Sunteți pe pagina 1din 31

CHAPTER

1 BASIC CONCEPTS

The outlook towards organic chemistry has undergone great changes in recent years. Besides
synthesis and the determination of the structure of molecules, the study, both chemical and
kinetic, of organic reactions has been given considerable attention with a view to understanding
the mechanism of reactions. Organic chemists continue to be engaged in the search for new
reagents and new reactions. Yet, simultaneously, old reactions are being reinvestigated in order
to establish their mechanisms. This has been possible by new experimental, analytical and
spectroscopic techniques. Lewis first presented the concept of covalent bond in chemistry. The
covalent bond was subsequently classified into polar and non-polar, depending on the
electronegativities of the bonded atoms. A knowledge of the nature and strength of bonds is
essential for the chemical investigation of organic molecules. Moreover, the properties of a
molecule are always influenced by its structure. A brief description of certain basic concepts
has been attempted in this chapter.

1.1 BOND DISSOCIATION ENERGY

It is a well known fact that when a covalent bond is formed between two atoms, a certain
amount of energy is released. For instance, when hydrogen atoms combine to form hydrogen
molecules, 104 kcal of heat is evolved for each mol of hydrogen formed. The bond dissociation
energy instead, is the energy required (∆H° at 25°C and 1 atm pressure) in cleaving a bond into
two radicals. In the case of hydrogen molecule, a certain amount of energy is expended in
splitting the molecule, i.e.,

H — H (g) 2H (g), ∆H ° = 104.0 kcal/mol
The dissociation energy of the H—H bond is thus 104 kcal/mol. Such a process is an
endothermic process, therefore, ∆H° will be a positive value. This information is obtained by
molecular spectroscopy, thermal methods and electron impact studies.1 Walsh argued theoretically
that the strength of a covalent bond increased with increase in the difference of electronegativity
of the bonded atoms. This is apparent from the values listed in Table 1.1 (cf. carbon—halogen
bonds).
The bond dissociation energy is not equal to the bond energy. The definition of bond
energy is though less precise, since it is the average value of the energy required to break a
bond between two atoms in a molecule. Let us, for instance, consider a water molecule which
contains two H — O bonds. It is possible to measure the bond dissociation energy for the
removal of one and then the second H atom from H2O. The result is computed as follows:
• •
H — O — H (g) O — H (g) + H (g), ∆H ° = D1 = 119.7 kcal/mol
• • •
O — H (g) O (g) + H (g), ∆H ° = D2 = 101.5 kcal/mol
1
2 ORGANIC REACTION MECHANISMS

It is apparent that the bond dissociation energy for the two steps is different but from
Hess’s Law one can state that their sum is equal to the energy of atomization (conversion of
a molecule to its atoms) of the water molecule, i.e., 221.20 kcal/mol.
• •
H2O (g) O (g) + 2H (g), D1 + D2 = 221.20 kcal/mol
In water molecule there are two O—H bonds present that are equivalent and energy value
can be assigned to each bond which is equal to the average value of D1 and D2.
1
(D1 + D2) 110.6 kcal/mol.
2
This is the bond energy and is often referred to as the mean theoretical bond energy and
should be distinguished from the bond dissociation energy.
For diatomic molecules such as H2, Cl2, HCl, etc. bond energies and bond dissociation
energies are equal. The bond dissociation energy for the same molecule may differ from molecule
to molecule (see Table 1.1). The higher the bond energy, the stronger is the bond. Also increasing
s-character shortens a bond so bond strength increases with s-character.
In a chemical reaction a new bond is usually formed while the old one is broken. It is,
therefore, important to know the bond dissociation energy, which is a measure of the strength
of a bond, to understand a chemical reaction. This may also reflect on the nature of the
reaction.
TABLE 1.1: Bond Dissociation Energies (kcal/mol)
Bond Bond Dissociation Bond Bond Dissociation
Energy (∆H°) Energy (∆H°)

D— D 106.0 Na — Cl 98.1
H— H 104.2 H— F 136
C— C 83.1 H — Cl 103
N— N 38.4 H — Br 88
O— O 33.2 H— I 71
F— F 36.6 I — Cl 50.3
Cl — Cl 58.0 Si — F 129.3
Br — Br 46.1 C— N 69.7
I— I 36.1 C— F 108
Si — Si 42.2 C — Cl 84
P— P 51.3 C — Br 70
S— S 50.9 C— I 56
Si — H 76.0 CH3 — Cl 81
H— O 102.4 CH3CH2 — Cl 81
CH3 — H 104 (CH3)2C — Cl 81
CH3CH2 — H 98 |
H
(CH3)2C — H 94
(CH3)3C — Cl 78
|
H C6H5CH2 — H 85
(CH3)3C — H 90 C6H5CH2CH2 — H 70
Si — C 85.7 CH2=CHCH2 — H 85
Si — Br 69.1
C— Si 69.3
Except for the values of H—O, D—D and Na—Cl, other values are taken from S W Benson, J. Chem. Educ., 42, 502
(1965). The values of H—O, D—D and Na—Cl are from K S Pitzer, J. Am. Chem. Soc., 70 2140 (1948).
BASIC CONCEPTS 3

It is noticed from the above Table that C — H bond dissociation energies for ethane
(98 kcal/mol) and for toluene (85 kcal/mol) are considerably different and smaller for toluene.
The obvious reason for this difference is that the benzyl radical produced by bond dissociation
in toluene is substantially stabilized by resonance.

