Big Chemical Encyclopedia

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

Articles Figures Tables About

Reversible radical reaction

The carbon-oxygen double bond of the carbonyl group is opened, and the hydrogen sulfite radical is added. An increase in temperature reverses the reaction more easily for ketones than for aldehydes. [Pg.1169]

As the temperature approaches the NTC zone, the reversibility of reaction 2 comes into play and the steady-state concentration of alkyl radicals rises. There is a competing irreversible reaction of oxygen with radicals containing an alpha hydrogen which produces a conjugate olefin (eq. 23). [Pg.339]

Most radicals are transient species. They (e.%. 1-10) decay by self-reaction with rates at or close to the diffusion-controlled limit (Section 1.4). This situation also pertains in conventional radical polymerization. Certain radicals, however, have thermodynamic stability, kinetic stability (persistence) or both that is conferred by appropriate substitution. Some well-known examples of stable radicals are diphenylpicrylhydrazyl (DPPH), nitroxides such as 2,2,6,6-tetramethylpiperidin-A -oxyl (TEMPO), triphenylniethyl radical (13) and galvinoxyl (14). Some examples of carbon-centered radicals which are persistent but which do not have intrinsic thermodynamic stability are shown in Section 1.4.3.2. These radicals (DPPH, TEMPO, 13, 14) are comparatively stable in isolation as solids or in solution and either do not react or react very slowly with compounds usually thought of as substrates for radical reactions. They may, nonetheless, react with less stable radicals at close to diffusion controlled rates. In polymer synthesis these species find use as inhibitors (to stabilize monomers against polymerization or to quench radical reactions - Section 5,3.1) and as reversible termination agents (in living radical polymerization - Section 9.3). [Pg.14]

A stereoelectronic requirement in radical addition to carbon-carbon double bonds first became apparent from studies on radical cyelization and the reverse (Fragmentation) reactions.54 56 It provides a rationalization for the preferential formation of the less thermodynamically stable exo-product (i.e. head addition) from the cyelization of to-alkcny] radicals (16 - Scheme 1.5). s 57 G4... [Pg.23]

Much has been learned in recent years about the 00 dimer , O2O2, produced in reaction 17. It is actually dichlorine peroxide, OOOCl its geometry is now well established from submillimeter wave spectroscopy (15). Photolysis of OOOO around 310 nm the atmospherically important wavelengths -- yields chlorine atoms and ClOO radicals (16), as given in reaction 18, rather than two OO radicals, even though QO-OQ is the weakest bond (it has a strength of about 17 Kcal/mol (17)). Thermal decomposition of QOOQ (the reverse of reaction 17) occurs very fast at room temperature, but more slowly at polar stratospheric temperatures. Hence, photolysis is the predominant destruction path for CIOOQ in the polar stratosphere and two Q atoms are produced for each ultraviolet photon absorbed. [Pg.32]

This comparison suggests that of these two similar reactions, only alkene additions are likely to be a part of an efficient radical chain sequence. Radical additions to carbon-carbon double bonds can be further enhanced by radical stabilizing groups. Addition to a carbonyl group, in contrast, is endothermic. In fact, the reverse fragmentation reaction is commonly observed (see Section 10.3.6) A comparison can also be made between abstraction of hydrogen from carbon as opposed to oxygen. [Pg.956]

It is important to select stoichiometric co-reductants or co-oxidants for the reversible cycle of a catalyst. A metallic co-reductant is ultimately converted to the corresponding metal salt in a higher oxidation state, which may work as a Lewis acid. Taking these interactions into account, the requisite catalytic system can be attained through multi-component interactions. Stereoselectivity should also be controlled, from synthetic points of view. The stereoselective and/or stereospecific transformations depend on the intermediary structure. The potential interaction and structural control permit efficient and selective methods in synthetic radical reactions. This chapter describes the construction of the catalytic system for one-electron reduction reactions represented by the pinacol coupling reaction. [Pg.65]

