CH 391: Unit 3: Stereochemistry: I.Stereochemistry of Nucleophilic Substitution Reactions at Saturated Carbon
CH 391: Unit 3: Stereochemistry: I.Stereochemistry of Nucleophilic Substitution Reactions at Saturated Carbon
Whatever the charge type of the nucleophile or the leaving group, these
reactions invariably occur with complete inversion of configuration at the saturated
(sp3) carbon undergoing substitution. One of the earliest (and especially elegant)
demonstrations of inversion of configuration in an SN2 reaction is illustrated in the
reaction of optically pure 2-iodooctane with radioactive iodide ion. The observed
reaction,t he incorporation of radioactive iodine into the substrate, is rigorously
second order (first order in iodide ion and first order in iodooctane), as it should be for
an SN2 mechanism. The key observation is that the rate of incorporation of radioactive
iodine into 2-iodooctane is exactly one-half of the rate of racemization of the optically
active substrate. This can only be the case if each reaction of radioactive iodide ion
with 2-iodooctane racemizes two molecule of this substrate. In turn, this can only be
the case if each reaction occurs with inversion of configuration, the inverted product,
combined with a molecule of the original product, then provides two molecules of
racemic (R + S) 2-iodooctane. If the reaction had occurred via a carbocation route,
racemization would have been the result of each reaction with radioactive iodide ion,
making the rate of racemization exactly equal to the rate of radioactive iodide ion
incorporation.
Of course the determination of absolute configurations of reactants
and products now makes the determination of reaction stereochemistry
relatively straightforward. The theoretical question of especial interest is
why the preference for inversion over hypothetical retention is so powerful
in all of these reactions. Incidentally, it clearly has nothing fundamentally
to do with the circumstance that a negatively charged nucleophile may
electrostatically repulse a leaving group, which in the molecule is uncharged
but which is also in the process of becoming negatively charged in the
transition state, in a retentive mechanism, since inversion is still strongly
preferred even when the leaving group (when present in the molecule) is
positively charged and the entering nucleophile is negatively charged
(resulting in an electrostatic attraction). The theoretical baisis for the highly
developed preference for an invertive reaction mechanism can be seen by
applying either the concept of transition state aromaticity/antiaromaticity or
by applying frontier orbital (HOMO/LUMO) theory. The former approach
has already been developed, and for the specific case of nucleophilic
substitution, in Unit 1. The explanation given there was that the transition
state for retentive substitution is unavoidably a cyclic one and it possesses
four electrons (two from the nucleophile and two from the C-L bond), so it
is antiaromatic, whereas the transition state for the invertive process is
acyclic, so that this TS is non-aromatic.
In contrast, the (R,R) and (S,S) stereoisomers of the bromoalcohol both afford
a racemic mixture of (2R,3R) and (2S,3S)-dibromobutane. This amounts to retention
of configuration in the conversion of the (2R,3R) alcohol stereoisomer to the (2R,3R)
dibromide, but in inversion of configuration at both C2 and C3 for the conversion of
the (2R,3R) alcohol to the (2S,3S) dibromide. All of these results are readily and
uniquely explicable if it is assumed that bromine acts as a neighboring group (i.e., as
an intramolecular nucleophile), displacing the leaving water molecule from C2 with
inversion of configuration (as in a normal SN2 reaction) thus forming an intermediate
epibromonium ion. The latter then opens up by reaction with the external nucleophile
(bromide ion), as epibromonium ions do in what is essentially an SN2 reaction, with
inversion. Since the epibromonium ion formed from either the (R,R) or (S,S) alcohol
is achiral (it has a plane of symmetry which bisects the central C2-C3 bond), it can
react with bromide ion at either carbon (the one which formerly was C3 or the one
which formerly was C2) equally to yield the two enantiomers of 2,3-dibromobutane.
The distinct epibromonium ions derived from the (2R,3S) and (2S,,3R) alcohol are
chiral (do not have a plane or other relevant element of symmetry), but their reaction
produces meso-2,3-dibromobutane, which is achiral. Thus, the products of all of these
reactions are optically inactive, even though they are formed from enantiomerically
pure starting alcohols.
