-diol derivatives (208). As a chiral auxiliary, diisopropyl (i ,i )-tartrate (209, 210) has been very popular. [Pg.25]

Since Sharpless discovery of asymmetric dihydroxylation reactions of al-kenes mediated by osmium tetroxide-cinchona alkaloid complexes, continuous efforts have been made to improve the reaction. It has been accepted that the tighter binding of the ligand with osmium tetroxide will result in better stability for the complex and improved ee in the products, and a number of chiral auxiliaries have been examined in this effort. Table 4 11 (below) lists the chiral auxiliaries thus far used in asymmetric dihydroxylation of alkenes. In most cases, diamine auxiliaries provide moderate to good results (up to 90% ee). [Pg.223]

In summary, the reaction of osmium tetroxide with alkenes is a reliable and selective transformation. Chiral diamines and cinchona alkakoid are most frequently used as chiral auxiliaries. Complexes derived from osmium tetroxide with diamines do not undergo catalytic turnover, whereas dihydroquinidine and dihydroquinine derivatives have been found to be very effective catalysts for the oxidation of a variety of alkenes. OsC>4 can be used catalytically in the presence of a secondary oxygen donor (e.g., H202, TBHP, A -methylmorpholine-/V-oxide, sodium periodate, 02, sodium hypochlorite, potassium ferricyanide). Furthermore, a remarkable rate enhancement occurs with the addition of a nucleophilic ligand such as pyridine or a tertiary amine. Table 4-11 lists the preferred chiral ligands for the dihydroxylation of a variety of olefins.61 Table 4-12 lists the recommended ligands for each class of olefins. [Pg.224]

Dipolar addition is closely related to the Diels-Alder reaction, but allows the formation of five-membered adducts, including cyclopentane derivatives. Like Diels-Alder reactions, 1,3-dipolar cycloaddition involves [4+2] concerted reaction of a 1,3-dipolar species (the An component and a dipolar In component). Very often, condensation of chiral acrylates with nitrile oxides or nitrones gives only modest diastereoselectivity.82 1,3-Dipolar cycloaddition between nitrones and alkenes is most useful and convenient for the preparation of iso-xazolidine derivatives, which can then be readily converted to 1,3-amino alcohol equivalents under mild conditions.83 The low selectivity of the 1,3-dipolar reaction can be overcome to some extent by introducing a chiral auxiliary to the substrate. As shown in Scheme 5-51, the reaction of 169 with acryloyl chloride connects the chiral sultam to the acrylic acid substrate, and subsequent cycloaddition yields product 170 with a diastereoselectivity of 90 10.84... [Pg.308]

Chiral active pharmaceutical ingredients, 18 725-726. See also Enantio- entries Chiral additives, 6 75—79 Chiral alcohols, synthesis of, 13 667-668 P-Chiral alcohols, synthesis of, 13 669 Chiral alkanes, synthesis of, 13 668-669 Chiral alkenes, synthesis of, 13 668—669 Chiral alkoxides, 26 929 Chiral alkynes, synthesis of, 13 668-669 Chiral ammonium ions, enantiomer recognition properties for, 16 790 Chiral ansa-metallocenes, 16 90 Chiral auxiliaries, in oxazolidinone formation, 17 738—739... [Pg.173]

Perlmutter used an oxymercuration/demercuration of a y-hydroxy alkene as the key transformation in an enantioselective synthesis of the C(8 ) epimeric smaller fragment of lb (and many more pamamycin homologs cf. Fig. 1) [36]. Preparation of substrate 164 for the crucial cyclization event commenced with silylation and reduction of hydroxy ester 158 (85-89% ee) [37] to give aldehyde 159, which was converted to alkenal 162 by (Z)-selective olefination with ylide 160 (dr=89 l 1) and another diisobutylaluminum hydride reduction (Scheme 22). An Oppolzer aldol reaction with boron enolate 163 then provided 164 as the major product. Upon successive treatment of 164 with mercury(II) acetate and sodium chloride, organomercurial compound 165 and a second minor diastereomer (dr=6 l) were formed, which could be easily separated. Reductive demercuration, hydrolytic cleavage of the chiral auxiliary, methyl ester formation, and desilylation eventually led to 166, the C(8 ) epimer of the... [Pg.233]

The inter- or intramolecular cyclopropanation of achiral alkenes with enantiome-rically pure diazoacetic esters [1016,1363,1364] or amides [1365,1366] does not usually proceed with high diastereoselectivity. A chiral auxiliary which occasionally gives good results is pantolactone (3-hydroxy-4,4-dimethyltetrahydro-2-furanone) [1016,1367,1368]. [Pg.219]

