Aliphatic Nucleophilic Substitution
Aliphatic Nucleophilic Substitution
GCU, LAHORE
+
H2O + CH3- S(CH3)2 CH3 OH2 + (CH3)2S
+
HO + CH3- S(CH3)2 CH3 OH + (CH3)2S
1
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
As a result of these reactions the alkyl group is attached to the nucleophile,
therefore, the nucleophile is said to be alkylated. Sometimes, the solvent itself functions
as a nucleophile; the reaction is then called solvolysis. Thus, a reaction in which water is
used as a solvent and it also acts as a nuc1eophile, is called hydrolysis while a reaction
involving methanol as a solvent and also as a nucleophile is known as methanolysis.
In nucleophilic substitution reactions the R-X bond of the substrate undergoes
heterolysis, and the unshared pair of the nucleophile is used to form a new bond to the R
group.
Nu + R X Nu R + X
Heterolysis
occurs here
The question arises when does the R-X bond break? Does it break at the same time that
the new bond between the nucleophile and R forms? Or does the C-X bond breaks first?
The answer can be explained by the conditioned dependent mechanisms.
MECHANISMS:
The mechanism of a reaction is the actual pathway through which the reaction
proceeds, i.e., which bonds are broken, which are formed, and in what order; how many
steps are involved and what is the relative rate of each step, etc.
The SN1 mechanism consists of two steps. The first step involves slow ionization
of the substrate resulting in the formation of a carbocation that rapidly combines with the
nucleophile to form the product in the second step. Mostly the solvent itself acts as a
nucleophile, therefore the SN1 reaction is generally called as solvolysis.
In a reaction involving more than CH3 CH3
one step the slowest step Step 1 H3C C I Slow
H3C C + I
determines the rate of the (Ionization)
reaction and is therefore called CH3 CH3
the rate-determining step. So, CH3
CH3
in this reaction the first step, i.e., H2O
OH2
the ionization of the substrate, is Step 2 H3C C
Fast
H3C C
2
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
Evidences:
There are some evidences that support this mechanism e.g.
a) Kinetic study
b) Stereochemistry
c) Rearrangement
a) Kinetic Study
b) Stereochemistry
Another strong evidence in favour of this mechanism also comes form the
stereochemical studies. This mechanism involves the formation of a carbocation as an
intermediate. Since the central carbon atom of the carbocation is sp2 hybridized, it is a
planar molecule. The unhybridized p orbital on carbon is perpendicular to the plane of the
molecule with one lobe on each side of the plane. The nucleophile can therefore attack
from either side (front or back) of the plane to form a bond with the carbocation. Since
the carbocation is a symmetrical molecule, the chances of attack form both sides are
almost equal, and if we start with an optically active substrate, we are expected to get a
racemic mixture of the product.
a a b
R1 R1
R1 R1
C X Y
C C Y + Y C
R3 R3 R3
R2 R2 R3 b -X R2 R2
Retention Inversion
(Predominates)
So, an aliphatic nucleophilic substitution is generally expected to be accompanied
by racemization if it proceeds by the SN1 mechanism. Anyhow the product is not
completely racemised because the inverted product predominates over its enantiomer.
The most probable answer to this question lies in the fact that attack of the nucleophile
occurs before the departing ion has completely left the neighbourhood of the carbonium
ion. Thus the departing ion shields the carbonium ion from the frontal attack (i.e. path a).
As a result back side (i.e. path b) is somewhat preferred.
3
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
c) Rearrangemet
CH3 CH3
OH OH2
The SN2 mechanism is involved in the substitution reactions that take place in one
step without the formation of any intermediate. In this mechanism the nucleophile attacks
the substrate from the side opposite to the leaving group, i.e., from the backside. The
bond formation between the nucleophile and the substrate and the bond cleavage
resulting in the removal of the leaving group occur simultaneously, i.e., the reaction is
concerted one.