Analogous results are found in other reactions in which a
suitably placed substituent having an unshared electron pair is present in the
molecule (e.g., alkoxy, alkylthio, and amino, and amido substituents. Even
certain electron-rich aryl substituents
(e.g. 4-methoxyphenyl = anisyl) are found to act as effective neighboring
groups, e.g., in the solvolysis (cleavage of solvent) of the 3-anisyl-2-butyl
tosylates, shown below. Note that the erythro/threo nomenclature for
distinguishing diastereoisomers is convenient to use here. The erythro
isomer of a compound containing two non-equivalent stereocenters is the
one (like erythrose) in which all of the like groups can be placed anti to one
another in one of the conformations accessible to the molecule. In the case
under discussion, hydrogens are anti to hydrogens, methyls to methyls, and
the tosylate (4-methylbenzenesulfonyloxy) and anisyl groups (the “like”
groups which are not identical) are also anti. In the single product obtained
in the acetolysis (solvolysis in acetic acid), the like groups are still anti to
one another, i.e., only the erythro isomer is obtained.
Further, only one enantiomer of the erythro diastereoisomer is
obtained, since neither the intermediate arenium ion nor the product are
achiral. In other words, both enantiomers of the erythro diastereoisomeric
tosylates give different, enantiomerically pure products which are the result
of retention of rigorous configuration at both stereocenters. In
contrast, either threo enantiomerr gives an intermediate arenium ion which
has a plane of symmetry and is therefore achiral. It generates racemic threo
acetate (a 50:50 mixture of the 2R,3S and 2S,3R enantiomers). See if you
can draw this out and show the structures of the two enantiomers of the
threo product.Incidentally, the cationic intermediate in these reactions has
been called a “phenonium ion”by Professor D.J.Cram, in whose research
group this work was performed. It is a special kind of arenium ion, i.e., the
type of cationic intermediate which is involved in electrophilic aromatic
substitution. Finally, the participation of an internal nucleophile in the rds
naturally results in rate enhancements in comparison to direct reactions with
external (solvent) nucleophiles.
In certain rigid bicyclic systems, such as the bicyclo[2.2.1]heptyl system (the
norbornyl system), even alkyl groups can act as neighboring groups. This is
illustrated in the acetolysis of optically pure exo- 2-norbornyl tosylate, which results
in the formation exclusively of the exo isomer, but in racemic form. Obviously, an
SN2 reaction would result in the formation, exclusively, of the endo acetate, whereas
an SN1 reaction should result in the formation of a mixture of endo and exo acetates.
Importantly, in the case of either of these mechanisms, an single enantiomer would
result, i.e., optically active products would be formed. Instead, only the exo product is
obtained, and it is 100% racemic. This results stem from neighboring group
participation by the C6 carbon atom to form a symmetrical (i.e., achiral), bridged
cationic intermediate, which has been called a nonclassical carbocation. The latter can
react at either of two equivalent carbon atoms (formerly C2 and C1) to yield both
enantiomers of the exo acetate in equal amounts.
The question of why the reaction does not occur via a concerted, syn
stereospecific path, instead of the stepwise carbocation path is an interesting one. The
carbocation path involves a rate-determining step in which two bonds are broken and
only one is formed. It is strongly endothermic. On the other hand, the concerted
addition pathway would forrm two bonds and break two bonds and is obviously
exothermic. Why is the stepwise path, which appears to be less favorable from a
thermodynamic standpoint, actually the lower energy pathway. If we look at the TS
for a concerted addtion of HCl to a pi bond we can see that the TS has antiaromatic
character, i.e., it has a cyclic system which contains four electrons. Consequently, the
stepwise path is of lower energy. It should be noted that concerted addition of HCl to
1,2-dimethylcyclohexene would give only the cis chloride product (i.e., the product in
which the methyl groups are cis o each other on the cyclohexane ring).
The resonance treatment of the transition state for this reaction has already
been given in Unit 1, but is repeated here. Note that in the TS the hydrogen which is
being transferred to carbon from boron is still partially linked to the boron atom.