For a simple displacement reaction involving atoms or radicals, such as A + BC -> AB + C, where the reaction is written in the exothermic direction, the activation energy is 5.5% of the dissociation energy of the bond which is broken. For the reverse endothermic reaction, the standard energy change of reaction must be added to this quantity in order to obtain the activation energy. [Pg.89]

Finally, Crich [31] and Danishefsky [32] demonstrated that enantioselective hexahydropyrroloindole formation could precede C3-quatemary center formation. The carbon-carbon bond formation could be accomplished through a stereoretentive cationic or radical process as demonstrated through the reverse-prenylation reaction in Danishefsky s synthesis of amouramine [32a, b] or the allylation reaction en route to Crich s synthesis of (+)-debromofhistramine [33], respectively. [Pg.218]

The theory of chain co-oxidation of binary mixtures of organic compounds was described in Chapter 5. The experimental study of co-oxidation of alcohols (HRiOH) and hydrocarbon R H opens the way to measure the rate constants of one chosen peroxyl radical R OO with several alcohols HRiOH and on the reverse, the chosen alcohol HR1 OH with several peroxyl radicals RiOO. The parameters of co-oxidation of alcohols and hydrocarbons are collected in Table 7.6. The absolute values of peroxyl radical reactions with alcohols were calculated from these data using the values of kp from Table 2.8 (see Table 7.7). [Pg.297]

By extrapolation from the behavior of diphenylpicrylhydrazyl, the dissociation of tetraaryl hydrazines into colored, reactive substances in solution at high temperatures is probably a radical reaction. It is only partly reversible, owing to disproportionation. [Pg.66]

The above three examples involved reactions where the electron transfer takes place from the metal to the organic substrate. The reverse scenario can also be used in radical reactions via oxidative generation of cationic radical species, which can undergo coupling reactions. Kurihara et al. have used chiral ox-ovanadium species as a one-electron transfer oxidant to silylenol ethers in a hetero-coupling process [165]. Treatment of 246 with a catalyst prepared in situ from VOCI3/chiral alcohol/MS 4 A followed by addition of 247 provided the coupling product 248 (Scheme 63). 8-Phenyl menthol 251 was found to be... [Pg.167]

At the end of section 3.1 we addressed a common circumstance in gas-phase thermochemical kinetics studies. For many reactions, there is not enough experimental information to determine A //°, and a negligible barrier for the product recombination reaction is often assumed. The same ideas can be applied for reactions in solution When D and E are radicals, it is frequently accepted that the reverse of reaction 3.33 is diffusion controlled and that A // , has a value of 8kJ mol-1. [Pg.45]

If we make the assumption that the reverse of reaction 15.5 is diffusion-controlled and assume that the activation enthalpy for the acyl radicals recombination is 8 kJ mol-1, the enthalpy of reaction 15.5 will be equal to (121 - 8) = 113 kJ mol-1. This conclusion helps us derive other useful data. Assuming that the thermal correction to 298.15 K is small and that the solvation enthalpies of the peroxide and the acyl radicals approximately cancel, we can accept that the enthalpy of reaction 15.5 in the gas phase is equal to 113 kJ mol-1 with an estimated uncertainty of, say, 15 kJ mol-1. Therefore, as the standard enthalpy of formation of gaseous PhC(0)00(0)CPh is available (-271.7 5.2 kJ mol-1 [59]), we can derive the standard enthalpy of formation of the acyl radical Af//°[PhC(0)0, g] -79 8 kJ mol-1. This value can finally be used, together with the standard enthalpy of formation of benzoic acid in the gas phase (-294.0 2.2 kJ mol-1 [59]), to obtain the O-H bond dissociation enthalpy in PhC(0)0H DH° [PhC(0)0-H] = 433 8 kJ mol-1. [Pg.222]

Table 6.23 presents calculated barriers for the cyclization of the but-3-enyl radical [i.e. the reverse of reaction (7.2)]. This reaction is an example of an intramolecular radical addition. A number of the features observed in the barriers for the intermolecular radical additions (e.g. methyl radical addition to ethylene, Table 6.14) are also seen here. [Pg.189]

