Vous êtes sur la page 1sur 33

Polytechnic University of Bucharest

Faculty of Engineering in Foreign Languages


Chemical Engineering Division



Student Communications
May 2003 Edition







S
S
t
t
e
e
r
r
e
e
o
o
c
c
h
h
e
e
m
m
i
i
c
c
a
a
l
l

A
A
s
s
p
p
e
e
c
c
t
t
s
s
o
o
f
f

P
P
e
e
r
r
i
i
c
c
y
y
c
c
l
l
i
i
c
c
R
R
e
e
a
a
c
c
t
t
i
i
o
o
n
n
s
s




Eugen S. Andreiadis










Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

2

1. Introduction


In organic chemistry there exist a number of reactions known as pericyclic

reactions, the
stereochemistry of which depends on the symmetry of the interacting molecular orbitals
(MOs) and not on the overall symmetry of the molecules. These reactions are situated
apart from the more common heterolytic and homolytic transformations since they
dont involve intermediates of any kind, ionic, radicalic or carbenic. The bond-breaking
and bond-formation process occur (not necessarily synchronous

) through a cyclic
transition state (TS) accounting for a so-called concert mechanism, the result of which is
the transposition of bonds along a closed curve. They are further characterised by the
fact that their rate is unaffected by the solvent polarity (unless the reactants are charged
see 1,3 dipolar cycloadditions), the presence of radical initiators (or inhibitors) or
catalysts (exceptions are some acid-catalysed Diels-Alder reactions), and that they take
place thermally or photochemically. The reactions have high (sometimes even total)
stereoselectivity and stereospecificity under kinetically controlled conditions and are of
special interest in organic synthesis. A classical example is the Diels-Alder reaction
between a diene and a substituted alkene (dienophile) in which the 4-electrons of the
diene and the 2-electrons of the alkene reorganise thermally through a set of
interacting MOs to the 4 and 2-electrons of a cyclohexene (reaction (1)).

Y
X
Y
X X
Y


Woodward and Hoffmann were the first to deduce a series of rules
19-22
to explain
and predict the stereochemistry of the products obtained and the reaction conditions
necessary for a certain transformation to take place. The rules are based on the principle
of conservation of orbital symmetry which may be briefly stated as follows: the
transformations in which the symmetry of the MO is conserved (i.e. the orbitals remain
in phase and thus maintain a degree of bonding during the process) involve a relatively
low energy TS and are called symmetry allowed. On the other hand, in the transformation
in which the symmetry of the orbitals is destroyed by bringing one or more orbitals out
of phase, the energy of the TS becomes very high due to an antibonding interaction and
the reaction is symmetry forbidden. Slight perturbations of the molecular symmetry
caused by a substituent (e.g. CH3) are generally ignored since the mechanism of the
process remains the same. Another important rule referring to the reaction conditions
states that a thermally allowed transformation is forbidden photochemically, and vice
versa, a photochemically allowed process is forbidden thermally, moreover thermal and

from Gr. Perikyklos (peri =around; kyklos =ring).

each bond undergoing change need not necessarily have been made or broken to the same extent by the time the
TS has been reached.

(1)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

3
photochemical reactions give opposite stereochemistry. However, it is important to note
that a forbidden reaction may still take place if nothing easier is available. Three cases
are possible
4
, with the choice depending on various factors:

a) a nonconcerted reaction may occur to form a discrete intermediate;
b) enough energy may be added to force the reaction to produce via the symmetry forbidden
pathway;
c) the reaction may follow a symmetry-allowed pathway to give an excited state of the
product.

What is meant by forbidden is that the interaction of the orbitals presents an
energy barrier that the allowed reactions do not have. It has turned out that, in many
cases, this energy is quite high examples of allowed transformations are abundant, but
forbidden reactions are few and far between. Equally, the statement that a reaction is
symmetry-allowed does not necessarily guarantee that it will proceed readily in
practice: the attainment of the required geometry in the TS could well be inhibited by
steric factors, such as the size of a ring or the presence of bulk substituents.

Three different models with different degrees of sophistication can now be used
to explain the results of pericyclic reactions: (a) the method based on the principle of
orbital symmetry conservation throughout the reaction (Woodward-Hoffmann); (b) the
method based on aromatic stabilisation of the TS according to Hckels MO theory
(Zimmerman and Dewar); and (c) the frontier molecular orbital (FMO) method
(Woodward-Hoffman, Fukui) which, although the simplest, is capable of explaining the
stereochemistry of all pericyclic reactions and is therefore adopted in the present text
5
.

In this model, one is interested only in the interaction between the highest
occupied molecular orbital (HOMO) of one reactant and the lowest unoccupied molecular
orbital (LUMO) of the other. When two molecules (or appropriate segments thereof)
approach each other in a reaction, pairs of filled MOs which are close in energy interact
to give pairs of hybrid MOs, one bonding and the other antibonding. The energy gained
by a bonding orbital is always slightly less than the energy lost by the antibonding one
so that the energy of the system increases slightly (four electrons are to occupy the two
hybrid MOs) and the cumulative effect constitutes the major part of the activation
energy of the reaction. At the same time, the HOMO of one molecule interacts with the
LUMO of the other but since there are only two electrons (usually), they are
accommodated in the bonding hybrid orbital lowering the activation energy to an
appreciable extent. A third factor, namely, coulombic interaction, has also to be
considered when dealing with charged reacting species. The three combined effects
account for the activation energy of reaction. In order to have appreciable interaction
between the HOMOs and the LUMOs (which is the major consideration in pericyclic
reactions), they must be of comparable energies and above all must belong to the same
symmetry type.

Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

4
Theoretically, all pericyclic transformations may be regarded as cycloaddition
reactions or their retrogressions, but they are usually classified into several categories:
electrocyclic, cycloaddition, sigmatropic, cheletropic, and group transfer reactions.
Following, only the first two of these divisions will be analysed and emphasis will be
put on their stereochemical aspects.


2. Electrocyclic reactions


An electrocyclic reaction is one in which the two terminals of a linear conjugated -
electron system are joined through a single bond, the net change being the conversion of
a -bond into a -bond, or a reversion of this process, i.e. the conversion of a -bond
into a -bond with concomitant ring-opening
1
. Such a transformation can be
represented generally as below (Figure 1), where n indicates the number of trigonal
carbon atoms in the conjugated polyene and also the number of -electrons.

n
CH
n-2
CH


Figure 1. Schematical representation of an electrocyclic reaction

2.1. Frontier Molecular Orbital Approach

In electrocyclic reactions, only one reactant (i.e. a polyene or in the reverse
reaction, a cycloalkene) is involved and thus only the HOMO needs to be considered. If
we take the example of the interconversion of 1,3,5-hexatriene and cyclohexadiene (2),
we can represent the HOMO of the triene as in Figure 2.

