Big Chemical Encyclopedia

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

Articles Figures Tables About

Electron deficient Lewis acid coordination

In solid Bep2, a complex network is formed with a Be atom coordination number of 4 (see Figure 3.7). BeCl2 dimerizes to a 3-coordinate structure in the vapor phase, but the linear monomer is formed at high temperatures. This monomeric structure is unstable due to the electronic deficiency at Be in the dimer and the network formed in the solid-state, the halogen atoms share lone pairs with the Be atom in an attempt to fill beryllium s valence shell. The monomer is still frequently drawn as a singly bonded structure, with only four electrons around the beryllium and the ability to accept lone pairs of other molecules to relieve its electronic deficiency (Lewis acid behavior, discussed in Chapter 6). [Pg.50]

Two types of interactions have been shown to be involved in superelectrophilic species. Superelectrophiles can be formed by the further interaction of a conventional cationic electrophile with Brpnsted or Lewis acids (eq 16).23 Such is the case with the further protonation (protosolvation) or Lewis acid coordination of suitable substitutents at the electron deficient site, as for example in carboxonium cations. The other involves further protonation or complexation formation of a second proximal onium ion site, which results in superelectrophilic activation (eq 17).24... [Pg.11]

More recently, Porta and co-workers [6] applied similar considerations of the polar effects to a new one-pot multicomponent process for the addition of nucleophilic radicals to aldimines, generated in situ in the presence of Ti(IV). In analogy with the Minisci reaction, Ti(IV), which acts as a Lewis acid, coordinates the nitrogen of the imine, strongly increasing the electron-deficient character of the carbon in the a-posilion and thus the reactivity of the imine toward nucleophilic radicals. This reaction, as well as the Minisci one, represents a useful route for the synthesis of a variety of poly-functionalized derivatives of chemical and biochemical relevance. [Pg.338]

Mewald M, Froehlich R, Oestreich M (2011) An axially chiral, electron-deficient borane synthesis, coordination chemistry, lewis acidity, and reactivity. Cbem Eur J 17 9406-9414... [Pg.227]

Jorgensen and coworkers reported a chiral Cu(ll)-catalyzed 1,3-dipolar cycloaddition reaction between electrophilic nitrones (304) and electron-rich alkenes (305) (Scheme 17.68) [99]. In contrast with the more widely studied cycloaddition of nitrones to electron-deficient alkenes in which Lewis acid coordination results in lowering of the alkene LUMO, this methodology utilizes Lewis acid coordination to electron-deficient nitrones (304) to lower the nitrone LUMO. The result is a change from a HOMOnitrone-LUMO controlled reaction path to a HOMOaikene-LUMOnitrone Controlled reaction path. Cu(OTf)2/t-Bu-BOX (13)-catalyzed cycloadditions proceed in moderate to good yields and selectivi-ties with a pentacoordinated complex (308) proposed to account for the observed stereoselectivity. [Pg.428]

Dipolar cydoadditions are one of the most useful synthetic methods to make stereochemically defined five-membered heterocydes. Although a variety of dia-stereoselective 1,3-dipolar cydoadditions have been well developed, enantioselec-tive versions are still limited [29]. Nitrones are important 1,3-dipoles that have been the target of catalyzed enantioselective reactions [66]. Three different approaches to catalyzed enantioselective reactions have been taken (1) activation of electron-defident alkenes by a chiral Lewis acid [23-26, 32-34, 67], (2) activation of nitrones in the reaction with ketene acetals [30, 31], and (3) coordination of both nitrones and allylic alcohols on a chiral catalyst [20]. Among these approaches, the dipole/HOMO-controlled reactions of electron-deficient alkenes are especially promising because a variety of combinations between chiral Lewis acids and electron-deficient alkenes have been well investigated in the study of catalyzed enantioselective Diels-Alder reactions. Enantioselectivities in catalyzed nitrone cydoadditions sometimes exceed 90% ee, but the efficiency of catalytic loading remains insufficient. [Pg.268]