However, (TMS)3Si radicals are found to add to a variety of double bonds reversibly and therefore to isomerize alkenes [19]. An example is shown for the interconversion of ( )- to (Z)-3-hexen-l-ol and vice versa by (TMS)3Si radicals (Reaction 5.1). Figure 5.1 shows the time profile of this reaction under standard experimental conditions (AIBN, 80 °C). The equilibration of the two geometrical isomers is reached in ca 10 h, and the percentage of Z/E = 18/82 after completion corresponds to an equilibrium constant of = 4.5. The difference in the stability of the two isomers in 2-butenes, i.e., AG°( -isomer) - AG° (Z-isomer) = — 3.1kJ/mol, corresponds to K = 3.5, since... [Pg.90]

Now, just the same sort of rationalization can be applied to the radical addition, in that the more favourable secondary radical is predominantly produced. This, in turn, leads to addition of HBr in what is the anti-Markovnikov orientation. The apparent difference is because the electrophile in the ionic mechanism is a proton, and bromide then quenches the resultant cation. In the radical reaction, the attacking species is a bromine atom, and a hydrogen atom is then used to quench the radical. This is effectively a reverse sequence for the addition process but, nevertheless, the stability of the intermediate carbocation or radical is the defining feature. The terminologies Markovnikov or anti-Markovnikov orientation may be confusing and difficult to remember consider the mechanism and it all makes sense. [Pg.330]

Addition of bromine to quinolizinium bromide results in formation of the perbromide salt (13). The formation of this salt is reversible, for treatment with an easily brominated solvent like acetone regenerates the original bromide. Van der Plas and coworkers (81H(15)213) have shown that if the perbromide salt (13) is heated at 200 °C it is converted into 1-bromoquinolizinium bromide (Scheme 3). While the high temperature suggests that this may be a radical reaction, the orientation (at a carbon atom of lower positive charge, although not the 3-position predicted above) is consistent with an electrophilic mechanism. [Pg.529]

Reversible redox reactions can initiate radical chemistry without a follow-up reduction or oxidation reaction. In successful reactions of this type, the redox step that produces the radical is thermodynamically disfavored. For example, Cu(I) complexes react reversibly with alkyl hahdes to give Cu(II) hahde complexes and an alkyl radical. The alkyl radical can react in, for example, an addition reaction, and the product radical will react with the Cu(II) hahde to give a new alkyl halide. This type of reaction sequence, which has been apphed in living radical polymerizations, is in the general family of nonchain radical reactions discussed earlier. ... [Pg.143]

Among electrochemical techniques,cyclic voltammetry (CV) utilizes a small stationary electrode, typically platinum, in an unstirred solution. The oxidation products are formed near the anode the bulk of the electrolyte solution remains unchanged. The cyclic voltammogram, showing current as a function of applied potential, differentiates between one- and two-electron redox reactions. For reversible redox reactions, the peak potential reveals the half-wave potential peak potentials of nonreversible redox reactions provide qualitative comparisons. Controlled-potential electrolysis or coulometry can generate radical ions for smdy by optical or ESR spectroscopy. [Pg.210]


See other pages where Reversible radical reaction is mentioned: [Pg.340]    [Pg.155]    [Pg.12]    [Pg.470]    [Pg.599]    [Pg.169]    [Pg.271]    [Pg.916]    [Pg.101]    [Pg.344]    [Pg.574]    [Pg.703]    [Pg.2]    [Pg.164]    [Pg.166]    [Pg.86]    [Pg.423]    [Pg.205]    [Pg.149]    [Pg.63]    [Pg.249]    [Pg.184]    [Pg.7]    [Pg.460]    [Pg.275]    [Pg.340]    [Pg.142]    [Pg.189]   
See also in sourсe #XX -- [ Pg.319 ]




SEARCH



Reaction reverse

Reaction reversible

Reactions, reversing

Reversibility Reversible reactions

© 2024 chempedia.info