1.2 ELECTRONEGATIVITY

A fundamental property of a chemical bond is polarity. The polarity of a bond depends on the
nature of the atoms linking that bond. The preferred tendency shown by an atom of a covalent
bond to attract the shared pair of electrons towards itself is termed as a measure of
electronegativity of that atom. This unequal attraction of electrons causes a charge separation
+ –
between the atoms and produces a dipole, indicated by a charge notation (δ or δ ) and shown
for a bond between two atoms, i.e., HCl molecule:
Hδ+— Clδ–
The partial negative charge on the chlorine atom indicates that the chlorine atom has a
greater tendency to attract the shared electron pair towards it. As a result the electron density
of the bonding electrons lies more towards Cl atom. Chlorine is more electronegative than
hydrogen. A complete transfer of electrons takes place in the case of ionic bonds. We often use
electronegativities as a guide to predict whether a given bond will be polar or not and also the
direction of its dipole moment.
Electronegativity values have been measured from a consideration of bond distances,
bond energies, dipole moments, and the data for some elements are listed in Table 1.2.

TABLE 1.2: Electronegativity Values of Different Elements


H
2.1
Li Be B C N O F
1.0 1.5 2.0 2.5 3.0 3.5 4.0
Na Mg Al Si P S Cl
0.9 1.2 1.5 1.8 2.1 2.5 3.0
K Ca Ge As Se Br
0.8 1.0 1.7 2.0 2.4 2.8
Cs Ba
0.7 0.9

Source: L Pauling, The Nature of the Chemical Bond, 3rd ed., Cornell University Press, Ithaca, (1960).

From these values the following conclusions may be drawn:


1. On going from top to bottom in a group, as the atomic size of the element increases
the electronegativity decreases.
2. In a given period, electronegativity increases with increasing atomic number.
3. The more non-metallic an element, the greater is the electronegativity. The value of
electronegativity for fluorine is the largest.
The electronegativity of carbon is almost similar to that of hydrogen, as a result C—H
bond is usually considered as non-polar. As we will study in Chapter 4, nucleophilicity decreases
with increasing electronegativity of the attacking atom. The electronegativity order for hybridized
carbon orbitals is sp > sp2 > sp3.
4 ORGANIC REACTION MECHANISMS

1.3 DIPOLE MOMENT

In the case of non-polar bonds, the electric charge is equally distributed between similar atoms
forming the bond. On the other hand, in case of polar bonds there is a charge displacement
towards the more electronegative atom (causing a charge separation and making the bond
polar). When the centers of positive and negative charges do not coincide, the bond may
possess a dipole moment designated by the symbol µ. The dipole moment is equal to the
product of the charge (e) and the distance (d) of separation between the atoms as shown in
equation (1.1). A molecule is considered to possess dipole moment when the two bonded atoms
µ = e × d …(1.1)
–10
differ in electronegativity. Since the electronic charge is 10 e.s.u. and the distance between
atoms is in angstrom units (10–8 cm), the dipole moment is obtained in Debye units,
D = 4.8 × 10–18 e.s.u. It is not possible to measure the dipole moment of an individual bond
within the molecule. We can measure only the total dipole moment of the molecule. Dipole
moment values for some compounds are listed in Table 1.3.
TABLE 1.3: Dipole Moments of Some Common Organic and Inorganic Compounds
(Values Determined in the Vapor State Unless Otherwise Mentioned)
Compound µ Compound µ
(in D) (in D)

C6H5CH3 0.37 CH3I 1.64


*
C6H5Si(CH3)3 0.44 CH3OH 1.69
BrCl 0.57 p-ClC6H4Si(CH3)3 1.70
HCl 0.65 C2H5OH 1.74$
HBr 0.79 CH3Br 1.79
H 2S 0.93 CH3F 1.81
†, †
CH3COOH 0.94 CH3Cl 1.83
HCOOH 1.06†, † H2 O 1.84
HCl 1.08 p-CH3C6H4Cl 1.95†
C6H5COOH 1.13†, ‡ p-CH3C6H4Br 1.96†
CH3Cl 1.22$ ClCH2COOH 1.97†,$
$
C2H5OC2H5 1.22 O
y
C6H5OH 1.40 HCH 2.23
O
y
NH3 1.47 CH3CCH3 2.76$
C6H5NH2 1.48 HCN 2.93
$
C6H5Br 1.55 CH3CN 3.94
$
C6H5Cl 1.58 C6H5NO2 4.01$
SO2 1.61 p-CH3C6H4NO2 4.42†

Source: J W Smith, Electric Dipole Moments, Butterworths Scientific Publications, London,


84–308, (1955). *J D Roberts et al., J. Am. Chem. Soc., 71, 2923 (1949); †In benzene solution; ‡Values for the dimeric
molecule; $In solution.
BASIC CONCEPTS 5

The dipole moment of a molecule will vary depending on the other atoms present in the
vicinity. It is a vector quantity. In polyatomic molecules the dipole moment of the molecule as
a whole is equal to the vector sum of the individual bond moments. It is thus expected that
a molecule can be non-polar but still contain polarized bonds. This is possible if the bonds are
so directed that the dipole moments of the individual bonds cancel out, i.e., their vector sum
is zero. In carbon tetrachloride, for instance, the C — Cl bond possesses bond moment of value
2.3 D. Since the bonds are symmetrical the dipole moment is cancelled out. Carbon tetrachloride
therefore, has a net zero dipole moment and is non-polar. Chloroform, on the other hand, has
a value of 1.22 D and is polar. Unshared pair of electrons make large contributions to the dipole
moment of water and ammonia. This is explained by the polarity of the lone-pair of electrons
present on O and N. A lone pair contributes to a large moment directed away from the central
atom.