Using (-)-lOO [46] as a chiral auxiliary tethered to the enolether, one face of the alkene can be specifically blocked by a n-n interaction of the phenyl rest for the [2 r5+2 r ] cycloaddition with a ketene [47], resulting in the highly diastereoselective formation of the cyclobutanone 102 (Scheme 15). The observed regio- and stereoselectivity is in accord with the stereochemical predictions made on the basis of the Woodward-Hoffmann... [Pg.57]

In a search for more effective approaches to the problem of stereoselective dioxygenation, alternative methods have been developed. While the inclusion of alkenes and O2 within a chiral cyclodextrin cavity furnished hydroperoxides with modest ee values, neighboring stereocenters and chiral auxiliaries could induce highly stereoselective dioxygenation. In 1987 Kropf and Reichwaldt, and three years later Adam and coworkers reported on the photooxygenation of phenyl-substituted alkenes 36 producing allyUc hydroperoxides 37 and 38 with high diastereoselectivity (dr 80/20). In the best example... [Pg.342]

Another approach that relies on asymmetric induction from the alkene part, uses chiral auxiliaries of various types, thereby leading to enantiomerically enriched or pure isoxazoline products. The complexity of some of these auxiliaries is high, and more economical solutions are desirable since the competition is the resolution of racemic cycloadducts with an overall efficiency up to 50% yield. With chiral nitrile oxides, the situation is much less satisfactory since asymmetric induction of the 1,4-type (with 1-alkenes) is minimal, and hardly better with a 1,3-relationship of inducing-forming stereocenters, when 1,2-disubstituted alkenes are employed (Scheme 6.22). Upon separation of the two diastereomers, however, another entry to pure optically active isoxazolines is available. [Pg.386]

The use of chiral auxiliaries to induce (or even control) diastereoselectivity in the cycloaddition of nitrile oxides with achiral alkenes to give 5-substituted isoxazolines has been investigated by a number of groups. With chiral acrylates, this led mostly to low or modest diastereoselectivity, which was explained in terms of the conformational flexibility of the vinyl-CO linkage of the ester (Scheme 6.33) (179). In cycloadditions to chiral acrylates (or acrylamides), both the direction of the facial attack of the dipole as well as the conformational preference of the rotamers need to be controlled in order to achieve high diastereoselection. Although the attack from one sector of space may well be directed or hindered by the chiral auxiliary, a low diastereomer ratio would result due to competing attack to the respective 7i-faces of both the s-cis and s-trans rotamers of the acrylate or amide. [Pg.393]

The 1,3-dipolar cycloadditions of 1,3-dipoles with chiral alkenes has been extensively reviewed and thus only selected examples will be highlighted here. We have chosen to divide this section on the basis of the different types of alkenes rather than on the basis of the type of 1,3-dipole. For 1,3-dipolar cycloadditions, as well as for other reactions, it is important that the chiral center intended to control the stereoselectivity of the reaction is located as close as possible to the functional group of the molecule at which the reaction takes place. Hence, alkenes bearing the chiral center vicinal to the double bond are most frequently apphed in asymmetric 1,3-dipolar cycloadditions. Examples of the application of alkenes with the chiral center localized two or more bonds apart from the alkene will also be mentioned. Application of chiral auxiliaries for alkenes is very common and will be described separately in Section 12.3. [Pg.835]

D. Stereoselective Cyclopropanation of Alkenes using Chiral Auxiliaries. . . 266... [Pg.237]


See other pages where Alkenes chiral auxiliaries is mentioned: [Pg.47]    [Pg.252]    [Pg.80]    [Pg.58]    [Pg.22]    [Pg.242]    [Pg.269]    [Pg.268]    [Pg.289]    [Pg.894]    [Pg.25]    [Pg.272]    [Pg.161]    [Pg.255]    [Pg.117]    [Pg.472]    [Pg.241]    [Pg.41]    [Pg.2]    [Pg.205]    [Pg.234]    [Pg.18]    [Pg.1489]    [Pg.395]   
See also in sourсe #XX -- [ Pg.266 , Pg.267 , Pg.268 , Pg.269 , Pg.270 , Pg.271 , Pg.272 ]




SEARCH



Alkenes chiral

Chiral auxiliaries stereoselective alkene cyclopropanation

Chiral auxiliaries, diastereoselectivity, asymmetric alkenes

Chirality alkenes

Chirality auxiliaries

© 2024 chempedia.info