Sunteți pe pagina 1din 5

https://www2.chemistry.msu.edu/faculty/reusch/virttxtjml/pericycl.

htm
PERICYCLIC REACTIONS
An important body of chemical reactions, differing from ionic or free radical
reactions in a number of respects, has been recognized and extensively studied.
Among the characteristics shared by these reactions, three in particular set them
apart.:
1. They are relatively unaffected by solvent changes, the presence of radical
initiators or scavenging reagents, or (with some exceptions) by electrophilic
or nucleophilic catalysts.
2. They proceed by a simultaneous (concerted) series of bond breaking and
bond making events in a single kinetic step, often with high stereospecificity.
3. In agreement with 1 & 2, no ionic, free radical or other discernible
intermediates lie on the reaction path.
Since reactions of this kind often proceed by nearly simultaneous reorganization of
bonding electron pairs by way of cyclic transition states, they have been
termed pericyclic reactions. The four principle classes of pericyclic reactions are
termed: Cycloaddition, Electrocyclic, Sigmatropic, and Ene Reactions. A general
illustration of each class will be displayed by clicking on the following diagram. The
cycloaddition and ene reactions are shown in their intermolecular format.
Corresponding intramolecular reactions, which create an additional ring, are well
known.

All these reactions are potentially reversible (note the gray arrows). The reverse of a
cycloaddition is called cycloreversion and proceeds by a ring cleavage and
conversion of two sigma-bonds to two pi-bonds. The electrocyclic reaction shown
above is a ring forming process. The reverse electerocyclic ring opening reaction
proceeds by converting a sigma-bond to a pi-bond. As shown, the retro ene reaction
cleaves an unsaturated compound into two unsaturated fragments. Finally,
sigmatropic bond shifts may involve a simple migrating group, as shown in the
example above, or may take place between two pi-electron systems (e.g. the Cope
rearrangement).
The general descriptions shown above provide a basis for reaction classification, but
care must be taken to assure that a given transformation is truly concerted.
Unfortunately, this is not a trivial determination, often requiring a combination of
https://www2.chemistry.msu.edu/faculty/reusch/virttxtjml/pericycl.htm
isotope labeling and stereochemical studies to arrive at a plausible conclusion. There
is also a subtle distinction to be made between a synchronous reaction in which all
bond-making and bond-breaking events take place in unison, and a multi-
stage concerted process in which some events precede others without generating an
intermediate state.
Although some pericyclic reactions occur spontaneously, most require the
introduction of energy in the form of heat or light, with a remarkable product
dependence on the source of energy used. An appreciation of the stereoselective
structural changes these reactions promote is best achieved by inspecting some
individual examples.
1. Cycloaddition Reactions
A concerted combination of two -
electron systems to form a ring of atoms
having two new bonds and two fewer
bonds is called a cycloaddition reaction.
The number of participating -electrons
in each component is given in brackets
preceding the name, and the
reorganization of electrons may be
depicted by a cycle ofcurved arrows -
each representing the movement of a pair
of electrons. These notations are illustrated in the drawing on the right. The ring-
forming cycloaddition reaction is described by blue arrows, whereas the ring-
opening cycloreversion process is designated by red arrows. Note that the number
of curved arrows needed to show the bond reorganization is half the number
total in the brackets.

The most common cycloaddition reaction is the [4

+2

] cyclization known as
the Diels-Alder reaction. In Diels-Alder terminology the two reactants are referred to
as the diene and the dienophile. The following diagram shows two examples of
[4

+2

] cycloaddition, and in the second equation a subsequent light induced [2

+2

]
cycloaddition. In each case the diene reactant is colored blue, and the new -bonds in
the adduct are colored red. The stereospecificity of these reactions should be evident.
In the first example, the acetoxy substituents on the diene have identical E-
configurations, and they remain cis to each other in the cyclic adduct. Likewise, the
ester substituents on the dienophile have a trans-configuration which is maintained in
the adduct. The reactants in the second equation are both monocyclic, so the
cycloaddition adduct has three rings. The orientation of the quinone six-membered
ring with respect to the bicycloheptane system (colored blue) is endo, which means it
is oriented cis to the longest or more unsaturated bridge. The alternative
configuration is called exo. Since the dienophile (quinone) has two activated double
bonds, a second cycloaddition reaction is possible, provided sufficient diene is
https://www2.chemistry.msu.edu/faculty/reusch/virttxtjml/pericycl.htm
supplied. The second cycloaddition is slower than the first, so the monoadduct
shown here is easily prepared in good yield. Although this [4+2] product is stable to
further heating, it undergoes a [2+2] cycloaddition when exposed to sunlight. Note
the loss of two carbon-carbon -bonds and the formation of two -bonds (colored
red) in this transformation. Also note that the pi-subscript is often omitted from the
[m+n] notation for the majority of cycloadditions involving -electron systems.

By clicking on this diagram two more examples of cycloaddition reactions will be
displayed. Reaction 3 is an intramolecular Diels-Alder reaction. Since the diene and
dienophile are joined by a chain of atoms, the resulting [4+2] cycloaddition actually
forms two new rings, one from the cycloaddition and the other from the linking
chain. Once again the addition is stereospecific, ignoring the isopropyl substituent,
the ring fusion being cis and endo. The fourth reaction is a [6+4] cycloaddition.
2. Electrocyclic Reactions
An electrocyclic reaction is the concerted
cyclization of a conjugated -electron system by
converting one -bond to a ring forming -bond.
The reverse reaction may be called electrocyclic
ring opening. Two examples are shown on the
right. The electrocyclic ring closure is is
designated by blue arrows, and the ring opening
by red arrows. Once again, the number of curved
arrows that describe the bond reorganization is
half the total number of electrons involved in the
process.