Consequently, the formation of the new C-H bond and the new C-B bond must take
place from the same face of the pi bond. It would be geometrically impossible for the
hydrogen to be transferred to the opposite face of the pi bond while still maintaining
significant bonding to boron. Essentially, this is why syn stereospecific reactions are
almost always concerted reactions and conversely, i.e., concerted mechanisms
essentially always result in syn stereospecific addition.
There is another interesting aspect of this hydroboration reaction. Since we
seem to have a cyclic TS with four electrons (two from the pi bond and two from the
B-H bond), this would seem to set up an antiaromatic TS, which should be strongly
disfavored. In fact, this TS is not rigorously cyclic when viewed in terms of the orbital
overlaps. In particular, the 2p AO on the carbon which is bonding to boron is bonding
to a 2p AO on boron, but the hydrogen which is being transferred to carbon is bonded
to boron via another AO (a boron sp2 AO) which is orthogonal to (has zero overlap
with) the 2p AO on boron. Consequently, the overlaps of the system do not continue
through boron, which acts as an insulator between the hydrogen and the carbon to
which boron is bonding. Consequently, it should be recalled that not all four-
membering ring transition states necessarily are antiaromatic.
Diels-Alder Cycloadditions. Perhaps the most famous, and also the most useful, of
all cycloaddition reactions is the Diels-Alder reaction of a dienophile with the s-
cis conformation of a conjugated diene. Again, the reaction is highly syn
stereospecific, both with respect to the dienophile and the diene. The stereospecific
reactions of diethyl maleate ( which has cis ester functions, E) and diethyl fumarate
(trans E’s) with 1,3-butadiene are depicted below (NP = nodal plane). In the specific
depictions, the dienophile is shown as approached the termini (C1 and C4) of the s-
cis conformation of the diene from above, so that the bottom lobes of the dienophile
orbitals overlap with the top lobes of the diene orbitals. The orbital signs suggested by
the light and dark lobes imply the interaction of the dienophile HOMO (which is
symmetric, like that of the BMO of ethene) with the LUMO of the conjugated diene
(this MO is also symmetrical and has two nodal planes, one between C1-C2 of the
diene and one between C3-C4 of the diene). Note also that the coefficients of the
diene LUMO are larger the termini (C1 and C4) than at the internal positions, as
implied by the sizes of the AO’s depicted. Most importantly, from the point of view
of FO theory, the interactions between both termini of the diene and the dienophile
are bonding, a circumstance which follows from the matched symmetrics of the
dienophile HOMO and the diene LUMO. The same favorable bonding situation is
found when examining the interactions of the diene HOMO and the dienophile
LUMO (try this as an exercise).
It is also of interest to note that for the cis dienophile there are two
diastereoisomeric transition states, both of which lead to the same cis Diels-Alder
adduct. The endo face of the TS is considered to be the one which lies directly (or
nearly directly) over the diene (especially C2 and C3), while the exo face is the one
which projects away from this diene moiety. In one cis TS both ester functions are
endo and in the other both ester functions are exo. In the single TS for the addition of
the trans dienophile, one ester function is endo and the other exo. We shall see
momentarily the reasons for the importance of recognizing the endo and exo faces of
a Diels-Alder transition state.
Not only is the addition to the dienophile syn stereospecific but, as the
drawing implies, it is also a syn stereospecific addition to the termini of the
conjugated diene, i.e., both new bonds to the dienophile are formed from the same
face, here shown to be the top face of the diene. The consequence of this
stereospecificity element is that different geometric isomers of the diene lead to
different adducts. For example, addition to E,E-2,4-hexadiene leads to adducts in
which the two methyl groups are cis to one another in the adduct, as shown below.
Again, the cis dienophile can orient the ester substitutents either endo or exo, but this
time the adducts are differentiated, and the endo one is slightly favored (more about
this later). The key points here are (1)the E methyl groups turn out to be cis to one
another in the adduct and (2) the majority endo product has the ester groups cis to the
methyl groups.