O N N
H H H H F F
H F
µ = 1.84 D µ = 1.5 D µ = 0.2 D

Water molecule has a dipole moment of 1.84 D. This proves that it is not a linear molecule.
Contribution to the net dipole moment arises from individual bond moments and the lone-pair
as shown above. Ammonia has a higher dipole moment than NF3. This is due to the fact that
in NH3 the N—H bonds and the lone-pair moments reinforce each other while in NF3 they
oppose each other.
Molecules containing dative bonds are expected to possess large dipole moments since the
formation of such bonds requires a marked separation of charges. Because of small differences
between the electronegativity of C and H, alkanes have very small dipole moments.
A knowledge of dipole moment is of considerable importance in the study of molecular
geometry of molecules. For example, geometrical isomers can be identified unambiguously,
from dipole moment data.

1.4 HYDROGEN BOND

Hydrogen bond is a weak bond formed by the attraction of a hydrogen on one atom with
another atom bearing an unshared electron pair. In other words, the hydrogen atom forms a
bridge between two electronegative atoms. Thus hydrogen can participate in hydrogen bonding
with O, N or F. (organic molecules do not contain H—F bonds). The hydorgen bond is denoted
O — H ····· O O — H ····· N
by a dotted line. This bond is much weaker than the covalent bond and energy of such a bond
in a molecule ranges from 2–10 kcal/mol. Compounds like alcohols, acids, water, ammonia, etc.
display hydrogen bonding.
The existence of this bond has been established by various techniques and molecular
properties. In n.m.r., for instance, a hydrogen-bonded hydroxyl group shows a downfield shift
of its proton. Similarly, i.r. studies reveal that the stretching vibration of the O—H bond
6 ORGANIC REACTION MECHANISMS

decreases in the presence of a hydrogen bond. Varied properties like mp, bp, heat of dilution
and solubility point etc. are due to the existence of hydrogen bond.
Two types of hydrogen bonds have been recognized, viz., intramolecular (within the same
molecule) and intermolecular (between two or more molecules). Both types markedly influence
the properties of molecules.2, 3 Hydrogen bonding plays an important role in many chemical
and biochemical processes as well.
Intramolecular hydrogen bonds are considered to influence the acidity of certain acids.
Thus, salicylic acid is about eighteen times more stronger than benzoic acid whereas
p-hydroxybenzoic acid is half as acidic as benzoic acid. The large acidity of salicylic acid is

OH OH COOH

COOH COOH HO OH

COOH
Ka –5 –5 –5 –2
6.25 × 10 105 × 10 2.9 × 10 6 × 10

attributed to the intramolecular hydrogen bonding which is capable of stabilizing the salicylate
ion.

OH H
OH O O
C C O
O +
+ H2O + H3O

The hydrogen atom of the hydroxyl group is sufficiently close to the oxygen atom of the
salicylate ion for it to interact electrostatically with the formation of a six-membered ring. This
is also sometimes referred to as chellation. p-Hydroxybenzoic acid, on the other hand, is unable
to enter into such a ring formation because of the remoteness of the p-OH group. The introduction
of a second OH group in the ortho position results in even greater increase in the value of the
dissociation constant of the acid.3 Because of H-bonding, 2,6-dihydroxybenzoic acid is stronger
than even phosphoric acid.
In many 2-substituted heterocyclic acids the hydrogen bond has the opposite effect, i.e.,
it reduces the strength of the acid. The dissociation constant, for instance, of picolinic acid (1)
is decreased because of the H-bonding taking place between the hydrogen atom of the carboxyl
group and the ring nitrogen atom. The hydrogen bonding stabilizes the acid because the
– COOH group acts as the proton donor and decreases its acidity.4

N O
C

H O
(1)
BASIC CONCEPTS 7

Between keto and enol forms of a molecule, the former is considered more stable. But the
enol form of β-diketones and β-keto esters are stabilized by intramolecular H-bonding and are
stable.
H
O O O O

H H

Enol/Keto > 50
H
O O O O

OCH3 OCH3
Enol/Keto > 29

Intermolecular hydrogen bonding also affects several additional properties such as boiling
point, solubility, vapor pressure of solvents and dielectric constant. As an example consider
ethyl alcohol (bp 78°C) and dimethyl ether (bp –24°C) which have the same molecular formula
C2H6O but different structures. The former, however, has a higher boiling point than the latter.
This difference is explained by the presence of intermolecular H-bond formation, also called
association, of ethyl alcohol in the following manner (2):

C2H5

O
C2H5 H H

O O O
H H H H

O
H5C2 H

O
H H
(2)

