Big Chemical Encyclopedia

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

Articles Figures Tables About

Diazo cyclopropanation with

The same difference in regioselectivity holds for cyclopropanation with ethyl diazoacetate 25 K It is assumed that Cu(OTf)2 or Cu(BF4)2 are reduced to the Cu(I) salts by the diazo compound the ability of CuOTf to form stable complexes with olefins may then explain why, with these catalysts, cyclopropanation is governed by the steric environment around a double bond rather than by its electron-richness. [Pg.80]

It has been pointed out earlier that the anti/syn ratio of ethyl bicyclo[4.1,0]heptane-7-carboxylate, which arises from cyclohexene and ethyl diazoacetate, in the presence of Cul P(OMe)3 depends on the concentration of the catalyst57). Doyle reported, however, that for most combinations of alkene and catalyst (see Tables 2 and 7) neither concentration of the catalyst (G.5-4.0 mol- %) nor the rate of addition of the diazo ester nor the molar ratio of olefin to diazo ester affected the stereoselectivity. Thus, cyclopropanation of cyclohexene in the presence of copper catalysts seems to be a particular case, and it has been stated that the most appreciable variations of the anti/syn ratio occur in the presence of air, when allylic oxidation of cyclohexene becomes a competing process S9). As the yields for cyclohexene cyclopropanation with copper catalysts [except Cu(OTf)2] are low (Table 2), such variations in stereoselectivity are not very significant in terms of absolute yields anyway. [Pg.108]

It has already been mentioned that prochirality of the olefin is not necessary for successful enantioselective cyclopropanation with an alkyl diazoacetate in the presence of catalysts 207. What happens if a prochiral olefin and a non-prochiral diazo compound are combined Only one result provides an answer to date The cyclopropane derived from styrene and dicyanodiazomethane shows only very low optical induction (4.6 % e.e. of the (25) enantiomer, catalyst 207a) 9S). Thus, it can be concluded that with the cobalt chelate catalysts 207, enantioface selectivity at the olefin is generally unimportant and that a prochiral diazo compound is needed for efficient optical induction. As the results with chiral copper 1,3-diketonates 205 and 2-diazodi-medone show, such a statement can not be generalized, of course. [Pg.166]

Use of a chiral diazo ester proved less rewarding in terms of enantioselective cyclopropanation. Only very low enantiomeric excesses were obtained when styrene was cyclopropanated with the carbenoid derived from diazoacetic esters 219 bearing a chiral ester residue 214). [Pg.171]

It has been widely accepted that the carbene-transfer reaction using a diazo compound and a transition metal complex proceeds via the corresponding metal carbenoid species. Nishiyama et al. characterized spectroscopically the structure of the carbenoid intermediate that underwent the desired cyclopropanation with high enantio- and diastereoselectivity, derived from (91).254,255 They also isolated a stable dicarbonylcarbene complex and demonstrated by X-ray analysis that the carbene moiety of the complex was almost parallel in the Cl—Ru—Cl plane and perpendicular to the pybox plane (vide infra).255 These results suggest that the rate-determining step of metal-catalyzed cyclopropanation is not carbenoid formation, but the carbene-transfer reaction.254... [Pg.249]

As expected, DMFL reacts with styrene to give the appropriate cyclopropane. Irradiation of the diazo-compound in benzene containing the labeled a-methylstyrene gives cyclopropane with essentially complete retention of stereochemistry (Table 6). [Pg.345]

Furthermore, cyclopropane structures have often served as intermediates in organic synthesis. For these reasons, olefin cyclopropanation has proved to be a useful tool for synthetic organic chemists. This has led to the development of several methods for cyclopropanation reactions,91 including the metal-catalyzed reactions of diazo compounds with olefins, as well as the Simmons-Smith reaction. [Pg.313]

In an approach to the stereocontrolled creation of the acyclic side-chain of tetracyclic triterpenoids and other natural products, Trost and his colleagues have converted the acyclic starting compound (29) into the cyclopropanoid intermediate (31) via the diazo-ketone (30). The key step in the scheme is the cleavage of the cyclopropane with lithium dimethylcuprate to give (32). The stereochemistry at C-7 is determined by the configuration of the double bond in (29). The c.d. and u.v. spectra of a series of triterpenoid olefins have been measured. The Scott-Wrixon rules can be used to correlate the sign of the c.d. curves with molecular structure. A... [Pg.154]