Antibonding
orbital
δ δ
Nu + C L Nu C L
Transition state
Nu C + L
As the nucleophile starts making the bond with the central carbon atom of the
substrate, the leaving group starts leaving it and in the transition state both the
nucleophile and the leaving group are loosely bound to the carbon atom such that at no
time the carbon atom has more than eight electrons in its outer shell. The hybridization of
the central carbon atom changes from sp3 in the substrate to Sp2 in the transition state and
again to sp3 in the product. Thus, the central carbon atom and the three non-reacting
groups are approximately coplanar in the transition state, with the unhybridized p orbital
perpendicular to the plane and having one lobe on either side of the plane; one lobe of the
p orbital is engaged by the nucleophile and the other by the leaving group.
H
H H
δ δ
HO + C X HO C I HO C + I
H H
H H H
H
4
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
Evidences:
Like SN1, this mechanism is also supported by some evidences e.g.
a) Kinetic study
b) Stereochemistry
a) Kinetic Study
The substitution reactions proceeding through the SN2 mechanism involve only
one step and that is the rate-determining step. In this step both the substrate and the
nucleophile are involved. The rate of the reaction should therefore depend on the
concentration of both, i.e., the reaction should follow the second-order kinetics with the
following rate law:
Rate α [(Substrate] [ Nucleophile ] Rate = k2 [ CH3-X ] [ OH ]
Since two molecules, i.e., the nucleophile and the substrate, are involved in the
formation of the transition state in the rate-determining step (the only step of the
reaction), this reaction is said to be bimolecular.
However, if the nucleophile is used in large excess, e.g., the solvent acts as a
nucleophile, the mechanism may still be bimolecular but the kinetics will appear as first-
order. This is because there will be no apparent change in the concentration of the
solvent, due to its large quantities, and the rate will depend only on the concentration of
the substrate. Such a reaction is said to be pseudo-first-order.
b) Stereochemistry
Much more convincing evidence for the SN2 mechanism comes from the
stereochemical study of the reaction. In this pathway since the nucleophile attacks the
substrate from the backside, inversion of configuration should take place. This has
actually been observed in the reaction of radioactive iodide ion with optically active 2-
octyl iodide. If we start with an optically pure R-2-octyl iodide, the substitution of I in the
substrate by *I will produce enantiomer of the substrate, i.e., S-2-octyl iodide.
One molecule of the product C6H13 C6H13
together with one molecule *I + C I *I C + I
CH3
of the starting substrate will H3C
H
form a racemate because H
they are enantiomers of each R-2-octyl iodide S-2-octyl iodide
other.
This means that the exchange of iodide in each molecule of the substrate will
result in a racemic mixture of two molecules. In other words the reaction is not only
accompanied by racemization but also the rate of racemization is twice the rate of
exchange. This is only possible when there is an inversion of configuration, i.e., the
nucleophile attacks the substrate from the backside.
5
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
iii) SNi (INTERNAL NUCLEOPHILIC SUBSTITUTION) MECHANISM:-
O O O H
R R H R
C OH + S C O S C O S
H H H Cl Cl
Cl Cl R' Cl Cl R'
R'
O
O
- HCl
R R
SNi R S
S
C Cl C O Cl
H - SO2 H C O Cl
H
R' R'
R' Alkyl chloro sulfite
(Retention)
Thionyl chloride first reacts with the alcohol to form an alkyl chloro sulfite (these
alkyl chlorsulphites can be isolated). In the second step the sulfite group is lost and just
like in an SN1 reaction an alkyl carbocation is created. The actual nucleophile in the third
reaction step is the chlorine atom attached to the sulfite group which recombines with this
carbocation and results in the retention of stereochemical configuration.
Evidence:
6
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
O O
N N
R S R S R
C O Cl C O N Cl C
H - Cl H H
R' R' R'
Alkyl chloro sulfite (Inversion)
The reaction between alcohol and thionyl chloride is second order, which is
predicted by this mechanism, but the decomposition by simple heating of ROSOCl is first
order.