Hydrogen bonding is thus expected to cause an elevation of boiling point. Hydrogen


bonding exists in H2O but is absent in H2S, H2Se and H2Te. Nitrobenzene is less soluble in
water than phenol or aniline though nitrobenzene has a large dipole moment, µ= 4.01 D.
Presumably nitrobenzene does not form hydrogen bonds with water as do the other two.
It has been reported that hydrogen bonds exist in carbanions,5 between OH and π electrons,
as in allylic phenols. Hydrogen bonds occupy a significant place in biologically important
molecules such as nucleic acids and proteins.6 A large quantity of water is retained by living
cells in plants and animals and most of it is attached to proteins by hydrogen-bonds.
8 ORGANIC REACTION MECHANISMS

1.5 HYPERCONJUGATION (Interaction Between σ- and π-Systems)

Baker and Nathan, during the reaction of organic halides with pyridine, observed the following
relative rates:
+ –
R CH2Br + N R CH2 N + Br

R H –CH3 –CH2CH3 –CH(CH3)2 –C(CH3)3


Relative rates 1 1.66 1.48 1.34 1.35

The rate was largest when R was a methyl group. Similar results were obtained by other
workers for the solvolysis of p-alkyl substituted benzyl halides, the order of reactivity being:
CH3 > CH2CH3 > CH(CH3)2 > C(CH3)3. This is the reverse of what would be expected if only
the inductive effects were involved.
These workers explained the results by suggesting that the methyl group can donate
electrons through a no-bond resonance or hyperconjugation. In resonance the overlap of π and
lone-pair electrons is involved. In contrast, in hyperconjugation structures are written involving
overlap of σ electrons with the adjacent π-bond as shown for toluene. The methyl group
comprises three σ bonds and can be conjugated with the ring in the following manner:
H H H H H
– + + + +
H C H H C H H C H H C H H C H

– –

Hyperconjugative structures of toluene

A methyl group adjacent to a carbon—carbon double bond acts as an electron donor.


A methyl group is a better electron donor than any of the other alkyl groups by virtue of
the fact that it has more hyperconjugative structures and can thus stabilize the transition state.
Similar structures for the t-butyl group would involve carbon—carbon bond electrons which
are less important. Hyperconjugation thus seems quite valid in explaining the above differences
in reaction rates. A critical analysis of hyperconjugation has been provided by Dewar.7
Like resonance, the concept of hyperconjugation affords adequate physical interpretations
for the behavior of molecules in chemical reactions. As we have seen above, it plays an
important role in reactions involving positive transition states. Hyperconjugative effects also
account for the higher resonance energies of compounds, for instance toluene has a resonance
energy (37.5 kcal/mol) which is 1.5 kcal/mol higher than that of benzene (36 kcal/mol). The
difference is ascribed to hyperconjugation in toluene which makes it more stable. Dipole moment
data support the same general view. Benzene has zero dipole moment whereas toluene possesses
an appreciable value (0.37D). The basicity of di-and tri-methyl substituted benzenes is greater
than toluene for the same reason.
The conjugative ability of a three-membered ring is less than that of a multiple bond. This
is explained on the basis that the dipole moment values of cyclopropyl ketones are between
those of alkyl and vinyl ketones, i.e., a cyclopropyl ketone is less polarized than a vinyl ketone.
BASIC CONCEPTS 9

CH 3 CH 3 CH 3
C O C O C O
(CH 3)2CH CH 2 CH

µ = 2.7 D µ = 2.84 D µ = 2.98 D

Further support for hyperconjugation comes from the carbon— carbon bond distance data.
The carbon— carbon bond distance in propene is 1.488 Å compared to 1.543 Å in ethane and
the C=C bond distance is 1.353 Å compared to 1.334 Å in ethylene. It is reasoned that the
C2— C3 bond acquires some double bond character and the C1— C2 bond a single bond character
because of hyperconjugation. This shows an electron density transfer from C—H to the
π−system with a net strengthening of the bond between C2— C3 but a weakening of the
C1— C2 π bond.
+ +
H H

– 3
2 1 3 2 1 –
H C CH CH2 H C CH CH2

H H
1.448 1.353
Hyperconjugative
structures of propene

There is little doubt about the general usefulness of the concept of hyperconjugation to
explain some characteristics of certain molecules. But it is still a controversial concept and is
used selectively because certain reactions such as aromatic electrophilic substitutions are
inconsistent with this phenomenon.

1.6 INDUCTIVE AND FIELD EFFECTS

A methyl group, when attached to a π-electron system acts as electron donor. The electrons
being transmitted in this case by conjugation. There is, however, another way of transmission
of electrons to the reaction site from a distant substituent. It is a polar effect and takes place
when the substituent attached to the carbon chain is electron-donating or electron-withdrawing
in nature. In the classical sense this takes place via the inductive effect in which the electrical
effects of the group are propagated in a chain by the successive polarization of carbon— carbon
σ bonds. This occurs by virtue of the difference in the electronegativity of the bonded atoms
in the chain.
δδδ+ δδ+ δ+ δ–
C C C X
The argument for this operation is that the atom X of the C — X bond, which is polar
(X may be F, Cl or any other electronegative group), acquires a formal negative charge because
the electron pair is shared more by the electronegative atom. This induces polarization among
the other bonds, and makes the carbon chain to attain a positive charge. This effect diminishes
as the length of the carbon chain increases. The resonance and inductive effects may not
necessarily operate in the same direction. In the ground state these effects are permanent and
are manifested in many properties of the molecules. One of the most ideal cases for correlation
of inductive effects is the solvolysis of 4-substituted bicyclo[2.2.2]octa-1-yl-brosylate (3) in
acetic acid at 75°C. The relative rates are:
10 ORGANIC REACTION MECHANISMS

