Big Chemical Encyclopedia

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

Articles Figures Tables About

Diels-Alder reactions thermodynamically

Stereochemistry of Diels-Alder Reactions. Thermodynamic vs. Kinetic Control... [Pg.276]

C, b.p. 170 C (decomp.), has a characteristic odour. It is the Diels-Alder product of cyclopentadiene reacting with itself, the exo-form being formed most rapidly but the endo-form is thermodynamically favoured. At temperatures above ISO C a retro-Diels-Alder reaction occurs and cyclopentadiene monomer is regenerated see diene reactions. [Pg.135]

Unfortunately, the number of mechanistic studies in this field stands in no proportion to its versatility" . Thermodynamic analysis revealed that the beneficial effect of Lewis-acids on the rate of the Diels-Alder reaction can be primarily ascribed to a reduction of the enthalpy of activation ( AAH = 30-50 kJ/mole) leaving the activation entropy essentially unchanged (TAAS = 0-10 kJ/mol)" . Solvent effects on Lewis-acid catalysed Diels-Alder reactions have received very little attention. A change in solvent affects mainly the coordination step rather than the actual Diels-Alder reaction. Donating solvents severely impede catalysis . This observation justifies the widespread use of inert solvents such as dichloromethane and chloroform for synthetic applications of Lewis-acid catalysed Diels-Alder reactions. [Pg.13]

For a discussion of the mechanistic course of the reaction, many aspects have to be taken into account. The cisoid conformation of the diene 1, which is in equilibrium with the thermodynamically more favored transoid conformation, is a prerequisite for the cycloaddition step. Favored by a fixed cisoid geometry are those substrates where the diene is fitted into a ring, e.g. cyclopentadiene 5. This particular compound is so reactive that it dimerizes easily at room temperature by undergoing a Diels-Alder reaction ... [Pg.89]

The Diels-Alder reaction of a diene with a substituted olefinic dienophile, e.g. 2, 4, 8, or 12, can go through two geometrically different transition states. With a diene that bears a substituent as a stereochemical marker (any substituent other than hydrogen deuterium will suffice ) at C-1 (e.g. 11a) or substituents at C-1 and C-4 (e.g. 5, 6, 7), the two different transition states lead to diastereomeric products, which differ in the relative configuration at the stereogenic centers connected by the newly formed cr-bonds. The respective transition state as well as the resulting product is termed with the prefix endo or exo. For example, when cyclopentadiene 5 is treated with acrylic acid 15, the cw fo-product 16 and the exo-product 17 can be formed. Formation of the cw fo-product 16 is kinetically favored by secondary orbital interactions (endo rule or Alder rule) Under kinetically controlled conditions it is the major product, and the thermodynamically more stable cxo-product 17 is formed in minor amounts only. [Pg.91]

C to the thermodynamically more stable exo adduct through a retro Diels-Alder reaction followed by re-addition (Scheme 1.10). [Pg.15]

The reactivity of heterocyclic dienes is determined by the nature and number of heteroatoms and, in the case of heteroaromatic compounds, also by the aromatic character. Furans undergo Diels-Alder reactions with strong dienophiles and generally afford cxo-cycloadducts which are thermodynamically more stable than the kinetically favoured c z/o-adducts. [Pg.40]

Diels-Alder reactions of 1-azadienes are less thermodynamically favorable [92] than the all-carbon analogs because of the stronger carbon nitrogen 7r-bond which is broken during the Diels-Alder reaction. [Pg.66]

As an approach to biomimetic catalysis, Sanders and colleagues [67] synthesized a series of 1,1,2-linked cyclic porphyrin trimers that allow the stereo- and regiochemistry of the Diels-Alder reaction of 84 and 85 within the molecular cavity to be controlled, thereby producing prevalently or exclusively the endo 86 or the exo 87 adduct. Two examples are illustrated in Scheme 4.18. At 30 °C and in the absence of 88, the reaction furnishes a mixture of diastereoisomers, while the addition of one equivalent of trimer 88 accelerates the reaction 1000-fold and the thermodynamically more stable exo adduct 87 is the sole detectable product. [Pg.172]

The different ratios of 52/53 produced by cycloadditions performed at atmospheric and high pressure, and the forma tion of the unusual trans adducts 53, have been explained by the facts that (i) Diels-Alder reactions under atmospheric pressure are thermodynamically controlled, and (ii) the anti-endo adducts 52 are converted into the short-lived syn-endo adducts 54 which tautomerize (via a dienol or its aluminum complexes) to 53. The formation of trans compounds 53 by induced post-cycloaddition isomerization makes the method more flexible and therefore more useful in organic synthesis. [Pg.212]

