Big Chemical Encyclopedia

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

Articles Figures Tables About

Transition structures cyclic

The orbital phase theory can be applied to cyclically interacting systems which may be molecules at the equilibrium geometries or transition structures of reactions. The orbital phase continuity underlies the Hueckel rule for the aromaticity and the Woodward-Hoffmann rule for the stereoselection of organic reactions. [Pg.95]

Scheme 17 Cyclic orbital interactions at the transition structures of electrophilic aromatic substitutions... Scheme 17 Cyclic orbital interactions at the transition structures of electrophilic aromatic substitutions...
Orbitals interact in cyclic manners in cyclic molecules and at cyclic transition structures of chemical reactions. The orbital phase theory is readily seen to contain the Hueckel 4n h- 2 ti electron rule for aromaticity and the Woodward-Hof nann mle for the pericyclic reactions. Both rules have been well documented. Here we review the advances in the cyclic conjugation, which cannot be made either by the Hueckel rule or by the Woodward-Hoffmann rule but only by the orbital phase theory. [Pg.111]

Summary of the Relationship between Diastereoselectivity and the Transition Structure. In this section we considered simple diastereoselection in aldol reactions of ketone enolates. Numerous observations on the reactions of enolates of ketones and related compounds are consistent with the general concept of a chairlike TS.35 These reactions show a consistent E - anti Z - syn relationship. Noncyclic TSs have more variable diastereoselectivity. The prediction or interpretation of the specific ratio of syn and anti product from any given reaction requires assessment of several variables (1) What is the stereochemical composition of the enolate (2) Does the Lewis acid promote tight coordination with both the carbonyl and enolate oxygen atoms and thereby favor a cyclic TS (3) Does the TS have a chairlike conformation (4) Are there additional Lewis base coordination sites in either reactant that can lead to reaction through a chelated TS Another factor comes into play if either the aldehyde or the enolate, or both, are chiral. In that case, facial selectivity becomes an issue and this is considered in Section 2.1.5. [Pg.78]

Scheme 6.21. Thermal Eliminations Via Cyclic Transition Structures... Scheme 6.21. Thermal Eliminations Via Cyclic Transition Structures...
Aldol addition and related reactions of enolates and enolate equivalents are the subject of the first part of Chapter 2. These reactions provide powerful methods for controlling the stereochemistry in reactions that form hydroxyl- and methyl-substituted structures, such as those found in many antibiotics. We will see how the choice of the nucleophile, the other reagents (such as Lewis acids), and adjustment of reaction conditions can be used to control stereochemistry. We discuss the role of open, cyclic, and chelated transition structures in determining stereochemistry, and will also see how chiral auxiliaries and chiral catalysts can control the enantiose-lectivity of these reactions. Intramolecular aldol reactions, including the Robinson annulation are discussed. Other reactions included in Chapter 2 include Mannich, carbon acylation, and olefination reactions. The reactivity of other carbon nucleophiles including phosphonium ylides, phosphonate carbanions, sulfone anions, sulfonium ylides, and sulfoxonium ylides are also considered. [Pg.1334]

This chapter also discusses several (3-elimination reactions that proceed through cyclic transition structures. [Pg.1336]

Tietze, L. F., Pfeiffer, T., Schuffenhauer, A., 1998, Stereoselective Intramolecular Hetero Diels-Alder Reactions of Cyclic Benzylidenesulfoxides and DFT Calculations on the Transition Structures , Eur. J. Org. Chem., 2733. [Pg.302]

The bonding hydrogen orbital overlaps simultaneously with the / -orbitals on both the terminal carbon atoms and a cyclic transition structure is formed in which the C-l becomes a / -orbital and... [Pg.74]

An extensive review of the hetero-Diels-Alder reactions of 1-oxabuta-1,3-dienes has been published. Ab initio calculations of the Diels-Alder reactions of prop-2-enethial with a number of dienophiles show that the transition states of all the reactions are similar and synchronous.Thio- and seleno-carbonyl compounds behave as superdienophiles in Diels-Alder reactions with cyclic and aryl-, methyl-, or methoxy-substituted open-chain buta-1,3-dienes.The intramolecular hetero-Diels-Alder reactions of 4-benzylidine-3-oxo[l,3]oxathiolan-5-ones (100) produce cycloadducts (101) and (102) in high yield and excellent endo/exo-selectivity (Scheme 39). A density functional theoretical study of the hetero-Diels-Alder reaction between butadiene and acrolein indicates that the endo s-cis is the most stable transition structure in both catalysed and uncatalysed reactions.The formation and use of amino acid-derived chiral acylnitroso hetero-Diels-Alder reactions in organic synthesis has been reviewed. The 4 + 2-cycloadditions of A-acylthioformamides as dienophiles have been reviewed. ... [Pg.475]

It is well accepted that the high diastereospecificify of aldehyde allylboration reactions is a consequence of the compact cyclic transition structure. Theoretical calculations have shown that the chairlike transition structure shown in Scheme 1 and Fig. 1 is the lowest in energy relative to other possibilities such as the twist-boat conformation. With boronate reagents, it has also been suggested that a weak hydrogen bond between the axial boronate oxygen and the hydrogen of the polarized formyl unit contribntes to the preference for the transition structme with the aldehyde substituent in the psendo-eqnatorial position. ... [Pg.9]