To form a C-C bond on cyclisation, the p orbitals on the terminal atoms have to
rehybridise to sp
3
orbitals, and each rotate through 90
0
to allow of their potential
overlap.
conrotatory

HOMO
(
3
)
HOMO
(
3
)
antibonding
interaction
bonding
interaction
disrotatory


Figure 2. Conrotatory and disrotatory thermal cyclisation of hexatriene
(2)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

5

This rotation could happen in the same direction (conrotatory), or in opposite directions
(disrotatory). Conrotatory movement results in the overlap of sp
3
orbitals of opposite
phase, leading to an antibonding situation, while disrotatory movement results in the
overlap of sp
3
orbitals of the same phase, leading to a bonding interaction and thus to
the formation of cyclohexadiene. In thermal electrocyclic reactions of a 6 -electron
system, therefore, only disrotatory motion is allowed and the stereochemistry follows
accordingly (see subchapter 2.2).

On photochemical ring-closure, irradiation results in the promotion of an
electron into the orbital of next higher energy level, i.e. 3 4 and the ground state
LUMO (4) thus now becomes the HOMO
*
(Figure 3). It is now conrotatory motion that
results in the overlapping of sp
3
orbitals of the same phase and a different
stereochemistry follows.
conrotatory
h
HOMO*
(
4
)
bonding
interaction
antibonding
interaction
disrotatory
HOMO*
(
4
) h

Figure 3. Conrotatory and disrotatory photochemical cyclisation of hexatriene

If we consider now the case of a 4-electron system, for example the
interconversion of 1,3-butadiene and cyclobutene (3),

it is this time conrotatory movement that results in a bonding situation for the ground-
state diene (Figure 4). In the case of photochemical interconversion (which tends to lie
over in favour of the cyclobutene), irradiation of the diene will result in the promotion
of an electron into the orbital of next higher energy level (i.e. 2 3) and the ground
state LUMO becomes the HOMO
*
. It is disrotatory motion that results now in a bonding
interaction and the formation of cyclobutene.
conrotatory

disrotatory
HOMO
(
2
)
HOMO
*
(
3
)
h

Figure 4. Thermal and photochemical cyclisation of butadiene

(3)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

6
This difference in behaviour derives from the way in which the phases of the
MOs at the termini of the conjugated electronic system are arranged, i.e. their
symmetry. The same phase demand disrotatory movement for bond-making/bond-
breaking to occur, while opposite phases demand conrotatory movement for bond-
making/bond-breaking to occur. The selection rules gathered in Table 1, derived by
Woodward and Hoffmann
19
, may be stated for all electrocyclic reactions.

Table 1 Woodward Hoffmann rules for electrocyclic reactions

n (no. of -
electrons)
Condition
s
Geometry
conrotator
y 4 k
h disrotatory
disrotatory
4 k + 2 h conrotator
y

Electrocyclic reactions, particularly the ring opening of unsaturated cycloalkenes
to polyenes, may be regarded as cycloadditions in which a bond and bond (or a
conjugated system) constitute the two components in addition reaction
1
. Some notations
that are used in cycloadditions reactions may be defined here to indicate the mode of
addition. If a component undergoes addition (i.e. forms bonds) on the same face, it is
called a suprafacial component, while if a component undergoes addition on opposite
faces, it is called antarafacial component. The two modes of addition are known as
suprafacial and antarafacial respectively. Following this definition, conrotation involves
an antarafacial and disrotation a suprafacial interaction between the two termini of the
reacting species in electrocyclic reactions, as can be clearly seen from Figure 5.

suprafacial interaction
disrotatory motion
antarafacial interaction
conrotatory motion


Figure 5. Suprafacial and antarafacial interactions in electrocyclic reactions

For example, in the thermal conrotatory ring opening of cyclobutene (Figure 6),
the -HOMO of the component interacts with the -LUMO of the component, the
former behaving as a suprafacial component and the latter as an antarafacial one. The
reaction is designated [2s+2a] (where s and a stand for supra- and antarafacial
respectively). Alternatively, the same reaction may be considered in terms of interaction
between the -LUMO (i.e. *) and the -HOMO and thus denoted as [2a+2s].

Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

7
[

2
s
+

2
a
]

-LUMO
-HOMO
a
s
[

2
a
+

2
s
]

-HOMO
-LUMO
s
a


Figure 6. Thermal ring opening of cyclobutene as a cycloaddition reaction.
Suprafacial / antarafacial nomenclature

2.2. Stereochemical aspects

The stereochemistry of an electrocyclic reaction derives from the conrotatory or
disrotatory mode of ring closure (or ring opening) as permitted by the system under
consideration and the conditions of reaction (thermal or photochemical). For each mode
of reaction, there are two distinct possibilities: the disrotatory process can occur inwards
or outwards, and similarly the conrotatory process can occur towards left or towards right
(Figure 7).










Figure 7. Different possibilities of disrotatory
and conrotatory movement and their
stereochemical outcome


In cases where the two possibilities of a mode give the same product due to an
inherent symmetry of the system or one possibility is forbidden by molecular geometry
or by steric interaction in the TS, almost total diastereoselectivity results.

2.2.1. Four-member rings. The dienes and the cyclobutenes can be interconverted
through electrocyclic transformations. However, the cyclobutenes are
thermodynamically less stable than their partners by some 50 kJ mol
1
, and so the ring
closure cannot practically occur thermally due to reversibility of the process. On the
other hand, the dienes are absorbing light at a much higher wavelength than that of
ordinary alkenes, and it is thus possible for the photochemical cyclisation of dienes to
occur irreversibly. The possible outcomes of such processes, considering the
stereochemistry of the products involved, are represented in Figure 8. (For uniformity,
all disrotatory motions are inward and all conrotatory motions are toward right.)
R R
disrotatory
inwards
R R
R R
R R
R R
disrotatory
outwards
conrotatory
twrds right
conrotatory
twrd left
R R
R
R
R
R
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

8

h
disrot
h
disrot
h
disrot
conrot
conrot
conrot



Figure 8. Stereochemistry of electrocyclic reactions: four-member rings

Some interesting transformations are depicted below (reactions (4) and (5)).

O
O
h
disrot O
H
H
O
conrot

O
O
impossible
1 2 3

H H
conrot
<100
o
6
4
h
disrot
H H
5


The lactone 1 undergoes irreversible photochemical cyclisation to give 2. The
reverse reaction cannot occur through the excited pathway because radiation of
convenient wavelength is unavailable, and neither can occur through the thermal
pathway since the conrotatory movement required leads to the unlikely looking lactone
3 containing a trans double bond. In reaction (5), the cyclobutene 5 obtained similarly
through an electrocyclic process from diene 4 is stable at temperatures above 250C
because the thermal conrotatory ring opening would also give a geometrically-
impossible trans-cyclohexenic ring. In comparison, the trans-substituted cyclobutene 6
readily undergoes thermal bond cleavage to yield the diene 4.

Dewar benzene 7 was obtained
23
through the following reaction (6), using
electrocyclic ring closure.

O
O
O
h
O
O
O
Pb(OAc)
4
7


The direct formation of Dewar benzene derivatives by irradiating the corresponding
aromatic derivatives
24
can be accomplished, in spite of the high conjugation energy of
(4)
(5)
(6)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

9
the latter, for compounds substituted with bulk groups for which the transformation in
Dewar benzene brings a steric relief (reaction (7)).

tBu
tBu
tBu
h
tBu
tBu
tBu


It is interesting to observe that, since benzene is with ~250 kJ mol
1
more stable
than Dewar benzene, the conversion should be facile. However, an activation energy of 96 kJ
mol
1
is required for such a transformation to take place
25
. The reason, as in the case of
compound 5, is the fact that the allowed thermal conrotatory ring opening produces a
cyclohexatriene with a trans double bond (reaction (8)).