An interesting phenomenon has been observed in the high pressure Diels-Alder reactions of the l-oxa[4.4.4]propella-5,7-diene (117) with 1,4-naphthoquinone, maleic anhydride and N-phenylmaleimide, where the diene 117 undergoes a rearrangement to the diene isomer 118 which, although thermodynamically less favored, exhibits a greater reactivity [40]. The reactivities of the three dienophiles differed since maleic anhydride and N-phenylmaleimide reacted only in the presence of diisopropylethylamine (DIEA) and camphorsulfonic acid (CSA), respectively (Scheme 5.15). The distribution of the adduct pairs shows that the oxygen atom does not exert a consistent oriental dominance on TT-facial selectivity. [Pg.224]

Lubineau and coworkers [18] have shown that glyoxal 8 (Ri = R2 = H), glyoxylic acid 8 (Ri = H, R2 = OH), pyruvic acid 8 (Ri = Me, R2 = OH) and pyruvaldehyde 8 (Ri = H, R2 = Me) give Diels-Alder reactions in water with poor reactive dienes, although these dienophiles are, for the most part, in the hydrated form. Scheme 6.6 illustrates the reactions with (E)-1,3-dimethyl-butadiene. The reaction yields are generally good and the ratio of adducts 9 and 10 reflects the thermodynamic control of the reaction. In organic solvent, the reaction is kinetically controlled and the diastereoselectivity is reversed. [Pg.258]

Despite the fact that the exo adduct is likely to be the more stable of the two thermodynamically, it is often (though not universally) found in Diels-Alder reactions that the endo adduct is the major, if not the sole, product. To explain this, it has been suggested that in endo addition stabilisation of the T.S. can occur (and the rate of reaction thereby speeds up) through secondary interaction of those lobes of the HOMO in, e.g. (32) and of the LUMO in (33) that are not themselves involved directly in bond-formation, provided these are of the same phase. Such interaction would not, of course, be possible in the T.S. for exo addition because the relevant sets of centres in (32) and (33) will now be too far apart from each other the endo adduct is thus the kinetically controlled product. It is significant in this connection that the relative proportion of exo... [Pg.350]

Domingo and coworkers [11] have contributed an important theoretical input for the understanding of domino reactions. An interesting example is the domino Diels-Alder reaction of 4-33 and 4-34, in which the products 4-37 and 4-38 could be formed via 4-35 and 4-36, respectively (Scheme 4.7). Visnick and Battiste [12] had shown that, at room temperature, only cycloadduct 4-37 is formed, whereas with heat 4-38 is obtained quantitatively. This is in line with the calculations showing that TS5 is higher in energy than TS4 (74.5 and 55.3 kj mol"1, respectively) on the other hand, cycloadduct 4-38 is more stable (-92.9 kj mol"1) than cycloadduct 4-37 (-78.7 kj mol"1), which explains the formation of 4-38 under thermodynamic control. Calculations have also been performed for the bisfuran system 4-28a [13]. [Pg.285]

Cycloadditions selectively afford the adducts on the 6,6-ring junctions [65], and the products occasionally undergo a facile retro-Diels-Alder reaction as a consequence of the low thermodynamic stability of the adduct. Very stable Diels-Alder cycloadducts have, however, been prepared by use of different substituted o-quinodimethanes, probably because of stabilization by aromatization of the resulting adducts [66],... [Pg.309]

The term Diels-Alder reaction in a general sense refers to the reaction between a diene and a dienophile. Retro Diels-Alder reaction is a process that, under certain conditions, produces diene and olefin or a compound containing a C=C bond. The application of flash vacuum pyrolysis to effect the retro Diels-Alder reaction, as shown in Schemes 5-46 and 5-47, has become the standard procedure since the introduction of the method by Stork et al.74 in the 1970s. Therefore, alkenes that are difficult to access by conventional methods may be obtained via retro Diels-Alder reactions.75 In particular, this reaction allows the preparation of thermodynamically less stable compounds such as 4,5-dialkyl cyclopenta-2-en-one. In this case, the alkene functional group can be regarded as being protected by cyclopentadiene (as shown in 154 or 157), which, after subsequent reaction, can easily be removed through quick pyrolysis. [Pg.306]

Hetero Diels-Alder reactions are very useful for constructing heterocyclic compounds, and many important chiral molecules have thus been synthesized. Although the retro Diels-Alder reaction does not itself involve the asymmetric formation of chiral centers, this reaction can still be used as an important tool in organic synthesis, especially in the synthesis of some thermodynamically less stable compounds. The temporarily formed Diels-Alder adduct can be considered as a protected active olefin moiety. Cyclopentadiene dimer was initially used, but it proved difficult to carry out the pyrrolytic process. Pentamethyl cyclopentadiene was then used, and it was found that a retro Diels-Alder reaction could easily be carried out under mild conditions. [Pg.322]

