Nazarov cyclization reaction

This is a good article. Click here for more information.
Source: Wikipedia, the free encyclopedia.

Nazarov cyclization
Named after
Ivan Nikolaevich Nazarov
Reaction type Ring forming reaction
Identifiers
Organic Chemistry Portal nazarov-cyclization
RSC ontology ID RXNO:0000209

The Nazarov cyclization reaction (often referred to as simply the Nazarov cyclization) is a

Ivan Nikolaevich Nazarov (1906–1957) in 1941 while studying the rearrangements of allyl vinyl ketones.[1]

The Nazarov cyclization reaction
The Nazarov cyclization reaction

As originally described, the Nazarov cyclization involves the activation of a

cationic 4π-electrocyclic ring closure which forms the cyclopentenone product (See Mechanism below). As the reaction has been developed, variants involving substrates other than divinyl ketones and promoters other than Lewis acids have been subsumed under the name Nazarov cyclization provided that they follow a similar mechanistic pathway
.

The success of the Nazarov cyclization as a tool in organic synthesis stems from the utility and ubiquity of cyclopentenones as both motifs in natural products (including jasmone, the aflatoxins, and a subclass of prostaglandins) and as useful synthetic intermediates for total synthesis. The reaction has been used in several total syntheses and several reviews have been published.[2][3][4][5][6][7]

Mechanism

The mechanism of the classical Nazarov cyclization reaction was first demonstrated experimentally by

tautomerization of the enolate produces the cyclopentenone product.[8][9]

Mechanism of the classical Nazarov cyclization activated by Lewis acid catalyst. Note that α- and β-substitution are not required for the reaction to occur and that more complex α-substitution is also possible.

As noted above, variants that deviate from this template are known; what designates a Nazarov cyclization in particular is the generation of the pentadienyl

conformation, placing the vinyl groups in an appropriate orientation. The propensity of the system to enter this conformation dramatically influences reaction rate, with α-substituted substrates having an increased population of the requisite conformer due to allylic strain. Coordination of an electron donating α-substituent by the catalyst can likewise increase the reaction rate by enforcing this conformation.[2]

Relevant conformations for the Nazarov cyclization; Lewis-acid chelation in the s-cis conformer

Similarly, β-substitution directed inward restricts the s-trans conformation so severely that

stereospecific. The example below uses triethylsilane to trap the oxyallyl cation so that no elimination occurs.[2] (See Interrupted cyclizations
below)

Stereoselectivity in the Nazarov cyclization evident via reductive trapping

Along this same vein, allenyl vinyl ketones of the type studied extensively by Marcus Tius of the

University of Hawaii show dramatic rate acceleration due to the removal of β-hydrogens, obviating a large amount of steric strain in the s-cis conformer.[6]

Allene substrates for Nazarov cyclization with lowered steric interactions

Classical Nazarov cyclizations

Though cyclizations following the general template above had been observed prior to Nazarov's involvement, it was his study of the rearrangements of allyl vinyl

olefin isomerized in situ to form a divinyl ketone before ring closure to the cyclopentenone product. The reaction shown below involves an alkyne oxymercuration reaction to generate the requisite ketone.[10]

Early investigation into the Nazarov cyclization

Research involving the reaction was relatively quiet in subsequent years, until in the mid-1980s when several syntheses employing the Nazarov cyclization were published. Shown below are key steps in the syntheses of Trichodiene and Nor-Sterepolide, the latter of which is thought to proceed via an unusual

allene isomerization that generates the divinyl ketone.[11][12]

Synthesis of Trichodiene using a classical Nazarov cyclization
Synthesis of Nor-Sterepolide using a classical Nazarov cyclization

Shortcomings

The classical version of the Nazarov cyclization suffers from several drawbacks which modern variants attempt to circumvent. The first two are not evident from the mechanism alone, but are indicative of the barriers to cyclization; the last three stem from selectivity issues relating to elimination and protonation of the intermediate.[2]

  1. Strong Lewis or protic acids are typically required for the reaction (e.g.
    BF3, MeSO3H
    ). These promoters are not compatible with sensitive functional groups, limiting the substrate scope.
  2. Despite the mechanistic possibility for catalysis, multiple equivalents of the promoter are often required in order to effect the reaction. This limits the atom economy of the reaction.
  3. The elimination step is not
    regioselective; if multiple β-hydrogens are available for elimination, various products are often observed as mixtures. This is highly undesirable from an efficiency standpoint as arduous separation
    is typically required.
  4. Elimination destroys a potential stereocenter, decreasing the potential usefulness of the reaction.
  5. Protonation of the enolate is sometimes not
    stereoselective, meaning that products can be formed as mixtures of epimers
    .
Shortcomings of the Nazarov cyclization reaction

Modern variants

The shortcomings noted above limit the usefulness of the Nazarov cyclization reaction in its canonical form. However, modifications to the reaction focused on remedying its issues continue to be an active area of

catalytic in the promoter, effecting the reaction with more mild promoters to improve functional group tolerance, directing the regioselectivity of the elimination step, and improving the overall stereoselectivity
. These have been successful to varying degrees.

Additionally, modifications focused on altering the progress of the reaction, either by generating the pentadienyl cation in an unorthodox fashion or by having the oxyallyl cation "intercepted" in various ways. Furthermore,

enantioselective
variants of various kinds have been developed. The sheer volume of literature on the subject prevents a comprehensive examination of this field; key examples are given below.