Some reactions of a given substrate under a given set of conditions display all the
characteristics of SN2 mechanism; other reactions seem to proceed by SN1 mechanism,
but cases are found that can not be characterized so easily. There seem to be some thing
in between, a mechanistic “borderline” region. At least two broad theories have been
devised to explain these phenomena. One theory holds that an intermediate mechanism is
caused by a mechanism that is neither pure SN1 nor pure SN2, but some “in between”
type. According to second theory there is no intermediate at all, and borderline behaviour
is caused by simultaneous operation, in the same flask, of both SN1 and SN2 mechanism;
that is, some molecules react by SN1, while others react by the SN2 mechanism.
Sneen et al., formulated an intermediate theory. According to it all SN1 and SN2
reactions can be accommodated by one basic mechanism i.e. the ion-pair mechanism.
The substrate first ionizes to an intermediate ion pair that is then converted into product.
An intimate ion pair is an ionic intermediate in which the cation and anion are in close
proximity and are not separated by solvent molecules.
k1 k2
RX R X Products
(Ion-pair)
The difference between the SN1 and SN2 mechanisms is that in the former case the
formation of the ion-pair (k1) is the rate determining, while in the SN2 mechanism its
destruction (k2) is rate determining. Borderline behaviour is found where the rates of
formation and destruction of the ion-pair are of the same magnitude.
Evidence:
One of the experiments that support the ion-pair mechanism is the hydrolysis of
4-methoxybenzyl chloride in 70 % aqueous acetone. In this solvent, hydrolysis (i.e.
conversion to 4-methoxybenzyl alcohol) occurs by an SN1 mechanism. When azide ions
are added, the alcohol is still a product, but now 4-methoxybenzyl azide is another
product. Addition of azide ions increases the rate of ionization (by the salt effect) but
decreases the rate of hydrolysis. If more carbocations are produced but fewer go to the
alcohol, then some azide must be formed by reaction with carbocations – an SN1 process.
However, rate of ionization is always less than the total rate of reaction, so some azide
must also form by an SN2 mechanism.
7
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
Thus, the conclusion is that an intermediate ion-pair is involved and SN1 & SN2
mechanisms operate simultaneously.
H2O
MeO CH2Cl MeO CH2 Cl MeO CH2OH
70 %
(Ion-pair) aq. acet.
H2O NaN3
v) SN1' MECHANISM:-
R CH CH CH2 R CH CH CH2
Of course, an allylic rearrangement is undetectable in the case of symmetrical allylic
cations, as in the case where R=H, unless isotopic labeling is used. This mechanism for
the rearranged product has been called the SN1' mechanism.
Evidences:
a) As with other SN1 reactions, there is clear evidence that SN1' reactions can
involve ion pairs. If the intermediate attacked by the nucleophile is a completely free
carbocation, then, (1) and (2) should give the same mixture of alcohols when reacting
with hydroxide ion, since the carbocation from each should be the same.
8
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
and in most others, the product spread is in the direction of the starting compound. With
increasing polarity of solvent, the product spread decreases and in some cases is entirely
absent. It is evident that in such cases the high polarity of the solvent stabilizes
completely free carbocations.
b) There is other evidence for the intervention of ion pairs in many of these
reactions. When CH2=CHCMe2Cl was treated with acetic acid, both acetates were
obtained, but also some ClCH2CH=CMe2, and the isomerization was faster than the
acetate formation. This could not have arisen from a completely free Cl - returning to the
carbon, since the rate of formation of the rearranged chloride was unaffected by the
addition of external Cl - ion. All these facts indicate that the first step in these reactions is
the formation of an unsymmetrical intimate ion pair that undergoes a considerable
amount of internal return and in which the counterion remains close to the carbon from
which it departed.
Thus, (1) and (2), for example, give rise to two different intimate ion pairs. The
field of the anion polarizes the allylic cation, making the nearby carbon atom more
electrophilic, so that it has a greater chance of attracting the nucleophile.
Nucleophilic substitution at any allylic carbon can also take place by an SN2
mechanism, in which case no allylic rearrangement usually takes place. However, allylic
rearrangements can also take place under SN2 conditions, by the SN2' mechanism, in
which the nucleophile attacks at the γ carbon rather than the usual position:
R R" R
R"
SN2' mechanism R C C C X R C C C + X
R"
R' R" Y R'
Y
Stereochemistry:
The stereochemistry of SN2` reactions has been investigated. It has been found
that both syn (the nucleophile enters on the side from which the leaving group departs)
9
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
and anti reactions can take place, depending on the nature of X and Y, though the syn
pathway predominates in most cases.