The coordination of the dienophile to a Lewis acid (in the calculations a proton was used as the Lewis acid) leads also to an increase in regioselectivity. The re-gioselectivity of reactions of electron-rich, or conjugated dienes, with electron-deficient dienophiles is also controlled hy the diene HOMO-dienophile LUMO interaction. From Fig. 8.2 it appears that the difference in magnitudes of the LUMO coefficients at carhon atoms 1 and 2 of acrolein (Ci -C2 = 0.20) is smaller than the same difference for protonated acrolein (Ci -C2 = 0.30-0.43) so that the reaction of the latter should he considerable more regioselective than the former in accordance with the experimental results [3]. [Pg.304]

Some experimental evidences are in agreement with this proposed mechanism. For example, coordinating solvents like diethyl ether show a deactivating effect certainly due to competition with a Lewis base (149). For the same reason, poor reactivity has been observed for the substrates carrying heteroatoms when an aluminum-based Lewis acid is used. Less efficient hydrovinylation of electron-deficient vinylarenes can be explained by their weaker coordination to the nickel hydride 144, hence metal hydride addition to form key intermediate 146. Isomerization of the final product can be catalyzed by metal hydride through sequential addition/elimination, affording the more stable compound. Finally, chelating phosphines inhibit the hydrovinylation reaction. [Pg.320]

It is clear from the above observations that pyridine molecule interacts on the catalyst surface in the following three modes (1) interaction of the N lone pair electron and the H atom of the OH group, (2) transfer of a proton from surface OH group to the pyridine forming a pyridinium ion (Bronsted acidity), and (3) pyridine coordination to an electron deficient metal atom (Lewis acidity). Predominant IR bands, vga and vigb, confirms that the major contribution of acidity is due to Lewis acid sites from all compositions. Between the above two modes of vibrations, Vsa is very sensitive with respect to the oxidation state, coordination symmetry and cationic environment [100]. A broad feature for v a band on Cu containing... [Pg.180]

The pretreatment temperature is an important factor that influences the acidic/ basic properties of solids. For Brpnsted sites, the differential heat is the difference between the enthalpy of dissociation of the acidic hydroxyl and the enthalpy of protonation of the probe molecule. For Lewis sites, the differential heat of adsorption represents the energy associated with the transfer of electron density toward an electron-deficient, coordinatively unsaturated site, and probably an energy term related to the relaxation of the strained surface [147,182]. Increasing the pretreatment temperature modifies the surface acidity of the solids. The influence of the pretreatment temperature, between 300 and 800°C, on the surface acidity of a transition alumina has been studied by ammonia adsorption microcalorimetry [62]. The number and strength of the strong sites, which should be mainly Lewis sites, have been found to increase when the temperature increases. This behavior can be explained by the fact that the Lewis sites are not completely free and that their electron pair attracting capacity can be partially modified by different OH group environments. The different pretreatment temperatures used affected the whole spectrum of adsorption heats... [Pg.227]

Being an electron deficient compound, boron trifluoride forms complexes with Lewis bases and compounds that have unshared pair(s) of electrons. With ammonia, it forms boron trifluoride ammonia. Similar coordination compounds are formed with monoethylamine, BF3-NH2C2H5 diethyl ether, CH3CH20(BF3)CH2CH3 and methanol, BF3—OHCH3. It forms a sohd complex HNO3-2BF3 with concentrated nitric acid. [Pg.135]

The oxonium ylide mechanism requires a bifunctional acid-base catalyst. The validity of the oxonium ylide mechanism on zeolites was questioned459,461,464 because zeolites do not necessarily possess sufficiently strong basic sites to abstract a proton from the trimethyloxonium ion to form an ylide. It should, however, be pointed out, as emphasized by Olah,447,465 that over solid acid-base catalysts (including zeolites) the initial coordination of an electron-deficient (i.e., Lewis acidic) site of the catalysts allows formation of a catalyst-coordinated dimethyl ether complex. It then can act as an oxonium ion forming the catalyst-coordinated oxonium ylide complex (10) with the participation of surface bound CH30 ions ... [Pg.121]