Conjugated dienes, styrenes and electron-rich alkenes are cyclopropanated with ethyl diazoacetate using a triarylamminium salt of appropriate oxidation potential as a cata-lyst/initiator (equation 96)185. These reactions are initiated by electron transfer from the unsaturated substrate to the amminium ion and the double additions of the diazo esters to the conjugated dienes are effectively suppressed. Cyclopropanes geminally bearing two... [Pg.290]

Electron-withdrawing substituents generally increase diazo compounds stability toward decomposition. Dicarbonyl diazomethane, which bears two carbonyl groups flanking the diazomethane carbon, are more stable than diazo compounds with only one carbonyl substituent. In general, metal catalysed decomposition of dicarbonyl diazomethane requires higher temperature than does monocarbonyl substituted diazomethane. As indicated before, rhodium(II) carboxylates are the most active catalysts for diazo decomposition. With dicarbonyl diazomethane, the rhodium(II) carboxylate-promoted cyclopropanation process can also be carried out under ambient conditions to afford a high yield of products. [Pg.676]

Stereosectivity is a broad term. The stereoselectivity in cyclopropanation which has been discussed in the above subsection, in fact, can also be referred to as diastereoselectivity. In this section, for convenience, the description of diastereoselectivity will be reserved for selectivity in cyclopropanation of diazo compounds or alkenes that are bound to a chiral auxiliary. Chiral diazoesters or chiral Ar-(diazoacetyl)oxazolidinone have been applied in metal catalysed cyclopropanation. However, these chiral diazo precursors and styrene yield cyclopropane products whose diastereomeric excess are less than 15% (equation 129)183,184. The use of several a-hydroxy esters as chiral auxiliaries for asymmetric inter-molecular cyclopropanation with rhodium(II)-stabilized vinylcarbenoids have been reported by Davies and coworkers. With (R)-pantolactone as the chiral auxiliary, cyclopropanation of diazoester 144 with a range of alkenes provided c yield with diastereomeric excess at levels of 90% (equation 130)1... [Pg.695]

A variation on the thermal reactions of diazo compounds with alkenes is the decomposition of salts of sulfonylhydrazones. This procedure, known as the BamfoTd-Stevens reaction, is believed to occur via the formation of diazo compounds. Subsequent 1,2-hydrogen shifts generally lead to the formation of alkenes as the final products. Cyclopropanes may also be formed as the result of intramolecular 1,3-C—H insertion reactions or when the original hydrazone substrate contains a remote alkenic group as a site for intramolecular cyclopropanation.10... [Pg.954]

Although C—H insertion reactions rarely occur in intermolecular reactions with diazoacetates, these are common side reactions with diazomalonates3132 (equation 10) and diazo ketones (with a-allyl vinyl ethers).33 Several mechanistic pathways are available to generate the products of an apparent direct C—H insertion reaction and these include dipolar intermediates, ir-allyl complexes and ring opening of cyclopropanes.1 Oxidative problems due to the presence of oxygen are common with copper catalysts, but these are rarely encountered with rhodium catalysts except in systems where the carbenoid is ineffectively captured.34... [Pg.1036]

Rh(II) carboxylates, especially Rh2(OAc)4> have emerged as the most generally effective catalysts for metal carbene transformations [7-10] and thus interest continues in the design and development of dirhodium(II) complexes that possess chiral51igands. They are structurally well-defined, with D2h symmetry [51] and axial coordination sites at which carbene formation occurs in reactions with diazo compounds. With chiral dirhodium(II) carboxylates the asymmetric center is located relatively far from the carbene center in the metal carbene intermediate. The first of these to be reported with applications to cyclopropanation reactions was developed by Brunner [52], who prepared 13 chiral dirhodium(II) tetrakis(car-boxylate) derivatives (16) from enantiomerically pure carboxylic acids RlR2R3CC OOH with substituents that were varied from H, Me, and Ph to OH, NHAc, and CF3. However, reactions performed between ethyl diazoacetate and styrene yielded cyclopropane products whose enantiopurities were less than 12% ee, a situation analogous to that encountered by Nozaki [2] in the first applications of chiral Schiff base-Cu(II) catalysts. [Pg.203]

Reaction of a ruthenium carbenoid species, formed from a diazo compound, with an alkene produces cyclopropanes. A variety of ruthenium catalysts, notably chiral catalysts, have been developed to lead to efficient asymmetric cyclopropanation of alkenes as in the following example [105] (Eq. 81). [Pg.33]