In the first case, trans,cis,trans-2,4,6-octatriene undergoes thermal ring closure
to cis-5,6-dimethyl-1,3-cyclohexadiene. The sterospecificity of this reaction is
demonstrated by closure of the isomeric trans,cis,cis-triene to trans-5,6-dimethyl-
1,3-cyclohexadiene, as noted in the second example.
https://www2.chemistry.msu.edu/faculty/reusch/virttxtjml/pericycl.htm
By clicking on this diagram two examples of thermal electrocyclic opening of
cyclobutenes to conjugated butadienes will be displayed. This mode of reaction is
favored by relief of ring strain, and the reverse ring closure (light blue arrows) is not
normally observed. Photochemical ring closure can be effected, but the
stereospecificity is opposite to that of thermal ring opening.
3. Sigmatropic Rearrangements
Molecular rearrangements in which a -bonded atom or group, flanked by one or
more -electron systems, shifts to a new location with a corresponding
reorganization of the -bonds are called sigmatropic reactions. The total number of
-bonds and -bonds remain unchanged. These rearrangements are described by two
numbers set in brackets, which refer to the relative distance (in atoms) each end of
the -bond has moved, as illustrated by the first equation in the diagram below. The
most common atom to undergo sigmatropic shifts is hydrogen or one of its isotopes.
The second equation in the diagram shows a facile [1,5] hydrogen shift which
converts a relatively unstable allene system into a conjugated triene. Note that this
rearrangement, which involves the relocation of three pairs of bonding electrons,
may be described by three curved arrows.

By clicking on this diagram two additional examples of thermal [1,5] hydrogen shifts
will be displayed. These reactions are particularly informative in that [1,3] hydrogen
shifts are not observed. The reactant in the first equation is a deuterium labeled
1,3,5-cyclooctatriene. On heating, this compound equilibrates with its 1,3,6-triene
isomer, and the two deuterium atoms are scrambled among the four locations noted.
If [1,3] or [1,7] hydrogen shifts were taking place, the deuterium atoms would be
distributed equally among all eight carbon atoms. On prolonged heating, or at higher
temperatures these cyclooctatrienes undergo electrocyclic ring opening to 1,3,5,7-
octatetraene and reclosure to vinyl-1,3-cyclohexadienes. The second example shows
another [1,5] hydrogen shift, from the proximal methyl group to the carbonyl oxygen
atom. The resulting dienol rapidly exchanges OH for OD before the [1,5] shift
reverses. In this manner the reactive methyl is soon converted to CD
3
. Since
hydrogens alpha to a carbonyl group are known to undergo acid or base catalyzed
exchange by way of enol intermediates, we might expect the '-CH
2
group to
https://www2.chemistry.msu.edu/faculty/reusch/virttxtjml/pericycl.htm
exchange as well. However, if care is taken to remove potential acid or base
catalysts, the thermal [1,3] shift necessary for the exchange is found to be very slow.
The [3,3] sigmatropic rearrangement of 1,5-dienes or allyl vinyl ethers, known
respectively as the Cope and Claisen rearrangements, are among the most
commonly used sigmatropic reactions. Three examples of the Cope rearrangement
are shown in the following diagram. Reactions 1 and 2 (top row) demonstrate the
stereospecificity of this reaction. The light blue -bond joins two allyl groups,
oriented so their ends are near each other. Since each allyl segment is the locus of a
[1,3] shift, the overall reaction is classified as a [3,3] rearrangement. The three pink
colored curved arrows describe the redistribution of three bonding electron pairs in
the course of this reversible rearrangement. The diene reactant in the third reaction is
drawn in an extended conformation. This molecule must assume a coiled
conformation (as above) before the [3,3] rearrangement can take place. The product
of this rearrangement is an enol which immediately tautomerizes to its keto form.
Such variants are termed the oxy-Coperearrangement, and are useful because the
reverse rearrangement is blocked by rapid ketonization. If the hydroxyl substituent is
converted to an alkoxide salt, the activation energy of the rearrangement is lowered
significantly.
The degenerate or self-replicating Cope rearrangement has been a fascinating subject
of research. For examples .

Two examples of the Claisen Rearrangement may be seen by clicking on the above
diagram. Reaction 4. is the classic rearrangement of an allyl phenyl ether to an ortho-
allyl phenol. The methyl substituent on the allyl moiety serves to demonstrate the
bonding shift at that site. The initial cyclohexadienone product immediately
tautomerizes to a phenol, regaining the stability of the aromatic ring. Reaction 5 is an
aliphatic analog in which a vinyl group replaces the aromatic ring. In both cases
three pairs of bonding electrons undergo a reorganization. By clicking on the above
diagram a second time two examples of [2,3] sigmatropic rearrangements will be
displayed. The allylic sulfoxide in reaction 6 rearranges reversibly to a less stable
sulfenate ester. The weak S-O bond may be reductively cleaved by trimethyl
phosphite to an allylic alcohol and a thiol (not shown). Reaction 7 shows a similar
rearrangement of a sulfur ylide to a cyclic sulfide. The [2,3]-Wittig rearrangement is
yet another example.

S-ar putea să vă placă și