Actually one orientation predominates (called high regioselectivity) and only one diastereoisomer is produced (called high stereoselectivity). The Diels-Alder reaction is reversible and may be carried out under thermodynamic or kinetic controlled conditions. [Pg.46]

Scheme 6.47 Calculated thermodynamic and kinetic data for the dehydro Diels—Alder reaction of butenyne with acetylene and non-l-ene-3,8-diyne. Scheme 6.47 Calculated thermodynamic and kinetic data for the dehydro Diels—Alder reaction of butenyne with acetylene and non-l-ene-3,8-diyne.
Dicyclopentadiene is the Diels-Alder reaction dimer of cyclopentadiene. It is the thermodynamically stable form of cyclopentadiene at room temperature, and is also a byproduct in the olefin cracking process. Industrially, it is isolated by distillation, and currently is readily available in North America. [Pg.704]

Marchand and coworkers102 reported a difference in site selectivity between the thermodynamically and kinetically controlled Diels-Alder reactions of cyclopentadiene with 2,3-dicyano-p-benzoquinone (126) (equation 37). Under kinetic conditions, the more reactive double bond of 126 reacted with cyclopentadiene affording 127, whereas the less substituted double bond reacted under thermodynamic conditions affording 128. Both reactions proceeded with complete endo selectivity. These findings were in agreement with ab initio HF/3-21G calculations. [Pg.361]

While many observations are well understood, e.g. those dealing with the reaction rate or with the selectivity, there are some factors which cannot be generalized. Many transformations of particular reactants or under unusual reaction conditions led to unexpected results. There are often singular explanations for such reactions but no overall concept. For instance, computations on Diels-Alder transition structures and thermodynamics of retro-Diels-Alder reactions confirmed that the activation volume of these [4 + 2]-cycloadditions is negative80. This result, pointing to the compact character of the transition structure, is used to explain the dependence of reactivity and selectivity on internal as well as external pressure81-83. These effects are only observed at relatively high external pressures (Table 5). [Pg.1041]

The pseudothermodynamic analysis of solvent elfects in 1-PrOH-water mixtures over the whole composition range (shown in Figure 7.3) depicts a combination of thermodynamic transfer parameters for diene and dienophile with isobaric activation parameters that allows for a distinction between solvent elfects on reactants (initial state) and on the activated complex. The results clearly indicate that the aqueous rate accelerations are heavily dominated by initial-state solvation effects. It can be concluded that for Diels-Alder reactions in water the causes of the acceleration involve stabilization of the activated complex by enforced hydrophobic interactions and by hydrogen bonding to water (Table 7.1, Figure 7.4). °... [Pg.164]

Approaches to oseltamivir phosphate (1) that were independent of ( )-shikimic acid as the raw material were also evaluated. The furan-ethyl acrylate Diels-Alder approach is shown in Scheme 7.8 (Abrecht et al., 2001, 2004). The zinc-catalyzed Diels-Alder reaction between furan and ethyl acrylate was heated at 50°C for 72 h to provide a 9 1 mixture favoring exo-isomer rac-43 over the enJo-isomer. The enJo-isomer was kinetically preferred, but with increased reaction times an equilibrium ratio of 9 1 was achieved favoring the thermodynamically preferred exo-isomer rac-43. The optical resolution of rac-43 was achieved via enantioselective ester hydrolysis using Chirazyme L-2 to give (—)-43 in 97%... [Pg.105]

Theoretical work on the gas-phase hetero-Diels-Alder reaction of A -sulfinyl dienophiles was used to study both endo- and o-modes of cycloaddition for both (E)-29 and (Z)-30 dienophiles at the B3LYP/6-31G level (Scheme 2) <2000JOC3997>. In summary, these calculations have predicted that (1) the A -sulfinyl dienophiles prefer the (Z)-30 orientation over (E)-29 stereochemistry by 5-7 kcalmoP, (2) the transition state is concerted but nonsynchronous, and (3) an lYo-transition state with diene 31 is favored over the fvo-approach both kinetically and thermodynamically. [Pg.517]


See other pages where Diels-Alder reactions thermodynamically is mentioned: [Pg.271]    [Pg.306]    [Pg.271]    [Pg.306]    [Pg.318]    [Pg.438]    [Pg.160]    [Pg.640]    [Pg.304]    [Pg.178]    [Pg.82]    [Pg.160]    [Pg.80]    [Pg.373]    [Pg.56]    [Pg.477]    [Pg.22]    [Pg.255]    [Pg.117]    [Pg.302]    [Pg.361]    [Pg.715]   


SEARCH



Diels-Alder reaction thermodynamic control

Reactions thermodynamics

Stereochemistry of Diels-Alder Reactions Thermodynamic vs. Kinetic Control

Thermodynamic reactions

© 2024 chempedia.info