Salomon RG, Kochi JK (1973) Copper(I) catalysis in cyclopropanations with diazo compounds. Role of olefin coordination, J Am Chem Soc 95 3300-3310... [Pg.592]

Figure 7.18 Catalytic cycle for the cyclopropanation of alkene by diazo compounds with a rhodium-porphyrin complex as the precatalyst. Figure 7.18 Catalytic cycle for the cyclopropanation of alkene by diazo compounds with a rhodium-porphyrin complex as the precatalyst.
CO)2Fe (THF) BFT A transition state model for the syn stereoselective cyclo-propanations of alkenes with diazoacetic ester by Rh-porphyrin catalysts has been proposed. Alkenes , conjugated dienes and enol ethers are stereoselectively cyclopropanated with Rh(II) -stabilized 1- (alkoxycarbonyl)vinyl carbenoids derived from the diazo precursors and Rh2(OAc)4 (equation 95). The Cu(acac)2-catalyzed reactions of Me3SiCH2COCHN2 with alkenes provide the expected adducts in good yields ". ... [Pg.290]

The control of carbene reactions from diazo compounds as precursors is classically mediated by copper catalysts and all the carbene reactions discussed in Section 8 are, in fact, improved by such catalysts. Moreover, the use of ligands on the metal allows some control of the stereochemistry of the products, the most striking example being the asymmetric synthesis of cyclopropanes with an excellent optical yield (>90S ce) [35]. [Pg.278]

The reaction is carried out at ambient temperature and nearly complete enantioselectivity (>99%) is observed for mono- and 1,1-disubstituted olefins with diazoacetates. With all copper catalysts, the transkis selectivities in the cyclopropanation of mono-substituted olefins are only moderate. The transkis ratio depends, in this case, mainly on the structure of the diazo ester rather than the chiral ligand (eq 2). It increases with the steric bulk of the ester group of the diazo compound. With the BHT ester, the more stable trans isomer is formed with selectivities up to >10 1. The steric hindrance usually prevents ester hydrolysis, but the BHT group can be removed by reduction with LiAlHj. The trans isomer is even enriched by the reduction procedure because the cis isomer reacts more slowly. [Pg.109]

MO calculations have been carried out on the isomerization of cyclopropane to propene, and the MNDO method has been used to study the reaction pathway and to optimize the structure of reactant, transition structure, and product of the ring opening reaction of bicyclo[1.1.0]butane. Various methods have been employed to estimate the rate constants for ring opening of the 2-cyclopropyl-2-propyl radical. 1-Acceptor-1-sulfenyl-substituted 2-vinylcyclopropanes of the type (430) have been found to afford 6-sulfenyl-a,jS y, -unsaturated carboxylic esters and nitriles (431) upon treatment with acid, by a process which involves C(l)—C(2) bond fission and a novel 1,5-sulfenyl rearrangement (see Scheme 110). It has been shown that the benzophenone-sensitized photolysis of vinyl norcaradiene derivatives, such as 5-(2-methylprop-l-enyl)-3-oxatricyclo[4.4.0.0 ]deca-7,9-dien-4-ones (432), results in the regioselective cleavage of only one of the cyclopropyl c-bonds to afford isochroman-3-one derivatives (433). It has been reported that the major product obtained from the reaction of structurally diverse a-diazo ketones with an electron-rich alkene in the... [Pg.577]

There are many cases where the [3 + 2]-cycloaddition leads to cycloadducts with high enantioselectivity. Cycloaddition of diazo esters with a cobalt catalyst having a chiral ligand leads to cyclopropane derivatives with good enantioselectivity. ... [Pg.1191]


See other pages where Diazo cyclopropanation with is mentioned: [Pg.210]    [Pg.210]    [Pg.219]    [Pg.88]    [Pg.177]    [Pg.245]    [Pg.15]    [Pg.552]    [Pg.610]    [Pg.340]    [Pg.532]    [Pg.590]    [Pg.261]    [Pg.290]    [Pg.304]    [Pg.692]    [Pg.953]    [Pg.1045]    [Pg.577]    [Pg.577]    [Pg.131]    [Pg.261]    [Pg.304]    [Pg.692]    [Pg.106]    [Pg.44]   
See also in sourсe #XX -- [ Pg.194 ]




SEARCH



Cyclopropanation with diazo compounds

Cyclopropane, diazo decomposition with

With cyclopropane

© 2024 chempedia.info