Syn Y R X Anti R X
R" R"
R C C C R C C C
R" R"
R' R'
Y
When a molecule in an allylic position has a nucleofuge capable of giving the SNi
reaction, it is possible for the nucleophile to attack at the γ position instead of the α
position. This is called the SNi' mechanism and has been demonstrated on 2-buten-1-ol
and 3-buten-2-ol, both of which gave 100 % allylic rearrangement when treated with
thionyl chloride in ether.
SOCl2
H3C CH CH CH2-OH H3C CH CH CH2 H3C CH CH CH2 + SO2
- HCl
2-buten-1-ol Cl O
Cl
S
Cl
10
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
1) KINETIC STUDIES:-
Hydrolysis
H3C H2C O CH2 CH2 Cl H3C H2C O CH2 CH2 OH
(in aq. dioxane)
Hydrolysis
H3C H2C S CH2 CH2 Cl H3C H2C S CH2 CH2 OH
(in aq. dioxane)
The hydrolysis of the sulfide (latter) is over 10,000 times faster than the ether
(former). This rate difference is far too great to be attributed to differences in inductive,
conjugative, or steric effects, but suggests rather that the hydrolysis of the sulfide (but not
the ether) proceeds through a cyclic onium ion (in this case, sulfonium ion).
Actually, oxygen being more electronegative can not donate its lone pair to
carbon but sulphur can easily donate its lone pair to carbon and thus assists in the
departure of the leaving group by forming a cyclic intermediate, consequently increasing
the rate of reaction over 10,000 times faster as compared to first.
very fast
CH2 Cl S CH2
S Et S CH2 CH2 OH + H
CH2 Cl Et CH2 H2O
Et
2) STEREOCHEMISTRY:-
The stereochemistry of a reaction might suggest that neighbouring groups become
involved. For example, the hydrolysis of the α-bromopropionate ion in water or dilute
base yields lactate with retention of configuration about the α-carbon.
H3C H3C
Hydrolysis
O Br O OH
C C
O H O H
Since nucleophilic substitutions at secondary carbon atoms almost invariably
result in partial or complete inversion of configuration, it was assumed that two
displacements were actually involved; the first a displacement of bromide by the
neighbouring –COO- to form the nonisolable α-lactone (A), and the second a very rapid
cleavage of the lactone by water. Inversion presumably occurred in both displacements,
the second inversion “nullifying” the first; hence we get retention of configuration.
11
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
H3C H3C (A) CH3 H3C
H2O
O C Br O C Br C
O O C OH2
C Br very fast
C C H C
O H O H O H
O
3) MOLECULAR REARRANGEMENT:-
The neighbouring group participation may lead to molecular rearrangement when
the neighbouring group remains bonded to the reaction center but breaks away from the
atom to which it was originally attached in the substrate. Thus, the chlorinated amine (B)
yields, upon basic hydrolysis, the rearranged aminohydrin (D), presumably because the
intermediate imonium ion (C) is attacked preferentially at the primary α-carbon atom
rather than the secondary β-carbon atom.
(B) Et
Et (C) Et
Et (D) Et
Et
CH Cl N CH
N N CH
CH2 Cl Et CH2 OH
Et Et CH2OH
4) SPECTROSCOPIC STUDIES:-
In intermediate 1 the hydrogens on two OCH3
is formed by participation.
Thus NMR also serves as an evidence
CH2 CH2
for neighbouring participation.