H
H

H H
not
H
H
7


As compared to the electrocyclic ring closure, the ring opening process of four-
member rings has no thermodynamic restrictions, since it goes towards the more stable
product. Thus, it can take place thermally as well as photochemically. The thermal
opening of cis-3,4-dimethylcyclobutene 8 by either of the two possible conrotatory
motions gives Z,E-1,4-dimethylbutadiene, while the photochemical process gives the
E,E-isomer only through a single disrotatory motion in which the substituents move
outwards, because the inward movement implies excessive steric repulsion (reaction
(9)).

H
H
Me
Me
h
Me
H
H
Me

Me
H
Me
H
E, E
Z, E
8
or


Similarly, the trans-cyclobutene 9 gives only E,E and not Z,Z-butadienes on thermal ring
opening, as shown in reaction (10). However, the outward (or inward) movement of a
substituent is not controlled only by steric factors, but also by electronic effects. Thus an
electro-releasing substituent such as F shows a high affinity for outward rotation
26
, as
illustrated by reaction (11) in which the fluorinated cyclobutene 10 undergoes ring
cleavage to yield preferentially the diene 11.

H
H
Me
Me
H
H
Me
Me

Me
Me
H
H
E, E Z, Z
9

x


(7)
(8)
(9)
(10)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

10
F
F
CF
3
F
CF
3
F
F
F

F
CF
3
F
F
10

x
11


If we replace the methyl groups of cyclobutene 8 with a polymethylene chain as
in bicycloalkene 12, conrotatory ring opening should give rise to a cyclodiene containing
one E double bond (see reaction (12)).
H
H
( )
n
conrot

H
H
( )
n
12


However, this would happen only if the chain is long enough. If it has less than 8
members (as in 13), ring opening takes place only at high temperature (13) through a
non-concerted mechanism which involves the disrotatory movement as the only one
possible geometrically.

H
H
Me
Me
disrot

Me
Me
13


2.2.2. Six-member rings. In hexatrienes-cyclohexadienes interconversion, the
reactions occur thermally disrotatory and photochemically conrotatory. The various
transformations of different isomers of 1,6-dimethyl hexatriene and their stereochemical
outcome are shown in Figure 9.


h
conrot
disrot

trans cis
cis
h
conrot
disrot

trans


Figure 9. Stereochemistry of electrocyclic reactions: six-membered rings

These reactions are somewhat different than their four-membered ring
counterparts. First, the difference in thermodynamic stability between the triene and the
cyclodiene is lower than in the previous case, and thus thermal process is an
equilibrium one. Moreover, both the triene and the cyclodiene absorb light at similar
wavelengths, making it impossible to shift the equilibrium eitherway (at least in usual
experimental conditions). However, the triene usually predominates in the reaction
(11)
(12)
(13)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

11
mixture, both for the thermal as well as for the photochemical transformations. Second,
the polyene must be so disposed as to bring the two termini within reaction distance,
which is equivalent to all-cis geometry except at the marginal double-bonds. However,
in the photochemical conditions, the cis-trans equilibrium occurs easily, and thus the
geometry of the double bonds is of less importance.

It can be observed from the Figure 9 that the disrotatory ring closure of any
polyene brings the two outer and the two inner substituents at the termini close to each
other, i.e. to the same side (cis) of the newly formed bond, while the conrotatory ring
closure does the opposite.

The steric repulsion mentioned in the reaction (9) also applies here, forcing the
disrotatory thermal ring opening of the cis isomer of 5,6-dimethyl-cyclohexadiene to
follow only one pathway with both the methyl substituents moving outwards and
forming two marginal trans double bonds. But when the two substituents are apart of a
ring as in the case of bicyclic diene 14, this is no longer the case and the thermal ring
opening takes place inwards, giving an all-cis cyclotriene 15 (reaction (14)). This
happens because there is no longer any steric hindrance and two trans double bonds
cannot be accommodated in a ring unless this is a large one. In the interconversion of
norcaradiene (14, n=1) and cycloheptatriene (15, n=1), the equilibrium is shifted towards
right
27
. For n>1, interconversious among 14 and other compounds, such as 16 and 17, are
possible.

(CH
2
)
n
(CH
2
)
n
(CH
2
)
n
h

(CH
2
)
n
h
14 15
16 17



The thermal and photochemical transformations of steroidic derivatives further
illustrate the stereochemistry of these reactions. For example, consider the following
scheme (15) of the processes taking place upon irradiation and heating of ergosterol 18,

(14)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

12
Me
Me
HO
R
3
2
4
5
6
Me
Me 1
HO
R
7
Me
HO
R
Me
Me
HO
R
Me
R
HO
Me
Me
Me
HO
R
Me
Me
HO
R
h
con
con
h
h
con
h
con

[1,7]
dis
18
19
20
24
23
22
21
(lumisterol)
(precalciferol)
(ergosterol)
(tachisterol)
(isopyrocalciferol) (pyrocalciferol)
(vit D
2
)




when a mixture containing ergosterol, precalciferol 19, vitamin D2 20, lumisterol 21 and
tachisterol 22 is obtained
28
. No matter the starting product, the same equilibrium is
reached, with ergosterol in the smallest yield. What is interesting is that the
photochemical ring closure of 19 to give ergosterol occurs reversibly, while the same
reaction of 22 to yield 21 is irreversible. The reversible process is to be expected, but 22
having a trans double bond cannot react directly, being necessary for it to be converted
first to an excited state in which the rotation around the trans double bond can be
accomplished. This rotation occurs in the direction depicted in the scheme, leading to
lumisterol in order to avoid the steric repulsion between the two methyl groups. Ring
opening of 21 is also a conrotatory process in photochemical conditions and the only
geometrically-allowed pathway gives the precalciferol 19. This can be thermally
converted through a [1,7] sigmatropic transformation to vitamin D2 20 which, on its
turn, can suffer thermal cyclisation to give 23 and 24 in molar ratio 1:1.

Eight-member and higher even-member rings can be treated in a similar fashion.
An interesting example is given in (16).

Me
Me
2 H
2
Me
Me
conrot

H
H
Me
Me
disrot

Me
Me
25
26 27


During the syn-hydrogenation of biacetilene 25, surprisingly, the bicycle-
octadiene 27 was isolated. This is explained by the initial formation of tetraene 26,
which undergoes two successive fast conrotatory and disrotatory cyclisations. In this
(15)
(16)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

13
case, the stereoselectivity of the reaction was greater than the sensitivity limits available
in conventional product analysis.

2.2.3. Three-member rings. In the case of the even-member rings, the
electrocyclic reaction requires neutral molecules, but for the odd-member rings, it is
necessary to have cations, anions, or radicals. The selection rules also apply to the
concerted reactions of such species. A cyclopropyl cation opens up to an allylic cation (a
2-electron system) through a thermal disrotatory mode, while a cyclopropyl anion (a
4-electron system) gives rise to an allylic anion through a thermal conrotatory mode.

For 2,3-disubstituted cations 28, the disrotatory ring opening may proceed in two
ways and after quenching of the allyl cation, lead to either an E or Z olefin (reaction
(17)).

