Big Chemical Encyclopedia

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

Articles Figures Tables About

Protonated chiral catalysts

Protonated chiral catalysts versatile tools for asymmetric synthesis (C. Bolm, 2005) [Ic]. [Pg.5]

The chiral catalyst 142 achieves selectivities through a double effect of intramolecular hydrogen binding interaction and attractive tt-tt donor-acceptor interactions in the transition state by a hydroxy aromatic group [88]. The exceptional results of some Diels-Alder reactions of cyclopentadiene with substituted acroleins catalyzed by (R)-142 are reported in Table 4.21. High enantio- and exo selectivity were always obtained. The coordination of a proton to the 2-hydroxyphenyl group with an oxygen of the adjacent B-0 bond in the nonhelical transition state should play an important role both in the exo-endo approach and in the si-re face differentiation of dienophile. [Pg.185]

The function of the molecular sieves in this case is believed to be as a base that sequesters the protons, which otherwise would promote a variety of side reactions. With chiral catalysts, the carbonyl ene reaction becomes enantioselective. Among the successful catalysts are diisopropoxyTi(IV)BINOL and copper-BOX complexes. [Pg.874]

LA represents Lewis acid in the catalyst, and M represents Bren sled base. In Scheme 8-49, Bronsted base functionality in the hetero-bimetalic chiral catalyst I can deprotonate a ketone to produce the corresponding enolate II, while at the same time the Lewis acid functionality activates an aldehyde to give intermediate III. Intramolecular aldol reaction then proceeds in a chelation-controlled manner to give //-keto metal alkoxide IV. Proton exchange between the metal alkoxide moiety and an aromatic hydroxy proton or an a-proton of a ketone leads to the production of an optically active aldol product and the regeneration of the catalyst I, thus finishing the catalytic cycle. [Pg.490]

Collins and co-workers have performed studies in the area of catalytic enantioselective Diels—Alder reactions, in which ansa-metallocenes (107, Eq. 6.17) were utilized as chiral catalysts [100], The cycloadditions were typically efficient (-90% yield), but proceeded with modest stereoselectivities (26—52% ee). The group IV metal catalyst used in the asymmetric Diels—Alder reaction was the cationic zirconocene complex (ebthi)Zr(OtBu)-THF (106, Eq. 6.17). Treatment of the dimethylzirconocene [101] 106 with one equivalent of t-butanol, followed by protonation with one equivalent of HEt3N -BPh4, resulted in the formation of the requisite chiral cationic complex (107),... [Pg.212]

Later, Corey reported that the bulky superacid triflylimide (Tf2NH) protonates chiral oxazaborolidines to form superactive, stable, chiral acids 11a and 11b, which are highly effective catalysts for a wide variety of enantioselective Diels-Alder reactions that were beyond the reach of synthetic chemists (Scheme... [Pg.364]

Bifunctional catalysts have proven to be very powerful in asymmetric organic transformations [3], It is proposed that these chiral catalysts possess both Brpnsted base and acid character allowing for activation of both electrophile and nucleophile for enantioselective carbon-carbon bond formation [89], Pioneers Jacobsen, Takemoto, Johnston, Li, Wang and Tsogoeva have illustrated the synthetic utility of the bifunctional catalysts in various organic transformations with a class of cyclohexane-diamine derived catalysts (Fig. 6). In general, these catalysts contain a Brpnsted basic tertiary nitrogen, which activates the substrate for asymmetric catalysis, in conjunction with a Brpnsted acid moiety, such as urea or pyridinium proton. [Pg.172]

The three acetylpyridines have been reduced in the presence of catalytic concentrations of different alkaloids in attempts to induce optical activity in the products 412 The reduction of 3-acetylpyridine gave optically inactive alcohols under all conditions employed, whereas optically active pyridyl-ethanols are produced from 2- and 4-acetylpyridine at 0°C, in a 1 1 aqueous-ethanolic acetate buffer with strychnine (5 x 10-4 M) as chiral catalyst. Under these conditions protonated, adsorbed strychnine is probably acting as a chiral acid. The pinacols obtained as side products were all optically inactive. [Pg.332]

Buffer catalysis has been applied to induce chiral induction by enantio-selective protonation remarkable enantiomeric excess was achieved in the photodeconjugation of a,/3-unsaturated ketones and esters by using chiral catalysts for the ketonization of photoenols in aprotic solvents.29... [Pg.340]

Several reports deal with aqueous media. Acid-base catalysis by pure water has been explored, using DFT, for the model aldol reaction of acetone and acetaldehyde.125 A Hammett correlation of nornicotine analogues (28) - a series of meta- and para-substituted 2-arylpyrrolidines - as catalysts of an aqueous aldol reaction shows p = 1.14.126 Also, direct aldol reactions have been carried out in water enantioselectively, using protonated chiral prolinamide organocatalysts.127... [Pg.14]

According to another NMR study, the mechanism of bifunctional activation in the asymmetric aza-Morita-Baylis-Hillman reaction (Scheme 7) involves rate-limiting proton transfer (116) in the absence of added protic species155 (in consonance with the report summarized in Scheme 5144), but exhibits no autocatalysis. Addition of Brpnsted acids led to substantial rate enhancements through acceleration of the elimination step. Furthermore, it was found that phosphine catalysts, either alone or in combination with protic additives, can cause racemization of the reaction product by proton exchange at the stereogenic centre. This behaviour indicates that the spatial arrangement of a bifunctional chiral catalyst for the asymmetric aza-Morita-Baylis-Hillman reaction is crucial not only for the stereodifferentiation within the catalytic cycle but also for the prevention of subsequent racemization.155... [Pg.317]