OCH3 OCH3 OCH3 OCH3 OCH3 OCH3
Cl
CH2 CH2 Cl
H2C CH2 H2C CH2 CH2 CH2 H2C CH2 CH2 CH2
12
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
O
Br 3 O H2O 4
H3C CH CH2 C O O
H3C
Br 5 O O
H2O 6
H3C CH (CH2)2 C O
O
H3C
O
7 O O
Ag2O 9
8 O +
Br (CH2)5 C OH HO (CH2)5 C O
H2O
10 O O
11
Br (CH2)6 C O HO (CH2)5 C O (no lactone)
H2O
Thus the conversion of γ-bromovalerate (5) in water to a five-membered ring
lactone, 6, proceeds by direct displacement, and same is probably true for the formation
of β-butyrolactone (4). But as is the case with other types of cyclization, formation of
rings of seven or more membered entails some difficulty, due to the low probability for
collision between opposite ends of a long chainlike molecule. When ε-bromocaproic acid
(7) is treated with Ag2O in water, ordinary solvolysis to give the ε-hydroxycaproic acid
competes with neighbouring-group participation (which gives the seven-membered ring
lactone, 8. Anion 10, yields only a hydroxyl acid (no eight-membered lactone).
When the carboxylate group is converted by protonation to the carboxyl group, -
COOH, it becomes very much less nucleophilic and loses a great deal of its effectiveness
as a participant. The substitution of carboxyl group for an α-methyl in (CH3)2CHBr
lowers the rate of solvolysis in water about a hundred-fold, whereas a similar substitution
in PhCHMeBr retards solvolysis by a factor of about 105. If any anchimeric assistance
effects are present, they are completely overshadowed by the strongly negative inductive
effect of the –COOH group. It has been found that the “deaminations” of α-amino acids
proceed with retention of configuration but in this case it is not clear whether paricipation
by the –COOH or the –COO- HNO 2 D-R CH COOH + N2
group is involved, since the D-R CH COOH OH
D-R CH COOH NO
NH2 C D-R CH COOH + N2
N N intermediate is a l
Cl
strong acid.
Whether or not an ester linkage functions as a neighbouring-group depends
largely upon how it is situated with respect to the reaction center in the substrate. The
solvolyses of CH3CHBrCOOEt and the deamination of CH3CH(NH2)-COOEt result in
inversion of configuration about the α-carbon, suggesting that there is negligible
participation by the –COOEt group when the reaction center lies alpha to the C=O bond
of the ester. This is what we might expect, for such participation would require the form-
ation of a strained three-membered ring X
13
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
section of the ester molecule; in each of these, the anchimeric intermediate contains a
five-membered ring.
The conversion of H3C
C O 12
trans-2-acetoxycyclohe-
O
xyl brosylate (12) to H3C
C O
trans-1,2-diacetoxycyl- 15
O
ohexane (15) proceeds O HO Acetoxonium ion
Ac
through the symmetric -B 14
Ac
O S O sO -
intermediate acetoxoni- O
O-
H3 C
C OAc
um ion 14, and similar
evidence points to the O OAc 17
same intermediate in the Br +
14
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
24
H3C H3C 26
Cl OH H3C
N H2C Cl CH2
N N H2C OH
H3C H3C or
C H3C
H2 25 CH2 H2O C
H2
As indicated, the five-membered ring ammonium ion 23 (the pyrrolidinium ion),
is stable. Ethyleneimonium ions, such as 25, are, as class, somewhat more reactive than
ethylene oxides, although nitrogen containing such ions can, with care, be isolated. As
might be expected, the rate at which the various cyclic ammonium ions are formed
depends markedly on the number of atoms in the ring. The five or six membered rings,
being stable, are readily formed.
These imonium ions undoubtedly also intercede in rearrangements, not only of
such aliphatic haloalkylamines as on page 12, but also such heterocyclic amines as N-
ethyl-2-chloromethylpyrolidine (27).
Et Et Et
27 CH2
N CH2 Cl N N
1 Cl Cl H2C CH2
H2C 5 2CH H2C CH
28 29
4 3
H2C CH2 H2C CH2 H2C CH
C Cl
H2
What has been said about neighboring-group participation by nitrogen may be
applied, with but slight modification, to sulfur. The hydorlyses of β- and δ-chlorosulfides
are thousands of times as rapid as those of the corresponding chloroethers and, when
carried out in the presence of base, are zero order in base. These reactions doubtless
proceed through cyclic sulfonium ions of the type 31 and 33, analogous to the cyclic
ammonium ions 25 and 23.