R
R
inwards outwards
R
R
A
-
R
R
A
Z
R
R
A
R
E
A
-
28


E-Z diastereoselectivity is also controlled by other factors, notably by the relative
configuration of the cyclopropane. The preferred direction is the one in which the
emerging lobes of the disappearing bond in cyclopropane can compensate (during the
course of the transformation of the -bond to a delocalised -orbital) for the gradual
development of a positive charge at the site of the leaving group (anchimeric assistance
of the -bond, see Figure 10). This requires that the substituents cis to the leaving group
(X) turn inwards (approach each other) following the Z pathway, whereas those which
are trans turn outwards (E pathway).

X
R
R
-X
-
R
R
-X
-
R
X R
R
X
R
E route
Z route


Figure 10. The anchimeric assistance of the -bond and its possible outcomes.

A proof for this is the difference in the rates of solvolysis of the two isomeric tosylates
29 and 30 where the approaching methyl substituents render the TS derived from the cis
tosylates unfavourable (reaction (18)). The relative acetolysis rates of 29 and 30 are
4500:1 at 150
0
C.


H
H
Me
Me
H
outwards
Me
H
OTs
H
H
Me
k
rel
=4500
29
Me
Me
H
H
H
inwards
Me
H
H
OTs
H
Me
k
rel
=1
30

(17)
(18)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

14

Even more dramatic is the behaviour of endo- and exo-isomers 31 and 32
(reaction (19)). The endo isomer smoothly undergoes solvolysis at 125
0
C with
concomitant ring cleavage, while the exo isomer is stable up to 210
0
C. In this case, it is
the outward movement of the substituents which is prevented since they are
incorporated in a ring. Furthermore, if such movement could occur, the final product
would be an E-cycloheptene (geometrically impossible).

Cl
H
H
H
inwards
k
rel
=11000
H
H
31 endo
H
Cl
H
H
outwards
k
rel
=1
32 exo
H
H


Another similar example is presented in reaction (20). Here, the leaving group is
the endo-Cl substituent, despite it being rather hindered and less reactive that the exo-
Br substituent.

H
Cl
OH
X
Cl
Br
H
H
AgNO
3
H
2
O
H
Br
OH
33


3-Membered rings with 4 participating -electrons behave, as expected, in the
opposite manner as compared with corresponding cations. An example (21) is the ring
opening of two diastereomeric aziridines. The resulting 1,3-dipolar ion is subsequently
trapped by a cycloadditions reaction with dimethyl acetylenedicarboxylate, a very good
dipolarophile.

Since the rate of interconversion of the intermediate zwitterions (by C=N
+

rotation or carbanion inversion) is slower compared to the rate of cycloaddition, the
later process preserves the geometry of the parent species, and thus the overall reaction
sequence is stereospecific.

(19)
(20)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

15
N
H
COOR
H
ROOC
Ar
N
COOR
H
H
ROOC
Ar
N
H H
COOR ROOC
Ar
N
H ROOC
COOR H
Ar
COOMe
COOMe
COOMe
COOMe
N
COOMe MeOOC
Ar
H H
COOR ROOC
N
COOMe MeOOC
Ar
H ROOC
COOR H
100
100
con
h
dis
dis
h
con





3. Cycloaddition reactions


A cycloaddition reaction is one in which the terminals of two (or more)
independent and conjugated -systems combine themselves to form new bonds and
thus an additional ring
1
. The reverse process is known as cycloreversion or retro-
cycloaddition and it proceeds with the breaking of bonds. Depending on whether the
two components belong to different molecules or the same, the cycloaddition may be
inter- or intra-molecular. The reaction can be schematically depicted as in Figure 11,
where m and n represent the number of conjugated electrons in the two components,
respectively, and also the number of trigonal (sp
2
) carbon atoms.

(CH)
m
(CH)
n
(CH)
m-2
(CH)
n-2


Figure 11. Schematically representation of a cycloaddition reaction


With such a notation, the process can be called a [m+n] addition. The
nomenclature can be extended to include the nature of the participating electrons (,
or , the latter being used for electrons occupying a single interacting atomic orbital, as
in carbenes, carbanions or carbocations) and the mode of addition (supra or
antarafacial) on each component, e.g. [ms+na] addition.
(21)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

16


3.1. Frontier molecular orbital approach

In cycloadditions two components are commonly involved, and the feasibility of
a particular process will be determined by whether overlap can take place between the
HOMO of one component and the LUMO of the other.

The thermal cycloaddition of butadiene and a substituted ethylene (dienophile)
is a classical example of a Diels-Alder reaction (1). The HOMO of butadiene (2) and the
LUMO of ethylene (
*
), as also the LUMO of butadiene (3) and the HOMO of ethylene
(), are represented in Figure 12.

HOMO

2
LUMO

LUMO

3
HOMO



Figure 12. HOMO-LUMO interaction in thermal [4S+2S] cycloaddition

It can be observed that, whichever component has the HOMO or the LUMO, this
situation is a bonding one and thus the [4S+2S] cycloaddition is thermally allowed.
Since suprafacial-suprafacial (s,s) and antarafacial-antarafacial (a,a) combination give
here the same result, [4a+2a] addition is also thermally allowed and proceeds
simultaneously although less effectively, due to unfavourable geometry.

In the cycloaddition of the two substituted ethylene molecules to cyclobutene,
the ground state HOMO-LUMO combination does not permit thermal (s,s) addition
(Figure 13a.)

HOMO

LUMO

HOMO
(h)

LUMO

(a) (b)


Figure 13. HOMO-LUMO interactions in thermal (a) and photochemical (b) [2s+2s] cycloadditions

However, the cyclodimerisation takes place in the excited state interaction
(photochemical conditions) as can be seen from Figure 13b. The radiant energy is used
to promote an electron of one component into the orbital of the next higher energy level,
i.e.
*
, and its ground state LUMO (
*
) thus now becomes the HOMO, making the
[2s+2s] cycloaddition photochemically allowed.
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

17

On the other hand, the ground state interaction allows the [2s+2a] cycloaddition
(dotted lines in the Figure 14). In order to afford the orbitals overlap, the process goes
through an orthogonal transition state, but even in this disposition the overlap is only
minimal and worsened by the interfering effect of the substituents X and Y. Thermal
[2s+2a] are thus very rare.

Y
X X
Y
X
Y Y
X
HOMO

LUMO

X
Y
X
Y
X
X
Y
Y
LUMO
HOMO
1
2
4
3
X
Y Y
X
X
X
Y
Y
1 2
3
X
X X
Y X
Y Y
4
Y


Figure 14. HOMO-LUMO interaction in thermal [2s+2a] cyclisations

Such observations lead to the conclusion that the system with total number of
(4k+2) conjugated -electrons undergo thermal (s,s) cycloaddition, while those with 4k
electrons undergo photochemical (s,s) cycloaddition (Table 2). This prediction is the
particular form of the generalised Woodward-Hoffmann rule according to which a ground
state pericyclic change is symmetry allowed (and so facile) when the total number of
(4i+2)s and (4j)a components, on short, A and B, respectively, is odd. Thus in the
cyclisation of butadiene and ethylene, ethylene serves as the (4i+2)s component (i=0) and
butadiene as the (4j)s component (j=1), i.e. A=0 and B=0 making the total odd and
therefore the thermal ground state addition is symmetry allowed.