OBS
(3)
where
O

OBS Br S O

O
R – CH3 – CH2CH3 – CH(CH3)2 – C(CH3)3
–5 –1
10 k sec 3.38 4.04 4.75 6.24
The reaction rates increase progressively from methyl to t-butyl group. This increase was
attributed8 to the inductive electron release by the alkyl groups. The t-butyl group is the most
effective because it releases electrons more efficiently than any other group in this series. It
must be noticed that inductive effect of the alkyl groups is opposite to that of the hyperconjugation
effect where the methyl group is more electron-releasing. The acidic and basic properties of
reagents have been attributed to the electron-releasing or withdrawing inductive effects of the
substituents.
The inductive effect of substituents has a significant effect on the p.m.r. of molecules.
Halogens, for instance, are electron-withdrawing, therefore neighboring hydrogen atoms suffer
a decrease in electron density and as a result their nuclei are deshielded. The order of this
effect parallels the change of electronegativity F > Cl > Br > I of the halogens. The proton thus
resonates at a lower field in the p.m.r.
Recently modified models have been proposed for the inductive effects. From a molecular
orbital approach, Pople and Gordon9 have suggested that the inductive effect of the electronegative
groups results in charge alteration rather than steady charge decay. According to this, the β
carbon atom attains a negative charge, the γ carbon atom attains a positive charge, and so on.
Dewar and Marchand,10 on the other hand, have concluded that the inductive effect is, in fact,
a direct field effect acting across the space by successive polarization of the intermediate bonds.
δδ+ δδ– δ+ δ–
C C C F
γ β α
Polar effects may be directly transmitted through space or through charge-dipole or dipole-
dipole interactions rather than along a carbon chain. Such effects are often called the field
effects. These are long-range polar interactions, i.e., those at more than two carbon— carbon
bond lengths away. Both the inductive and field effects operate in the same direction. It is,
therefore, difficult to separate these two effects. However, an attempt has been made in this
respect. The ionization constants for instance, of polycyclic acids of type (4) were measured11
and the values of pKCl and pKCOOCH were found to be higher than the value of pKH. Electron-
3
withdrawing groups are found to decrease the acidity. On the inductive model an increase in
acid strength should have been expected. The higher the pKa, weaker the acid. The conjugative
and hydrogen-bonding effects can be precluded in this system and the inductive electron-
BASIC CONCEPTS 11

X COOH
X = H, pKH = 6.04
X = Cl, pKCl = 6.25
X = COOCH3, pKCOOCH = 6.20
3
(4)

withdrawing effects of these two groups are also minimal. The difference in the pK values is
explained only by invoking the field effects, i.e., the polar Cl or COOCH3 groups destabilize the
carboxylate ion through space. The Cl group has a better destabilizing effect. Nitration of
triptycene has been explained in terms of the inductive effect.12 The deuterium exchange rate
of fluorobenzenes has been interpreted by a simultaneous operation of both the field and the
inductive effects. It thus follows that a clear-cut separation of the inductive and field effects
has to await further work.

1.7 STERIC EFFECTS

The steric effects arise from the interactions between atoms present in a molecule that are not
bonded to each other. They are also called non-bonded interactions. This happens when bulky
groups are located in the vicinity of the reaction site and the reaction is hindered resulting in
the diminished rate of a reaction. Such a hindrance to reaction was first noted by Kehrman and
this phenomenon was named steric hindrance by Meyer. A vivid illustration of steric blocking
of a reaction with the resultant decrease in reactivity, is provided by the nitration
(HNO3 + H2SO4) of alkyl substituted benzenes. Nitration, which presumably involves the NO2+
ion, yields the ortho- and para-nitro derivatives. The o/p ratio for the following hydrocarbons
in nitration was found to be:
H CH 3 CH 3 CH 3

H C H H C H H C CH 3 CH 3 C CH 3

o/ p ratio 1.57 0.93 0.48 0.21

A drop in the o/p ratio indicates a marked decrease in the formation of the ortho isomer
as one approaches the t-butyl group. This decrease is interpreted in terms of the steadily
increasing steric requirements of the alkyl groups. The size of the substituent group hinders the
approach of the NO+ 2 ion to the ortho position.
Another example that illustrates a decreased reactivity of a molecule involves the
dissociation of carboxylic acids. For example, the dissociation constant value in 50%
H2O — 50% CH3OH of acetic acid is almost 25 times higher than that of methyl iso-butyl
neopentylacetic acid (5).
O C(CH3)3
O
CH3 C OH (CH3)3CCH2 C C
OH
CH3
(5)
–6 –7
Ka 2.7 × 10 1.1 × 10
12 ORGANIC REACTION MECHANISMS