15
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
R R
30
32 S
S
R S H2C Cl H3C S CH2
H2C α H2C Cl H2C CH2
β 33
C α Cl 31 CH2 β γ δ Cl
H2 H2C CH2 H2C CH2
The most familiar (and the most extensively studied) chloroalkyl sulfide is
mustard gas (34) [β,β′-dichlorodiethyl sulfide], the hydrolysis of which involves the
formation of the ethylenesulfonium ion 35, followed by the much faster destruction of this
ion by water. As befits a reaction sequence of this type, the first-order rate constant for
the hydrolysis will “drift downward” as the reaction proceeds, due to the reversal of the
cyclization when intermediate 35 is attacked by Cl- accumulating in solution (the mass
law effect). Moreover, if the reaction is carried out in the presence of an additional
nucleophile, even one as weakly nucleophilic as formate or chloroacetate, much of
intermediate 35 may be converted to a substitution product (for example, formate 37)
rather than to the hydrolysis product, 36.
Starting material
Cl
CH2 34 CH2 CH2 OH
CH2 35 H2O
CH2 Cl S 36
S S CH2
HC
Cl CH2 CH2 Cl
CH2 CH2 Cl CH2 CH2 Cl _
OO
O
Mustard gas
Rate determining CH2 CH2 O C H
S 37
Product
determining CH2 CH2 Cl
The fact that such relatively weak nulceophiles as chloride and format may, even
at low concentration, compete with the solvent water for the cationic intermediate is
further evidence that this intermediate is indeed sulfonium ion, rather than the very much
more reactive primary carbonium ion, ClCH2–CH2-S-CH2CH2+. While the latter is not
excluded by the kinetics, its lifetime would be extremely short, and all but a very small
fraction of such ions would be converted to the hydrolysis product 36 (by action of a
solvent molecule from their solvent “cages”) before colliding with a nucleophilic species
present in small concentration in solution.
As we have noted, anchimeric assistance by negatively charged oxygen or by an
ester linkage requires that the assisting group and the leaving group lie trans to each
other. The fact that the trans forms of 2-chlorocyclohexyl phenyl sulfide (38) and the
corresponding cyclopentyl derivative are hydrolyzed 105 to 106 times as rapidly as their
respective cis isomers indicates that trans orientation is required for β-sulfur participation
also. But trans orientation, although necessary, is apparently not sufficient; for it has been
ArS
H
SPh Cl
40
38 39 H SAr
Ar = Me
H
Cl H
Cl
16
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
found that the trans-chlorosulfide 39 is solvolyzed (in 85 percent aqueous ethanol at very
nearly the same (very low) rate as is its cis isomer, and that the trans-chlorosulfide 40 is
solvolyzed similarly slowly. The rigidity of the structure in compounds 39 and 40 will
prevent a coplanar arrangement of the four atoms involved in formation of the sulfonium
bridge (the sulfur atom, the chlorine atom, and the α– and β-carbons), it is reasonable to
infer that the stereochemical requirements for anchimeric assistance by β-sulfur (and
probably for assistance by β-nitrogen and β-halogen also) are similar in nature to those
for facile bimolecular elimination. In elimination, the same the rigidity of the ring
systems in these compounds prohibits the formation of the transition states necessary for
trans elimination, thus leading to very low rates of biomolecular dehydrochlorination,
even when hydrogen and chlorine atoms on adjacent carbons lay trans to each other.
Although more data are clearly needed, it may be tentatively assumed that
effective anchimeric assistance requires a transition state in which the α- and β-carbons,
the leaving group, and the “assisting atom” lie in or near a common plane.
Neighboring-group participation by a β-bromo or β-iodo group results in the
formation of a cyclic (three-membered ring) bromonium or iodonium ion. Although such
halogenonium ions are too reactive to be isolated, the stereochemistry of conversions of
the diastereomeric 3-bromo-2-butanols (41) to 2,3-dibromobutanes (44), and that of the
acetolyses of these dibromides in the presence of Ag+, point to the intervention of
bromonium ion 43 in the erythro-meso series, and to the intervention of the cis isomer of
43 in the threo-d, l series.