Table 2 Woodward - Hoffmann selection rules for cycloadditions

m+n (no. of
-electrons)
Condition
s
Geometry
h supra - supra
4 k supra -
antara
supra - supra
4 k + 2 h supra -
antara

A few relevant points emerge from the above discussion
1
:
a) for a two component cycloaddition the maximum number of modes of addition is 2
2
:
(s,s), (s,a), (a,s), (a,a); for p components, the number is 2
p
.
b) only in the (s,s) mode of the addition the two -systems approach in parallel planes.
In all other modes of addition, the planes are perpendicular.
c) the configuration of the marginal atoms is retained in the case of a suprafacial
component, and is inverted in the case of an antarafacial one.
d) for m>2 or n>2, there are two modes of (s,s) additions, one giving an endo product
and the other an exo product.
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

18
e) the [4S+2S] addition is the most facile closely followed by the [4a+2a] addition.


3.2. Stereochemical aspects

Concerted cycloadditions are almost totally stereospecific and the
stereochemistry is decided by the mode of addition as permitted by the Woodward-
Hoffmann rules. Some examples are discussed in the following pages.

3.2.1. [2+2] Additions. This type of reactions may involve two alkene fragments
undergoing dimerisation to give cyclobutane derivatives, or one alkene fragment and
another compound containing a double bond (such as ketenes or cumulenes). For the
particular case of the interaction between two molecules of cis-2-butene, Figure 15
presents all the theoretical modes of [2+2] addition and their stereochemical outcome
2
.


Me Me
Me Me
s,s
h
Me Me
Me Me
Me Me
Me Me
s,a
Me Me
Me Me

Me Me
Me Me
a,a
Me Me
Me Me
h
Me Me
Me Me
s,s
h
Me Me
Me Me
Me Me
Me Me
s,a
Me Me
Me Me

Me Me
Me Me
a,a
Me Me
Me Me
h


Figure 15. Possibilities of [2+2] addition in the case of cis-2-butene

Thermal [2s+2a] cycloaddition leading to cyclobutanes are rare, however,
because of the geometrical restraint in the orthogonal TS discussed before. But if the
double bond in the reacting species is twisted about its axis so that the two p orbitals are
no longer parallel and coplanar (as in compound 34 obtained by the photochemical
[4s+4s] addition of benzene and butadiene), there is better overlap of the FMOs and the
addition becomes facile (reaction(22)).

+
h
2 mol
34
H
H
35


(22)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

19
Actually, the intermediate adduct 34 spontaneously dimerises to 35. In this final
product one side of the cyclobutaric ring retains the trans configuration of the
suprafacial component, while the other becomes cis because of the antara interaction of
the copartner
29
(the upward hydrogen atoms are depicted by thick points).

Many of the photochemical [2+2] cycloadditions proceed through biradical or
ionic (in any case, non-concerted) mechanisms, particularly with chloro- and fluoro-
olefins
2
. A few synthetically useful [2s+2s] additions are known (e.g. reaction (23)).

Me
Me
O
h
O
H
Me H
Me 36


The process takes place from left to right in spite of the formation of strained
cyclobutanes 36. The reverse reaction is also governed by the W-H rules, and to go back,
the product would have to absorb light. But since it has now lost its bond, it doesnt
have any low-lying empty orbitals into which light can promote electrons.

The photochemical cycloreversion of butane can, however, be accomplished by
incorporating into the molecule a double bond serving as photon trap (37, reaction (24)).

h
H
H
H H 37


Conformational constrains limit the possibilities in the intramolecular reactions.
Formation of the cycloadduct 38 with the methyl group on the cyclobutanic ring on
adjacent carbon (1,2 rather than 1,3) would demand significant twisting of the
methylene tether, corresponding to a higher energy TS (reaction (25)).

O
Me
Me
Me
h
Me
Me
Me
O
not
Me
Me
O
Me
38


Ketenes, cumulenes or isocyanates with one or more sp hybridised carbon atoms
lacking a pair of interfering substituents at one of the reacting termini also undergo
[2+2] cycloadditions with olefins relatively easy. The ketenes and cumulenes behave as
antarafacial and the alkenes as suprafacial components. Two stereochemical features
are usually observed: the olefin expectedly retains its configuration but the adduct is the
one which is sterically more congested. The latter fact may be explained considering the
orthogonal TS characteristic to the (s,a) interaction (see figure 16). Consider for example,

(23)
(24)
(25)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

20
Me
H H
Me
C C O
EtO
H
Me
H
Me
H
EtO H
O
OEt
Me Me
H
H H
O
39


The arrangement in the TS is such that the sp carbon (less hindered) of the ketone is
directed towards the more hindered side of the olefinic double bond (i.e. C=O of the
ketene and the two methyl groups of cis-2-butene are on the same aide). As a result, the
kinetically controlled product of addition 39 becomes sterically more hindered (OEt
ends up on the same side as the Me groups). The stereochemistry thus reflects the
stability of the transition state and not that of the product.

Me
H
Me
H
LUMO
HOMO
O
p
z
p
x
HOMO
LUMO
(a) (b)


Figure 16. Molecular orbital interactions for the case of a [2s+2a] addition.
(a) top-view; (b) side view.


The intramolecular version of this reaction is usually more efficient than the
intermolecular one. Take for example the following reactions.

Cl
O
R
NEt
3
7
6
5
4
3
2
C
1
O
R H

5
6
2
3
4
7
1
O
H
R

Me
COOH
Me
C
O

O
H
Me


The intramolecular allene cycloadditions are also important, not only for the
generation of quaternary centers, but also for the construction of highly bridged
molecules. An example of a photochemical transformation from Weisners synthesis
of 12-epilycopodine is given in reaction (29).

(26)
(27)
(28)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

21
N Me
O
h
N Me
O


A final example of [2+2] cycloaddition is the reaction between an alkene and a
singlet

carbene (Figure 17.a). The HOMO of the carbene, a (usually denoted )


orbital with two electrons, adds antarafacial through a non-linear -approach to the
same side (i.e. suprafacial) of the LUMO of the olefin (Figure 17.b). It is thus a [2a+2s]
addition. In this orientation, the p orbital of carbene is pointing towards the electrons of
the bond, but this approach is however possible without the creation of a huge
amount of strain in the TS because the p orbital of carbene is empty. The alternative
linear and more symmetrical approach (Figure 17.c) is ruled out since it leads to an
antibonding interaction.

R
H
(a)
R
H
LUMO
HOMO
(b)
R
H
LUMO
HOMO
(c)


Figure 17. Addition of a carbene to an ethylene

The addition is stereospecific. Z-butene gives 80% cis-cyclopropane 40 upon
treatment with bromoform in a strong basic medium (reaction (30)), while E-butene
forms in the same conditions 68% trans- isomer.

Br
Br
H
Me
H
Me
40
H
Me Me
H
CHBr
3
tBuO
-
K
+


3.2.2. [4+2] Addition. Diels-Alder reaction. Several processes can be included in
this category, but the Diels-Alder transformation is the best known [4s+2s]
cycloaddition, proceeding stereospecifically syn with respect to both the diene and
dienophile, as expected from a concerted (s,s) mode of addition. An example was
previously given in reaction (1). The concerted nature of the addition of butadiene and
ethylene has been confirmed using suitably deuterated compounds (reaction (31)), the
product being almost exclusively (> 99%) the cis-isomer 41
30
.