It is apparent that in the latter, large alkyl groups hinder the release of a proton. The role
of steric effects in organic chemistry has been reviewed by Newman.13
The steric effects do not always decrease the rate of a reaction but can often accelerate
it. This happens when the reactants in the ground state are compressed and strain is relieved
in attaining the transition state. This phenomenon is termed steric acceleration and has been
observed in the reactions of flexible aliphatic and alicyclic systems.14 The formation of isobutylene
in the solvolysis of t-butyl chloride has been argued on the basis of relief of strain from the
intermediate t-butylcarbocation. In the rate determining step of the reaction a rehybridization
at the reaction center from sp3 to sp2 occurs. By the same token three and four-membered ring
compounds form open-chain products in deamination and solvolysis reactions requiring low
energies of activation. Opening of a small ring involves the release of ring-strain and an enhanced
rate. Brown and Hammer,15 however, maintain that such a relief to strain may not be possible
in rigid ring systems. In addition to the effect on reaction rates, the steric factors have implications
in organic synthesis as well. They reduce resonance in molecules (steric inhibition of resonance),
play a dominant role in preventing the dimerization of aroxyl radicals and explain the optical
isomerism of diphenyl derivatives.

1.8 BREDT’S RULE

This is an empirical rule deduced from experimental observations by Bredt. According to this
rule, in small bicyclic compounds a C=C bond cannot exist to the bridgehead atoms. Thus
elimination of bromine from bicyclo[2.2.1]hexane (6) gives no bicyclo[2.2.1]hex-1-ene (7).

– +
t-BuO k

t-BuOH
Br
(6) (7)

Compound (7) is highly strained molecule and is incapable of existence* (note the double
bond is cis in the six-membered ring).
Introduction of a carbon—carbon double bond to the bridgehead compresses the molecule
into a planar conformation because of the planar geometry of the sp2 carbons being formed. As
a result the transition state energy for elimination required is considerably high and a small
bicyclic system cannot accommodate the resultant strain. Therefore, the alkene (7) cannot
accommodate a C=C bond and thus very unstable. However, bridgehead alkenes should be
stable when the rings are large enough to accommodate the strain. How large must the ring
systems be before they defy the Bredt’s rule. Fawcett16 defined the limits of the rule and
suggested that the sum (S) of the numbers of atoms S in the bridge of a bicyclic compound
determines whether the ring system is stable or not. He tentatively gave a numerical value of
9 or more to S, i.e., S ≥ 9.
The synthesis of molecules which violates Bredt’s rule has been a challenge to chemists.
Eventually in 1967 the synthesis of bicyclo[3.3.1]non-1-ene (9), from a β-lactone (8) which has
a value of S = 7, was reported simultaneously by two groups of workers.17, 18 Since this
compound is stable, the value of S has to be lowered for defining stability, i.e., S ≥ 7 for stable
ring systems.

*H K Hall, Jr. and A E Shekeil, J. Org. Chem., 45, 5325 (1980).


BASIC CONCEPTS 13

a b c a
b
O
O
(8) (9)
17
Wiseman and Pletcher have suggested that the strain of a bridgehead is closely related
to the strain in trans-cycloalkenes. Therefore, a bicyclic system with a bridgehead double bond
should be isolated, provided the double bond is trans in the larger of the two rings in which
it is endocyclic and has at least eight carbon atoms. In cycloalkene (9), the three rings systems
ab, bc and ac are present. The double bond is trans and endocyclic in the ring ac which also
contains eight atoms and this ring system is thus possible.
Several other bicyclic bridgehead alkenes (10 to 12) with S = 7 have subsequently been
prepared.

(10) (11) (12)

The Bredt’s rule has proved to be of immense value in mechanistic interpretations and in
the assignment of structures. The rule has been successfully employed to account for the
failure of β-keto acids containing the carboxyl group at the bridgehead to decarboxylate; for
alkyl halides which resist elimination; alcohols which do not dehydrate; and for certain ketones
which do not exchange their α-hydrogen for deuterium. Thus bicyclo[2.2.1]heptan-7-one-1-
carboxylic acid (13),19 (with S value of 5) resists decarboxylation even on heating to 500°C. The
reason being that a β-keto acid is believed to decarboxylate through an enol and in the case
O

500°C
No reaction
COOH
(13)

of (13) this involves the formation of a bridgehead enol which is impossible. The bicyclo[5.3.1]
undecan-11-one-1-carboxylic acid (14) on the other hand, which has a value of S = 9,
decarboxylates rather easily even on heating only at 132°C because the ring system can support
a double bond at the bridgehead.
COOH

(14)

The rule applies to heterocyclic systems as well.


14 ORGANIC REACTION MECHANISMS

The cyclic amide (15) shows basic properties (pK a = 5.3) unlike the open-chain
amide.20 This is attributed to the inhibition of a dipolar structure which would unfavorably put
a carbon—nitrogen double bond at the bridgehead, and it will violate Bredt’s rule.

CH3 CH3
CH3 CH3 –
N O N O
+
(15)

For the same reason the following quinclidinone (16) will form an oxime.