OH 42 OH2 43 H
Me 41 Me Me
+
H
H H H H
H2O H Me
Br Erythro Me Br Me Br trans
_
c Ag Br
OAc O A Me Br
Me 45
44
H H H H
Erythro Meso
Br Me Br Me
17
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
corresponding bromobrosylates the trans isomer is likewise solvolyzed more rapidly than
the cis, but here the effect (Ktrans/Kcis = 800) is much less striking.
The substrate structure is the major factor that controls the mechanism and the
rate of a substitution reaction. As we have seen, the SN1 mechanism involves the
intermediate formation of a carbocation in the rate-determining step. The ease of
formation of the carbocation depends on its stability. The greater the stability of the
carbocation, the more rapidly it is formed. The order of stability of the carbocations is:
tertiary > > secondary > primary > methyl. So, the SN1 reaction is more favourable for
the tertiary substrates than for the secondary or primary substrates.
The SN2 reactions proceed through a transition state that involves five groups
attached to the central carbon atom and becomes rather over-crowded. The alkyl groups
attached to the central carbon atom will make the transition state more over-crowded, and
thus more unfavourable, than would do the hydrogen atoms. Also, the approach of a
nucleophile from the backside to a tertiary carbon atom is sterically more hindered than
to a primary carbon atom. Thus, the SN2 reactions are more favourable for the primary
substrates and less favourable for the tertiary substrates. In fact, the tertiary substrates
seldom react by the SN2 mechanism; even the neopentyl systems in which substitution
occurs at primary carbon atom, react so slowly as to make it synthetically useless because
the approach of the nucleophile from the backside is severely hindered by three β-methyl
groups.
Generally, the primary substrates react by the SN2 mechanism, whereas the
tertiary substrates react by the SN1 mechanism. The secondary substrates could proceed
either by SN1 or SN2 or both mechanisms depending on the reaction conditions, though
mostly they also follow the SN2 mechanism. If the reaction conditions are not suitable for
SN1 mechanism in the case of tertiary substrates, elimination usually predominates.
Similarly, when the rate of the SN2 reaction is slowed down by the structural effects,
alternate side reactions (mostly elimination) begin to compete.
Comparison of rate constants (Table 1) for SN1 reactions shows an increase by a
factor of 106 in changing the substrate form primary to tertiary. Similarly, rate constants
for SN2 reactions decrease as the number of alkyl groups surrounding the central carbon
atom increases.
18
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
Table 1: Effect of the Substrate Structure on the Relative Rates of SN1 and SN2 Reactions
R k1(SN1) k2(SN2)
H3C 1 30
CH3CH2 1 1
(CH3)2CH 12 0.02
(CH3)3C 1,200,000 0
(CH3)3CCH2 0.00001
Any species, whether neutral or negatively charged, that has an unshared pair of
electrons can function as a nucleophile. Nucleophile is not involved in the rate-
determining step of an SN1 reaction. It is only the second step when the nucleophile
comes into the picture and combines with the intermediate carbocation to form the
product. The rate of the SNI reaction is therefore not influenced by the nucleophile. For
example, the rate of hydrolysis of tert.butyl bromide which follows SN1 mechanism is not
affected by changing the nucleophile from H2O to HO-. On the other hand, the SN2
reaction requires a nucleophile to push off (displace) the leaving group in the rate-
determining step (the only step). The rate of this reaction therefore greatly depends on the
nucleophile. For example, the rate of hydrolysis of methyl bromide that proceeds by SN2
mechanism increases by more than 5000 times when the nucleophile is changed from
H2O to HO- because the latter is a better nucleophile than the former.
Similarly the rate in SN2 reaction is also dependent on the concentration of both
the substrate as well as the attacking nucleophile so if we double the concentration of the
nucleophile keeping the concentration of the substrate same, the rate in this case also
becomes double.
19
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
a better base. The opposite order of nucleophilicity and basicity in this case can be
explained as follows:
The outer electrons of the larger atoms are spread over greater volume than are
those of smaller atoms. The less-localized electrons form weaker bonds to a proton. The
larger atoms are therefore less basic. The outer electrons of the larger atoms are also less
tightly held by the nucleus, i.e., they are more polarizable. In other words, they are more
available for forming bonds to a carbon atom. The larger atoms are therefore more
nucleophilic.