Triplet carbenes behave more like radicals and do not give concerted cycloadditions.
(29)
(30)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

22
D
D
D
D
H D
H D

D D
D D
H
H
D
D
41



A few characteristic features need to be emphasised:

a) The diene must be able to react in the s-cis conformation. Those structural elements that
perturb this conformation (voluminous substituents, rings, a.s.o.) inhibit or retard the
reaction. Thus bulky 1-cis substituents slow the process, whereas bulky 2-substituents
speed it up (scheme (32)).

2
3
4
1
R
H
R
H
4
3
1
H
2
R
H
R
cisoid transoid cisoid transoid


b) The HOMO of the diene and the LUMO of the dienophile are generally closer in
energy than the HOMO of the dienophile and the LUMO of the diene. Those structural
features which raise the energy of the diene HOMO (e.g. electro-releasing substituents)
and lower the energy of the dienophile LUMO (e.g. electro-withdrawing substituents)
make the reaction procede faster

. The process works only poorly, if at all, in the


absence of the latter. Thus maleic anhydride reacts with butadiene at a much faster rate
than ethylene. In reaction (33), a double bond made strongly dienophile by an adjacent
carbocation gives an intermolecular Diels-Alder reaction
31
. (Syn or suprafacial addition
of trans alkenes gives a trans ring junction).

Me Me
OH
CF
3
SO
3
H
-23C
Me Me
H
H


c) Regioselectivity in Diels-Alder additions (ortho and para effects) is another
characteristic property. Thus dienophiles bearing electro-withdrawing substituents
react with 1-substituted butadienes to give 3,4-disubstituted cyclohexenes, independent
of the nature of the diene substituents (reaction (34)). (Of course, electro-releasing
groups on the diene increase both the rate and the selectivity, by a pull-push
mechanism.) Table 3 sums up a series of experimental results
3
related to this behaviour.
For example, 1-methoxybutadiene reacts (35) with acrolein to give exclusively the ortho
but not the meta adduct.

A rarer type of Diels-Alder addition is the reverse electron demand reaction in which the dienophile has electro-
releasing groups and the diene has a conjugated E group. The interaction is between the HOMO of the dienophile
and the LUMO of the diene.
(31)
(32)
(33)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

23

R
Z
R R
Z
Z
major product


OMe
CHO
X
CH
O
OMe OMe
CHO
meta ortho



Table 3. Ortho effect in Diels-Alder reactions

Diene subst. (R) Alkene subst. (Z) ortho:meta
Me CHO 8:1
Me CN 10:1
Me COOMe 6.8:1
iPr COOMe 5:1
nBu COOMe 5.1:1

The ortho effect is valid for 1-substituted dienes. In the case of 2-substituted
dienes, one can observe a so-called para effect (see Figure 18).


X
Z
X
Z
Z
Z
X X
ortho
para


Figure 18. Regioselectivity in Diels-Alder reactions.
(X stands for an electro-releasing substituent,
while Z denotes an electron-accepting substituent)

Although in many cases the regioselectivity may be determined form the
consideration of electrophilic and nucleophilic centers at the two components (see the
curved arrows in reaction (35)), the effect is primary determined by the coefficient of the
HOMO and LUMO orbitals
5
.

d) One of the most important characteristics of the Diels-Alder reaction is the endo
rule or endo-selectivity. When both the diene and dienophile are substituted, the endo
isomer is the major kinetically controlled product (reaction (36)) even though it may be
thermodynamically less stable than the exo isomer (into which it may be converted by
prolonged heating).
(34)
(35)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

24

X
X
X
endo exo


Table 4 presents some experimental results
3
showing the preference of the
reaction towards the endo product for various X-substituted olefins.

Table 4. Endo rule in Diels-Alder reactions















This behaviour can be explained by the presence of a secondary orbital
interaction between non-bond-forming orbital lobes of the same sign in the frontier
MOs. In the reaction between cyclopentadiene and acrolein (Figure 19), this interaction
is shown by a dotted line which is clearly absent in the formation of the exo product.

CHO
H
CHO
CHO
H
endo exo


Figure 19. Endo selectivity in a Diels-Alder reaction

For the thermally allowed [6s+4s] cycloaddotion, this interaction has an
antibonding character so that the exo isomer is the kinetically controlled product. A
more general rule for predicting such behaviour was deduced by K.N.Houk
3
. It states
that, for all allowed [4k+2] cycloaddtions between a (4k) -electrons polyene and a
Substituent X endo:exo
COOH 75:25
COOMe 76:24
CONH2 10:1
CHO Only endo
CN 60:40
CH2OH 80:20
CH2Br Only endo
CH2NH2 Only endo
CH2CN Only endo
CH2COOH Only endo
NO2 Mostly endo
OAc 81:19
OCHO Only endo
Br Mostly endo
(36)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

25
substituted alkene having extended conjugation, the endo transition state is stabilised
by secondary orbital interactions (Figure 20.a), whereas for the allowed [4k+4l+2]
cycloadditions of two polyenes (l 0) the endo transition state is destabilised (Figure
20.b).

(4k)
-el
2 -el
HOMO
LUMO
(a)
(4k)
-el
HOMO
LUMO
(b)
(4l+2)
-el


Figure 20. Secondary orbital interactions in Diels-Alder reactions.
Houks general approach

e) Site selectivity is that selectivity shown by a reagent towardss one site (or more) of
a polyfunctional molecule, when several sites are, in principle, available
5
. In
cycloadditions, site selecivity always involves a pair of sites, e.g. the Diels-Alder
reactions of anthracene generally take place across the 9,10-position than across the 1,4
or 3,9a. The highest coefficients in the HOMO of anthracene are at the 9,10-position, but
the more familiar explanation is that in this way are created two isolated benzene rings,
whereas reaction at the 1,4-position would create a naphthalene ring, less stable an
arrangement than two benzenic rings.

Reaction (37) illustrates an intramolecular Diels-Alder reaction giving a product
which is ultimatelly converted into brexane-2-one 44
32
. Due to [1,5] H-sigmatropic shifts
(see later), the side chain in the cyclopentadiene may approach any of the five carbon
atoms but it participates in the form 42 with the side chain at C-5 giving the product 43
exclusively.

COOMe
COOEt
H
115
H
H
COOEt
H
H
H
O 42 43 44



Intramolecular Diels-Alder reactions. If the diene and alkene are connected by a
chain of atoms (usually carbon), the Diels-Alder reaction occurs intermoleculary. This
has become one of the most powerful methods in organic synthesis for constructing
carbon bonds with high regio- and stereo-selectivity. Close approach of the diene and
olefin is sometimes inhibited by the atoms which link the two (the tether). The geometry
of the diene and dienophile (cis / trans) also plays a role. Consider for example the
structures 45 and 46.

(37)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

26
H
H
45 46


The former compound attains the proper geometry for the reaction, while the latter
requires significant distortion of diene and alkene.