Recently, the first examples of catalytic enantioselective preparations of chiral a-substituted allylic boronates have appeared. Cyclic dihydropyranylboronate 76 (Fig. 6) is prepared in very high enantiomeric purity by an inverse electron-demand hetero-Diels-Alder reaction between 3-boronoacrolein pinacolate (87) and ethyl vinyl ether catalyzed by chiral Cr(lll) complex 88 (Eq. 64). The resulting boronate 76 adds stereoselectively to aldehydes to give 2-hydroxyalkyl dihydropyran products 90 in a one-pot process.The diastereoselectiv-ity of the addition is explained by invoking transition structure 89. Key to this process is the fact that the possible self-allylboration between 76 and 87 does not take place at room temperature. Several applications of this three-component reaction to the synthesis of complex natural products have been described (see section on Applications to the Synthesis of Natural Products ). [Pg.39]

Intramolecular additions generally follow the same trends of stereoselectivity as observed in the bimolecular reactions. Eor example, allylic boronates ( )- and (Z)-118 provide the respective trans- and cis-fused products of intramolecular aUylation. As shown with allylboronate ( )-118, a Yb(OTf)3-catalyzed hydrolysis of the acetal triggers the intramolecular aUylboration and leads to isolation of the trans-fused product 119 in agreement with the usual cyclic transition structure (Eq. 96). [Pg.49]

A masked allylic boron unit can be revealed through a transition-metal-catalyzed borylation reaction. For example, a one-pot borylation/allylation tandem process based on the borylation of various ketone-containing allylic acetates has been developed. The intramolecular allylboration step is very slow in DMSO, which is the usual solvent for these borylations of allylic acetates (see Eq. 33). The use of a non-coordinating solvent like toluene is more suitable for the overall process provided that an arsine or phosphine ligand is added to stabilize the active Pd(0) species during the borylation reaction. With cyclic ketones such as 136, the intramolecular allylation provides cis-fused bicyclic products in agreement with the involvement of the usual chairlike transition structure, 137 (Eq. 102). [Pg.52]

FIGURE 26. Transition structures for the epoxidation of selected cyclic alkenes with peroxyformic acid (PEA), optimized at the B3LYP/6-31+G(d,p) level of theory. The classical activation barriers are given at B3LYP/6-311+G(3df,2p)//B3LYP/6-31+G(d,p)... [Pg.61]

Cyclic allylic alcohols have different steric requirements than the acyclic substrates discussed above. Sarzi-Amade and coworkers addressed the mechanism of epoxida-tion of 2-cyclohexen-l-ol by locating all the transition structures (TSs) for the reaction of peroxyformic acid (PFA) with both pseudoequatorial and pseudoaxial cyclohexenol con-formers. Geometry optimizations were performed at the B3LYP/6-31G level, and the total energies were refined with single-point B3LYP/6-311- -G //B3LYP/6-31G calculations. [Pg.67]

The observed stereochemical outcome suggested that a cisoid configuration of the crotyl metal was kinetically favored in the cyclic transition state. Indeed, recent calculations supported that transition structures with cisoid crotylmetals are more stable whatever the olefinic configuration of the alkenyl metal142. [Pg.915]

Various bimolecular assemblies that have been proposed for the transition state are shown in Scheme 13 (14, 19a, 20g). Bicyclic transition state A involves transfer of bridging alkyl group (R) to the terminally located aldehyde, while transition structure B involves reaction between terminal R and bridging aldehyde. The reaction may proceed via mono-cyclic, boat-like six-membered transition state C. Transition structures of types B and C were originally proposed for the reactions of orga-noaluminum compounds and carbonyl substrates (26, 27). Ab initio calculations suggest that methyllithium dimer reacts with formaldehyde through a bicyclic transition state related to A (28). The dinuclear Zn... [Pg.141]

Pericyclic reactions are the third distinct class. They have cyclic transition structures in which all bond-forming and bond-breaking takes place in concert, without the formation of an intermediate. The Diels-Alder reaction and the Alder ene reaction are venerable examples. The curly arrows can be drawn in either direction—clockwise, as here, but equally well anti- clockwise. They could even be drawn with fishhook arrows, and would still... [Pg.2]

All the reactions described so far have mobilized six electrons in the transition structure. Other numbers are possible, notably a few [8+2] and [6+4] cycloadditions involving ten electrons in the cyclic transition structure. It is no accident, as we shall see in the next chapter, that these reactions have the same number of electrons (4n+2) as aromatic rings. [Pg.15]


See other pages where Transition structures cyclic is mentioned: [Pg.83]    [Pg.122]    [Pg.473]    [Pg.596]    [Pg.1335]    [Pg.1337]    [Pg.1341]    [Pg.41]    [Pg.123]    [Pg.561]    [Pg.175]    [Pg.218]    [Pg.52]    [Pg.55]    [Pg.11]    [Pg.22]    [Pg.24]    [Pg.41]    [Pg.43]    [Pg.68]    [Pg.83]    [Pg.744]    [Pg.148]    [Pg.67]    [Pg.151]    [Pg.157]    [Pg.151]    [Pg.21]   
See also in sourсe #XX -- [ Pg.162 , Pg.163 , Pg.189 , Pg.217 , Pg.225 , Pg.245 ]




SEARCH



Cyclic structures

Transition cyclic

© 2024 chempedia.info