Nitrones can be activated mainly in two different ways for the 1,3-dipolar cycloaddition with alkenes. In the reaction between a nitrone and an electron-deficient alkene, such as an 7.,p-unsaturated carbonyl compound (a normal electron-demand reaction), it is primarily controlled by the interaction between HOMOnitrone-LUMOalkene (Scheme 12.64). By coordination of a Lewis acid (LA) catalyst to the cc,p-unsaturated carbonyl compound, the LUMOaikene energy decreases and a better interaction with the nitrone can take place (16,17). [Pg.710]

Cince the catalytic activity of synthetic zeolites was first revealed (1, 2), catalytic properties of zeolites have received increasing attention. The role of zeolites as catalysts, together with their catalytic polyfunctionality, results from specific properties of the individual catalytic reaction and of the individual zeolite. These circumstances as well as the different experimental conditions under which they have been studied make it difficult to generalize on the experimental data from zeolite catalysis. As new data have accumulated, new theories about the nature of the catalytic activity of zeolites have evolved (8-9). The most common theories correlate zeolite catalytic activity with their proton-donating and electron-deficient functions. As proton-donating sites or Bronsted acid sites one considers hydroxyl groups of decationized zeolites these are formed by direct substitution of part of the cations for protons on decomposition of NH4+ cations or as a result of hydrolysis after substitution of alkali cations for rare earth cations. As electron-deficient sites or Lewis acid sites one considers usually three-coordinated aluminum atoms, formed as a result of dehydroxylation of H-zeolites by calcination (8,10-13). [Pg.242]

Lewis acidity, arising from electron deficient centers, does not exist in silica. In alumina this acidity is associated with surface aluminum ions which are coordinated in a tetrahedral manner. Such sites may be generated either by removal of water coordinated to an aluminum center or by dehydration of two adjacent hydroxyl groups above 500°C. [Pg.238]

Compounds 4.75 and 4.76 formally contain tetravalent silicon, which is not electron-deficient. It is, however, a strong Lewis acid, and on reaction with F , five-coordinate silicon compounds are obtained. In the case of 4.75, the fluoride anion is localised mainly on the boron atom, although it does display dynamic behaviour involving hopping between boron and silicon. Compound 4.76 as the KF adduct contains two five-coordinate silicon atoms that chelate the F anion. The coordination sphere of the K+ counter-ion is completed by a molecule of [18] crown-6 (Figure 4.32a). [Pg.304]

The selective functionalization of heterocycles is of particular importance, because of the ubiquity of these structures in natural products and pharmaceutical agents. Direct utilization of a C-H bond [1] of heterocycles is a promising method for the preparation of heterocycles because no pre-functionalization is required. Although Friedel-Crafts acylation is the most commonly used method for introduction of keto functionality on an aromatic ring, it is not often applicable to N-heteroarenes because of deactivation of the Lewis acids by the coordination of N-heteroarenes and the electron-deficient aromatic character of N-heteroarenes. [Pg.175]


See other pages where Electron deficient Lewis acid coordination is mentioned: [Pg.226]    [Pg.11]    [Pg.344]    [Pg.234]    [Pg.118]    [Pg.152]    [Pg.203]    [Pg.183]    [Pg.8]    [Pg.214]    [Pg.99]    [Pg.14]    [Pg.295]    [Pg.152]    [Pg.794]    [Pg.93]    [Pg.159]    [Pg.11]    [Pg.146]    [Pg.151]    [Pg.84]    [Pg.602]    [Pg.102]    [Pg.183]    [Pg.164]    [Pg.124]    [Pg.47]    [Pg.141]    [Pg.283]    [Pg.302]    [Pg.299]    [Pg.76]   
See also in sourсe #XX -- [ Pg.317 ]




SEARCH



Coordinates electron

Electron deficiency

Electronic coordinate

© 2024 chempedia.info