The length of the tether is also important. Small rings are impossible to be
obtained, whereas large rings (8-11 member atoms) are formed with difficulty.

At least 2 different regioisomers are possible for internal Diels-Alder reactions,
depending on the orientation of the alkene as it approaches the diene. If the tether is
short, only 47 is possible, but with longer chains 48 can become a side product or even
the major one.

47 48


In general, E-dienes do not give products such as 48, although Z-dienes do.
When the reactive termini approach, as in 49, the alkene arm can assume an exo mode
(product 50) or an endo mode (product 51). The exo mode leads to a trans product,
while the endo TS leads to a cis product (scheme (39)). The final outcome depends on
the conformational energies of 50 and 51; in general, the exo TS is preferred.

exo
endo
H
H
50
H
H 51
H
H
H
H
H
H
H
H
49



The great synthetic advantage of the Diels-Alder reaction is its generality; the
variety of different dienophiles that can be used is very wide indeed (possible variations
in the diene are somewhat less wide), and conditions can usually be found to make the
great majority of such reaction proceeds in good yields and with high stereoselectivity.


3.2.3. Other [2+4] Additions.

(38)
(39)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

27
Dipolar cycloadditions. The 4-electrons component in a [4+2] addition is not
necessary a four-atom system (as in 1,3-dienes), nor it involves carbon atoms only, as
long as the HOMO-LUMO interaction can be achieved. The most common of these non-
dienic 4-electrons systems involve three atoms and have one or more dipolar
mesomeric structures, hence the term 1,3-dipolar additions. The only significant
difference between these cycloadditions and the Diels-Alder reaction is that they
involve less symmetrical, somewhat dipolar, transition states. The process is thermally
allowed and is highly stereospecific (syn addition to olefins). A general example is given
in reaction (40).

A B

B A


The alkene or alkyne derivative is called a dipolarophile. A variety of dipolar reagents
can be used, such as diazoalkanes (RCH=N
+
=N

), nitrile ylids (RCN


+
CH2), nitrile
oxides (RCN
+
O

), azides (RN=N
+
=N

), nitrones (R1R2C=N
+
RO

), and even ozone (

OOO
+
). In general, the low energy interaction is the one between the LUMO of the
dipole and the LUMO of the alkene. Following the FMO approach, the LUMO interacts
with the HOMO both suprafacially, as can be seen from Figure 21.a for the case of the
diazomethane.

N N
H
H
HOMO
LUMO
(a)
N N
H
H
LUMO
(b)
N N
H
H
LUMO
HOMO
(c)


Figure 21. Molecular orbital interactions for 1,3-dipolar cycloadditions

For the linear dipoles, the LUMO has a node through one of the terminal atoms
(Figure 22.b). If one needs to consider the alternative HOMO alkene-LUMO dipole
interaction, one must use the next lower unoccupied molecular orbital (NLUMO) instead
(see Figure 22.c).

Both intermolecular and intramolecular reactions are possible. An isomeric
bridged cycloadduct can be produced in this latter process if the tether is long enough
(reaction 41).

A B
B A
( )
n
( )
n
B
A
( )
n


Using the energies and orbital coefficients for the HOMO and LUMO involved,
accurate predictions of reactivity and regioselectivity can be made. However, these
values are sensitive to structural modifications and it is risky to give general
(40)
(41)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

28
predictions, therefore each case should be considered individually. Take for example
the following reactions.

H
2
C N N
COOMe
N
N
COOMe


HN N N
COOMe
HN
N
N
COOMe


N
O
N
O
N
O


O
O
O
H
OH
O
O
OH
O
O



Cheletropic reactions. A cheletropic reaction is a cycloaddition in which one
component interacts through a single atom, i.e. with electrons. The previously
discussed [2a+2s] addition of a carbene to olefins is such a cheletropic reaction. If the
olefin is replaced by a diene, it will be a 6-electron system and thermal [4s+ 2s]
addition would take place. Although the reaction is not known for carbene addition,
sulphur dioxide does react with a diene, e.g. E,E-1,4-dimethylbutadiene (reaction (46))
to form stereoselectively the cis-sulphione 52.

Me
Me
SO
2

SO
2
Me
Me
H
H
52


The reverse reaction (exclusion of SO2 from the sulphone) is more facile. Similarly, E,Z-
1,4-dimethylbutadiene gives the trans-sulphone or vice versa. One can envisage a
disrotatory motion in the diene component bringing the two methyl groups to their
respective configuration (Figure 22).

Me
H
H
Me
S
O
O
LUMO
HOMO


Figure 22. Molecular orbital interaction for a cheletropic reaction.

(42)
(43)
(44)
(45)
(46)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

29

3.2.4. Higher cycloadditions. Periselectivity

Among the higher cycloadditions, a photochemical [4s+4s] addition of
butadiene to benzene has already been mentioned (reaction (22)). Another example is
provided below.

[4+4]


[6+4] Cycloadditions are rare and go by the (s,s) mode, as illustrated in the addition of
cyclopentadiene and tropone 53 (reaction 48).

CO
X
7
2
3
O
CO

H
H
endo exo
53
[6+4]
[4+2]
not
O
54


The product is exclusively the exo isomer which is explained by an unfavourable
interaction between lobes of unlike sign in the HOMO-LUMO combination when
forming the endo TS (see previously Houks general approach for secondary orbital
interactions, Figure 20). Moreover the [6+2] addition takes place in preference to a Diels-
Alder [4+2] reaction between the butadiene unit and the C2-C3 bond (both are thermally
allowed). This phenomenon, known as periselectivity, is explained by the fact that the
coefficients of the frontier orbitals of the LUMO of tropone are highest at atoms C2 and
C7.
In general, the ends of conjugated systems carry the largest coefficients in the
frontier orbitals, and thus the pericyclic reactions use the longest part of conjugated
system according to the Woodward-Hoffmann rules. This proves to be true up to a
point, with the provision that the reactions have also to be geometrically reasonable.
The following cases are all ones where the largest possible numbers of electrons have
beer used, although smaller (but equally allowed) cycloadditions might have taken
place.

N
COOMe MeOOC
N
COOMe
MeOOC
H
H
N
COOMe
MeOOC
- H
2
[8+2]

(47)
(48)
(49)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

30

COOMe
COOMe
[8+2]
COOMe
COOMe


N
COOEt
N EtOOC
NCOOEt
NCOOEt

N
N
COOMe
COOMe
[8+2]
NCOOMe
NCOOMe


In open-chain and in some cyclic systems we can predict that FMO control will
lead the reaction to take the path which uses the longest part of the conjugated system
consistent with a symmetry-allowed reaction, but several other factors, spatial, entropic,
a.s.o., have to be taken into account. However, in some more complicated cases, the
longest conjugated system is not always the one preferred. The following examples
illustrate same of the complexities of such systems.

The reaction (53) between tetracyanoethylene 56 and heptafulvalene 55 is one in
which the geometrical factors outweight the contribution of the frontier orbitals
5
.