Steric considerations are also important in determining the nucleophilicity of a
nucleophile. When a Lewis base interacts with a proton, steric requirements are normally
important because proton is a small species. However, approach of the same reagent to a
tetracoordinate carbon atom can involve severe steric interactions. For example, whereas
methoxide ion can easily approach the carbon atom during a nucleophilic substitution
reaction, tert-butoxide ion, being very bulky, is sterically hindered to approach the central
carbon atom. Thus, the ability of tert-butoxide ion to function as a nucleophile is greatly
reduced, although its basicity is similar to that of the methoxdie ion.
I + R-OH No reaction
Under the acidic conditions necessary to protonate the leaving group the
nucleophile would also be converted to its conjugate acid. It would therefore no more act
as a nucleophile. Since SN1 reactions do not require powerful nucleophiles but do require
good leaving groups, most of them take place under acidic (or neutral) conditions. On the
other hand, SN2 reactions which do require powerful nuclephiles, mostly take place under
20
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
basic (or neutral) conditions.
Another way of converting –OH to a better leaving group is to convert it to a
reactive ester such as sulfonic ester. The sulfonic ester group like tosylate (p-
toluenesulfonate), is a better leaving group than halide and is frequently used in the
nucleophilic substitution reactions.
O O
PTSA
R-OH R O S CH3 R-I + O S CH3
O I
O Tosylate ion
The nucleophilic substitution reactions are heterolytic processes that usually take
place in solution. Each reaction involves either creation, dispersion or destruction of
charges. The reaction medium (solvent) can therefore play an important role in
determining the mechanism and the rate of these reactions through dipolar interactions
with the reactants, the intermediates and even the transition states.
The polarity of a solvent, usually determined from its dielectric constant (Table
2), is an indication of how well it can stabilize the charged species. The greater the
polarity of a solvent the greater its ability to stabilize a charged species, i.e., the greater
its ionizing power. The SN1 reactions, in which carbocations are formed in the rate-
determining step, are more favourable in polar solvents. For example, the rate of
hydrolysis of tert-butyl chloride increases more than 1000 times as the solvent is changed
from ethanol to water because water is more polar than ethanol.
H2O 81 (CH3)2SO 45
HCOOH 59 CH3CN 38
33
CH3OH HCON(CH3)2 37
24
C2H5OH CH3COCH3 23
6 C2H5OC2H5 4
CH3-COOH
21
CHEM-2201 ORGANIC SECTION
GCU, LAHORE
solvent, whereas the reactions in which the charges are either destroyed or dispersed in
the transition state are favoured by decreasing the solvent polarity. Generally, the
influence of a solvent is large when the charges are either created or destroyed; the
effects are small when there is a dispersal of charges.
If we want to favour the reaction of an alkyl halide by an SN2 mechanism, we should use
a relatively unhindered alkyl halide, a strong nucleophile, a polar aprotic solvent, and a
high concentration of the nucleophile. For substrates, the order of reactivity in SN2
reaction is
The effect of the leaving group is the same in both SN1 and SN2 reactions: alkyl
iodides react fastest; fluorides react slowest. (Because alkyl fluorides react so slowly,
they are seldom used in nucleophilic substitution reactions.)
Reference Books
1. M. Younas “A Text Book of Organic Chemistry”, 2nd edition 2005, Al-Hajaz
Printing Press, 18-A Darbar, Ilmi Kitab Khan, Urdu Bazar, Lahore, Pakistan.
2. Arun Bahl, B. S. Bahl, “A Text Book of Organic Chemistry (For B. Sc.
Students)”, 18th edition 2006, Rajendra Ravindra Printers (Pvt.) Ltd., S. Chand
and Company Ltd., New Delhi-110055, India.
3. Michael B. Smith and Jerry March, “March’s Advanced Organic Chemistry,
Reactions, Mechanisms, and Structure”, 5th edition, John Wiley and Sons, USA.
22