Silicon-directed cyclization

The earliest efforts to improve the selectivity of the Nazarov cyclization took advantage of the

cis isomer was selected for to varying degrees.[13]

Silicon-directed Nazarov cyclization

The silicon-directed Nazarov cyclization reaction was subsequently employed in the synthesis of the natural product Silphinene, shown below. The cyclization takes place before elimination of the benzyl alcohol moiety, so that the resulting stereochemistry of the newly formed ring arises from approach of the silyl alkene anti to the ether.[10]

Synthesis of Silphinene using Nazarov cyclization

Polarization

Drawing on the substituent effects compiled over various trials of the reaction, Professor Alison Frontier of the

copper triflate and regioselective elimination. In addition, the electron withdrawing group increases the acidity of the α-proton, allowing selective formation of the trans-α-epimer via equilibration.[14]

Polarized Nazarov cyclization

It is often possible to achieve catalytic activation using a donating or withdrawing group alone, although the efficiency of the reaction (yield, reaction time, etc.) is typically lower.

Alternative cation generation

By extension, any pentadienyl cation regardless of its origin is capable of undergoing a Nazarov cyclization. There have been a large number of examples published where the requisite cation is arrived at by a variety of rearrangements.[2] One such example involves the silver catalyzed cationic ring opening of allylic dichloro cylopropanes. The silver salt facilitates loss of chloride via precipitation of insoluble silver chloride.[15]

Nazarov cyclization initiated by dichlorocyclopropane rearrangement

In the

stannane likewise generates a pentadienyl cation via ring opening of the resultant epoxide.[16]

Nazarov cyclization initiated by oxidation

Interrupted cyclization

Once the cyclization has occurred, an oxyallyl cation is formed. As discussed extensively above, the typical course for this intermediate is

diastereoselectivity.[18]

Nazarov cyclization cationic cascade

Enolate trapping with various

ligand exchange with the carbonyl ester oxygen atom before the actual Michael addition takes place to the opposite face of the R-group.[19]

Tandem Nazarov cyclization and Michael reaction

Enantioselective variants

The development of an

Diastereoselective cyclizations are also known, in which extant stereocenters direct the cyclization. Almost all of the attempts are based on the idea of torquoselectivity; selecting one direction for the vinyl groups to "rotate" in turn sets the stereochemistry as shown below.[2]

Torquoselectivity in the Nazarov cyclization

Silicon-directed Nazarov cyclizations can exhibit induced diastereoselectivity in this way. In the example below, the silyl-group acts to direct the cyclization by preventing the distant alkene from rotating "towards" it via unfavorable

steric interaction. In this way the silicon acts as a traceless auxiliary. (The starting material is not enantiopure but the retention of enantiomeric excess suggests that the auxiliary directs the cyclization.)[2]

Silicon as a chiral auxiliary for the Nazarov cyclization

Tius's allenyl substrates can exhibit axial to tetrahedral chirality transfer if enantiopure allenes are used. The example below generates a chiral diosphenpol in 64% yield and 95% enantiomeric excess.[2]

Axial to tetrahedral chirality transfer in the Nazarov cyclization of allenyl vinyl ketones

Tius has additionally developed a

trifluoroethanol as solvent.[2][20]

Synthesis of roseophilin using asymmetric Nazarov cyclization

The first chiral Lewis acid promoted asymmetric Nazarov cyclization was reported by Varinder Aggarwal and utilized copper (II) bisoxazoline ligand complexes with up to 98% ee. The enantiomeric excess was unaffected by use of 50 mol% of the copper complex but the yield was significantly decreased.[2]

Lewis-acid mediated asymmetric Nazarov cyclization

Related Reactions

Extensions of the Nazarov cyclization are generally also subsumed under the same name. For example, an α-β, γ-δ unsaturated

vinylogous Nazarov cyclizations.[22][23]

Retro-Nazarov reaction

Because they overstabilize the pentadienyl cation, β-electron donating substituents often severely impede Nazarov cyclization. Building from this, several electrocyclic ring openings of β-alkoxy cyclopentanes have been reported. These are typically referred to as retro-Nazarov cyclization reactions.[2]

Retro-Nazarov reaction

Imino-Nazarov reaction

Nitrogen analogues of the Nazarov cyclization reaction (known as imino-Nazarov cyclization reactions) have few instances; there is one example of a generalized imino-Nazarov cyclization reported (shown below),[24] and several iso-imino-Nazarov reactions in the literature.[25][26] Even these tend to suffer from poor stereoselectivity, poor yields, or narrow scope. The difficulty stems from the relative over-stabilization of the pentadienyl cation by electron donation, impeding cyclization.[27]

Imino-Nazarov cyclization

See also

References

  1. ^ Nazarov, I.N.; Zaretskaya, I.I. (1941), Izv. Akad. Nauk. SSSR, Ser. Khim: 211–224 {{citation}}: Missing or empty |title= (help)
  2. ^
  3. ^ Denmark, S.E.; Habermas, K.L.; Jones, T. K. (1994), "The Nazarov Cyclization", Organic Reactions, 45: 1–158
  4. ^
  5. ^ .