[14+2]
1' 1
2
3
4
CN NC
CN NC
NC CN
NC CN
55 56


The HOMO coefficients for 55 are highest at the central double bond, but any reaction at
his site would have to be antarafacial on one of the components, and this is sterically
unreasonable. (However, in a carbene cycloaddition, in which an antarafacial element
can be taken up by carbene, it is this central bond that is attacked). The next best
possibility from the FO view point would be a Diels-Alder reaction across the 1,4-
positions, but this was shown not to occur probably because the carbon atoms in the
seven-member cycle are too distant

. The only remaining possibility is the site of lowest


orbital electron density, i.e. the antarafacial reaction across the 1,1

-positions.

Sesquifulvalene 57 presents another case where frontier orbital control is not the
only factor that governs periselectivity.

This is known to influence the rate of Diels-Alder reactions, e.g. cyclopentadiene reacts much faster than
cyclohexadiene, which in its turn reacts much faster than cycloheptatriene.
(50)
(51)
(52)
(53)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

31
5 4
3
2
1
7
6
8
12
11 10
9
57


The sesquifulvalene 58 does give the adduct 59 with tetracyanoethylene by a
[8+2] addition (reaction (54)), as expected from the coefficients of the HOMO of the
unsubstituted system 57, which are greatest for positions C2 and C8.

CN NC
CN NC
[8+2]
CN
CN
CN
CN
58 59


However, the sesquifulvalene 60 gives the [4+2] adduct 61 instead, by an
addition (55) across the positions C2 and C5, while the sequifulvalene 62 gives the [12+2]
adduct 63 (reaction (56)) using the orbitals located at C2 and C12. Furthermore, the
sesquifulvalene 64 gives yet another kind of adduct 65 in Diels-Alder reaction (57) when
treated with acetylenedicarboxylic ester. This examples serve to emphasize the pitfalls
of a too easy application of FMO theory.

Ph
Ph Ph
Ph
CN NC
CN NC
[4+2]
60
61
Ph
Ph Ph
Ph
CN
CN
NC
CN


CN NC
CN NC
[12+2]
62
CN
CN
CN
CN
H
H
63


(54)
(55)
(56)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

32
[4+2]
64
COOMe
COOMe
COOMe
COOMe
65


Cycloadditions of multiple components can also occur, indeed very rarely, in a
bi- or even uni-molecular manner by incorporating two or three of the -conjugated
systems (in the proper orientation) in a molecular structure. Additions of the type
[2s+2s+2s] are thermally allowed, while those of the type [2s+ 2s+ 2s+ 2s] are
photochemically allowed. Some examples are presented below (reactions (58), (59) and
(60)).

O
O
O
H
H

O
O
O
H
H

O O
NC CN
CN NC
O O
NC CN
NC CN


COOMe MeOOC
h
MeOOC COOMe




Conclusion

It is clear from the preceding sections that a wide range of cyclic and polycyclic
molecules can be prepared by pericyclic reactions. In addition, stereocontrolled
syntheses of acyclic molecules are possible via initial cycloaddition, followed by
cleavage of the ring. The power of the reaction lies in the ability to make carbon bonds
with high regio- and stereo-selectivity and specificity. Few synthetic methods allow
access to such a large number of products, with compatibility to a wide range of
functional groups. In conclusion, for power and versatility, few reaction types rival
pericyclic transformations.

(57)
(58)
(59)
(60)
Stereochemical Aspects of Pericyclic Reactions Eugen S. Andreiadis

33


References

1. D. Nasipuri, Stereochemistry of Organic Compounds. Principles and Applications, Wiley Eastern Ltd., New
Delhi, 1991.
2. M. Nogradi, Stereochemistry. Basic Concepts and Applications, Pergamon Press, New York, 1981.
3. M.B. Smith, Organic Synthesis, McGraw-Hill, New York, 1994.
4. J.M. Harris, C.C. Wamser, Fundamentals of Organic Reaction Mechanisms, John Wiley & Sons, New
York, 1976.
5. I. Fleming, Frontier Molecular Orbitals and Organic Chemical Reactions, John Wiley & Sons, London,
1976.
6. J. Clayden, N. Greeves, S. Warren, P. Wothers, Organic Chemistry, Oxford University Press, New
York, 2001.
7. P. Sykes, A Guidebook to Mechanism in Organic Chemistry, Longman, New York, 1982.
8. F. Badea, Reaction Mechanisms in Organic Chemistry (Mecanisme de Reactie in Chimia Organica),
Editura Stiintifica, Bucharest, 1973.
9. J. Hendrickson, D. Cram, G. Hammond, Organic Chemistry (Chimie Organica), Editura Stiintifica si
Enciclopedica, Bucharest, 1973.
10. M. Jones, Organic Chemistry, W.W. Norton & Company, New York, 1997.
11. S.N. Ege, Organic Chemistry, D.C. Health and Company, Lexington, 1984.
12. D.C. Neckesr, M.P. Doyle, Organic Chemistry, John Wiley & Sons, New York, 1977.
13. R.O.C. Norman, Principles of Organic Synthesis, Chapman and Hall, London, 1978.
14. Y. Jean, F. Volatron, J. Burdett, An Introduction to Molecular Orbitals, Oxford University Press, New
York, 1993.
15. A. Williams, Concerted Organic and Bio-Organic Mechanisms, CRC Press, New York, 2000.
16. W. Carruthers, Cycloaddition Reactions in Organic Synthesis, Tetrahedron Organic Chemistry Series, vol
8, Pergamon Press, London, 1991.
17. A.J. Bellamy, An Introduction to Conservation of Orbital Symmetry, Longman, London, 1974.
18. E.L. Eliel, S.H. Wilen, M.P. Doyle, Basic Organic Stereochemistry, Wiley-Interscience, New York, 2001.
19. R.B. Woodward, R. Hoffmann, J. Amer. Chem. Soc., 1965, 87, 395.
20. R.B. Woodward, R. Hoffmann, J. Amer. Chem. Soc., 1965, 87, 2046.
21. R.B. Woodward, R. Hoffmann, J. Amer. Chem. Soc., 1965, 87, 2511.
22. R.B. Woodward, R. Hoffmann, Angw. Chem., Int. Ed., 1969, 8, 781.
23. E.E. van Tamelen, S.P. Pappas, J. Amer. Chem. Soc., 1963, 85, 3297.
24. E.E. van Tamelen, S.P. Pappas, K.L. Kirk, J. Amer. Chem. Soc., 1971, 93, 6092.
25. J. Amer. Chem. Soc., 1972, 94, 5906.
26. W.R. Dolbier, H. Koroniak, D.J. Burton, P. Heinze, Tetrahedron Lett., 1986, 27, 4387.
27. G. Maier, Angw. Chem., Int. Ed., 1967, 6, 402.
28. E. Havinga, R.J. de Kock, M.P. Rappold, Tetrahedron, 1960, 11, 278.
29. k. Kraft, G. Koltenburg, Tetrahedron Lett., 1967, 8, 4357.
30. K.N. Houk, Y.-T. Lin, F.K. Brown, J. Amer. Chem. Soc., 1986,108, 554.
31. P.G. Gassman, D.A. Oingleton, J. Org. Chem., 1986, 51, 3075.
32. A. Nickon, A.G. Stern, Tetrahedron Lett., 1985, 26, 5915.




Work presented at the Student Communications, May 2003 Edition
Copyright 2003, 2004 by Eugen S. Andreiadis, all rights reserved
Contact: andreiadis@dap.ro

Vous aimerez peut-être aussi