The general reaction mechanism of the Michael reaction is given below (Scheme 4). First, deprotonation of the Michael donor occurs to form a reactive nucleophile (A, C). This adds enantioselectively to the electron-deficient olefin under the action of the chiral catalyst. In the final step, proton transfer to the developed enolate (B, D) occurs from either a Michael donor or the conjugate acid of a catalyst or a base, affording the desired Michael adduct. It is noteworthy that the large difference of stability between the two enolate anions (A/B, C/D) is the driving force for the completion of the catalytic cycle. [Pg.350]

The chemistry of asymmetric protonation of enols or enolates has further developed since the original review in Comprehensive Asymmetric Catalysis [1], Numbers of literature reports of new chiral proton sources have emerged and several reviews [2-6] cover the topics up to early 2001. This chapter concentrates on new examples of catalytic enantioselective protonation of prochiral metal enolates (Scheme 1). Compounds 1-41 [7-45] shown in Fig. 1 are the chiral proton sources or chiral catalysts reported since 1998 which have been employed for the asymmetric protonation of metal enolates. Some of these have been successfully utilized in the catalytic version. [Pg.141]

Muzart and coworkers have reported a new catalytic enantioselective protonation of prochiral enolic species generated by palladium-induced cleavage of p-ketoesters or enol carbonates of a-alkylated 1-indanones and 1-tetralones [21 ]. Among the various (S)-p-aminocycloalkanols examined, 17 and 18 were effective chiral catalysts for the asymmetric reaction and (J )-enriched a-alkylated 1-indanones and 1-tetralones were obtained with up to 72% ee. In some cases, the reaction temperature affected the ee. [Pg.145]

Figure 9.15 Catalytic cycle for asymmetric nitroaldol condensation reaction with 9.12 as the chiral catalyst. The La-O bond marked by an arrow opens up due to protonation of the O atom by nitromethane. Figure 9.15 Catalytic cycle for asymmetric nitroaldol condensation reaction with 9.12 as the chiral catalyst. The La-O bond marked by an arrow opens up due to protonation of the O atom by nitromethane.
Despite the obvious potential of cinchona alkaloids as bifunctional chiral catalysts of the nucleophilic addition/enantioselective protonation on prochiral ketenes, no further contribution has appeared to date and only a few papers described this asymmetric reaction with other catalysts [13], When the reaction is carried out with soft nucleophiles, the catalyst, often a chiral tertiary amine, adding first on ketene, is covalently linked to the enolate during the protonation. Thus, we can expect an optimal control of the stereochemical outcome of the protonation. This seems perfectly well suited for cinchona analogues and we can therefore anticipate successful applications of these compounds for this reaction in the near future. [Pg.177]

Leitner and coworkers [27] found that triphenylphosphine either alone or in combination with protonic additives could cause racemization of the aza MBH product by proton exchange at the stereogenic center, but the chiral catalyst 19a developed by Shi s group did not induce any racemization on a similar time scale. [Pg.412]

Muzart and coworkers have succeeded in a catalytic asymmetric protonation of enol compounds generated by palladium-induced cleavage of 3-ketoesters or enol carbonates under nearly neutral conditions [47,48]. Among the various optically active amino alcohols tested, (-i-)-e do-2-hydroxy-endo-3-aminoborn-ane (25) was effective as a chiral catalyst for the enantioselective reaction. Treatment of the P-ketoester of 2-methyl-1-indanone 58 with a catalytic amount of the amino alcohol 25 (0.3 equiv) and 5% Pd on charcoal (0.025 equiv) under bubbling of hydrogen at 21 °C gave the (P)-enriched product 59 with 60% ee... [Pg.1229]

Nakai and a coworker achieved a conceptually different protonation of silyl enol ethers using a chiral cationic palladium complex 40 developed by Shibasaki and his colleagues [61] as a chiral catalyst and water as an achiral proton source [62]. This reaction was hypothesized to progress via a chiral palladium enolate which was diastereoselectively protonated by water to provide the optically active ketone and the chiral Pd catalyst regenerated. A small amount of diisopropylamine was indispensable to accomplish a high level of asymmetric induction and the best enantioselectivity (79% ee) was observed for trimethylsilyl enol ether of 2-methyl-l-tetralone 52 (Scheme 11). [Pg.1230]

In 2002, Huang and Rawal found that the hetero Diels-Alder reaction of aminosiloxydienes with aldehydes was accelerated in alcoholic solvents [65], They subsequently elucidated that TADDOL (19) is an efficient chiral catalyst for the hetero-Diels-Alder reaction (Figure 10.17, Equation 10.33) [66]. The internal hydrogen bond in TADDOL observed in its crystal structure is expected to render the hydroxy proton more acidic, hence enabling it to participate better in intermolecular hydrogen bonding with the carbonyl group of the dienophile [67]. The Mukaiyama aldol reaction was also reported [68]. [Pg.323]


See other pages where Protonated chiral catalysts is mentioned: [Pg.312]    [Pg.312]    [Pg.402]    [Pg.575]    [Pg.153]    [Pg.134]    [Pg.191]    [Pg.216]    [Pg.229]    [Pg.290]    [Pg.146]    [Pg.140]    [Pg.382]    [Pg.709]    [Pg.1106]    [Pg.202]    [Pg.103]    [Pg.775]    [Pg.1225]    [Pg.156]   
See also in sourсe #XX -- [ Pg.5 ]




SEARCH



Catalysts protonation

Chiral catalysts

Chiral proton catalysts

Chiral protonation

Chiral protons

Hydrogen chiral proton” catalysts

© 2024 chempedia.info