O
(16)

The question arises whether or not the rule applies to systems containing atoms other
than carbon and nitrogen, such as sulfur or silicon. Compounds containing such atoms seem
to defy this rule as in the case of bicyclotrisulfone (17). Doering and Levy prepared (17) which
dissolves in aqueous sodium bicarbonate solution indicating that it is acidic in nature. The
H
– –
O
O2S SO2
SO2
O2S SO2 S SO2
SO2 SO2
etc.
+ O
–H

(17) (17a) (17b)

anion formed can be stabilized by resonance as shown in structures (17a) and (17b). The
structurally analogous compound, bicyclo[2.2.2]octan-2,6,7-trione (18), on the other hand, shows
no acidity21 because the corresponding anion cannot be similarly stabilized by resonance.
H

O O
CO C C

(18)

Thus a carbon-sulfur double bond at a bridgehead is possible. Sulfur belongs to the VIB
group of the periodic table and has the following electronic configuration:
1s 2s 2p 3s 3p

In the divalent state, the 3py and 3pz orbitals are filled by electron sharing but there are
five unoccupied d orbitals. It has been argued22 that a sulfur atom utilizes the d orbitals to form
a bond, described as the pπ–dπ bond.
BASIC CONCEPTS 15

1.9 AROMATICITY

We often say that certain organic molecules are aromatic in character whereas others are not.
The simple substance benzene, for instance, is a classical example of the former class of
compounds. Benzene is planar and the π-electrons are delocalized and it is stable. Benzene thus
has aromatic character also known as aromaticity.
Cyclooctatetraene, on the other hand, is non-aromatic because it lacks at least these two
fundamental properties even though it contains alternating single and double bonds. It is found
impossible to construct an unstrained planar model of cyclooctatetraene. Actually
cyclooctatetraene behaves like an unsaturated compound and possesses a non-planar “tub-like”
configuration.

Planar Tub shaped


Cyclooctatetraene

Cyclobutadiene (19) might be expected to be aromatic because like benzene, it also has
resonance structures. However, there is no effective overlap of p orbitals required for aromaticity.
Cyclobutadiene is thus not aromatic rather it is antiaromatic. Apparently there must be some
other way to predict the aromaticity23 of compounds.

(19) (20)

A method to define aromaticity was developed by Hückel24 using the molecular orbital
theory. He found that the relative stability of conjugated monocyclic compounds depended on
the number of electrons in the system. His calculations have been translated into the form of
an expression called the Hückel’s rule. This rule states that monocyclic, conjugated and coplanar
systems that possess (4n + 2)π electrons (where n is an integer) are aromatic. The benzenoid
hydrocarbons fit all criteria of aromaticity, i.e., they have high resonance energies. Applications
of the PMO method also leads to the prediction that the benzenoid hydrocarbons should be
especially stable. Since the inception of this rule the aromaticity of a large number of compounds
has been predicted.25 We shall discuss several examples in the following pages.
n = 0: The simple case of 2π electron system where n = 0 is ethylene but this molecule
is non-cyclic and also non-aromatic. Hückel’s rule predicts that monocyclic ions with (4n + 2) π
electrons will be aromatic, but those with 4nπ electrons will not. The cyclopropenium cation
(21) is cyclic and complies with the Hückel’s rule. The overlap of the three equivalent p-orbitals
(21a) delocalizes the π electrons over the three carbon atoms (21b) to give a stabilized cation.
Thus aromaticity is not restricted to neutral molecules but a number of ions are aromatic as
well. In contrast to (21) the cyclopropenyl anion is non-stabilized by resonance and thus non-
aromatic.
16 ORGANIC REACTION MECHANISMS
+
+ +
H H H H H H H H
AgBF4
+ +
– –
+ BF4–
H Cl H
H – H

(21) (21a) (21b)


n = 1: When n = 1, the cyclic system should have 6π electrons. This is true in the case
of a large number of compounds like benzene, phenol, pyridine, cyclopentadiene anion, etc.
These are aromatic and possess high energies of resonance.
Cyclopentadienyl anion has six π electrons, a negative charge is considered equal to two
electrons and is aromatic. It is stable like other carbanions. It can be prepared by the abstraction
of an acidic proton from cyclopentadiene using a strong base.
H
H H –
– +
(CH3)3O k
+ (CH3)3COH

In cyclopentadiene there is no continuous ring of p-orbitals as one – CH2 group is sp3


hybridized. Deprotonation of – CH2 group leaves an orbital occupied by a pair of electrons. This
orbital can rehybridize to a p-orbital and a ring containing 6π electrons results. The anion now
obeys the Hückel’s rule and it is aromatic.



H (CH3)3CO
H + (CH3)3COH
H

Benzene is a cyclic ring with a continuous ring of overlapping p-orbitals. There are 6π
electrons in benzene. It is a (4n + 2) π system and complies with the Hückel’s rule. It is an
unusually a stable compound.
The Hückel’s rule predicts that the cycloheptatrienylium ion (22) would be stable and
aromatic. This ion, also called the tropylium cation, was prepared by Doering and Knox by the
thermal dissociation of the corresponding bromide and, as predicted, is stabilized by resonance.

+
H
(22) (22a)

Ions with 10π electrons have also been prepared and studied. Dianions of hydrocarbons
are difficult to make but some of them can be easily prepared because they are stable and
aromatic. Katz and Coworkers26 prepared the cyclooctatetraene dianion (23) as a solid by
BASIC CONCEPTS 17

treating cyclooctatetraene with potassium metal in tetrahydrofuran. This provides another


excellent example for the verification of the Hückel’s rule. This also supports the view that
aromatic stabilization leads to usually stable hydrocarbon anions.

THF +
+ 2K –2 + 2K
Reflux –

(23) (23a)

n = 2: For n = 2, the system would have 10π electrons and expected to be aromatic.
Cyclodecapentaene (24) itself could not be prepared because of the large steric interaction

H
H
2
1
3
4
6
5
(24) (25)

between the internal hydrogen atoms at C1 and C6. With the prediction that an addition
methylene group at 1,6-position may make the molecule planar and strain free, Vogel and
Coworkers27 synthesized 1,6-methanocyclodecapentaene (25) which is a conjugated 10π electron
system and it behaves like aromatic compounds. In this molecule, the steric problem has been
avoided with slightly less planarity in the π-system.
n = 3: It is interesting to note that Hückel’s rule applies equally well to certain polycyclic
compounds such as pyrocyclene (26), naphthacene (27) and coronene (28).

(26) (27) (28)

Sondheimer28 has synthesized a new series of conjugated cyclic hydrocarbons which are
called annulenes. The name annulene was also suggested as a general name for monocyclic
compounds containing alternating system of single and double bonds. The ring size of an
annulene is indicated by a number in brackets. According to this nomenclature, benzene
becomes [6] annulene and cyclooctatetraene as [8] annulene. In these molecules the number
of π electrons is equal to the number of carbon atoms in the ring. Large annulenes can attain
planar configurations and are aromatic. Thus cyclooctadecanonane or [18] annulene [29] contains
18 π electrons and obeys Hückel’s rule. It is a planar molecule and the n.m.r. spectrum indicates
that it has a ring current characteristic of aromatic compounds. The resonance energy of (29)
from combustion data has been measured to be 100 kcal/mol.
18 ORGANIC REACTION MECHANISMS

H H

H H

H H H
H H
H H
H H H

H H

H H

(29)

Cyclotetradecaheptaene or [14] annulene (30) contains 14π electrons and is aromatic.29


However, it has been argued that this hydrocarbon must have low resonance energy because
the ring is a bit distorted from planarity. However, the dehydro compound [14] annulene (31)
also containing 14π electrons is planar and aromatic.30

H H H H
H H H H
H
H H H
H H H HH
H
H H H H
H H H H

(30) (31)

The term dehydro implies compounds with greater degree of unsaturation. The triple
bond in these compounds (31) is considered to constitute two π electrons for the Hückel’s rule.
There are, however, compounds which do not contain the benzene ring but are still
aromatic. Such compounds are known as non-benzenoid aromatics.31 Examples of such systems
are azulene (32), N-phenylsydnone (33), pyrrole (34), ferrocene (35), tropolone (36) etc. These
compounds behave like normal aromatic systems.


O –
O

+ N 2+
N N Fe

C6H5 H O
– OH

(32) (33) (34) (35) (36)

Despite the wide utility of the Hückel’s rule a number of systems are known which do
not contain (4n + 2) π electrons but are still aromatic. Well-known examples of these anomalous
cases are heptalene (37), pyrene (38), acenaphthene (39), etc., and they possess considerable
delocalization energies. A set of rules has been proposed to distinguish between the aromatic
and non-aromatic natures of such polycyclic systems.
BASIC CONCEPTS 19

(37) (38) (39)

Cyclic conjugated systems in which electrons are delocalized are aromatic. Breslow32 has
defined a new term called anti-aromaticity to describe conjugated, cyclic and planar systems
in which overlap of p-orbitals is destabilizing. They are thermodynamically less stable than
their corresponding acyclic counterparts. This can be best illustrated by the ions (40) to (42)
all of which contain 4nπ electrons and oxirene (43) and cyclobutadiene. All these species are
destabilized by resonance.
C6H5 C6H5

– O
+ –
C6H5
(40) (41) (42) (43)

Cyclooctatetraene is not antiaromatic, though it has a continuous cycle of 4nπ electrons,


it is not planar. Therefore, it is nonaromatic.
Recently33 a conformational criterion for aromaticity and anti-aromaticity has also been
suggested.

1.10 THE HAMMETT EQUATION

Groups located in a benzene ring or other π-systems have considerable influence on the properties
of the molecule. A substituent can function as a neighboring group, an electron-donor or
acceptor by resonance and inductive effect or alternatively offer steric effect. These effects can
cause changes in the reactivity of molecules. For instance, nitration of phenol in dil nitric acid
yields a mixture of isomeric o- and p-nitrophenols. This is a result of the increase in electron
density on the phenol ring due to conjugative effects. We would like to know as to what part
of the influence is due to each of the effects mentioned above. This was first accomplished by
Hammett.34 He used the ionization of benzoic acid in aqueous solution at 25°C as the reference
reaction and then measured the effect of substituents placed at the m- or p-position on the
observed ionization constant of these substituted benzoic acids.

COOH COO

KO +
+ H2O + H3O


COOH COO

K +
R + H2O R + H3O
20 ORGANIC REACTION MECHANISMS

The basic assumptions were that the steric effect of the m- or p-substituents was absent
and the influence on the ionization was exclusively due to the electronic nature of the groups
R. Hammett thus obtained a quantitative substituent effect denoted by σR (equation 1.2) for the
effect of a given substituent (R) on the ionization of benzoic acids.
FG K IJ
log
HK K
0
= σR ...(1.2)

where, K is the ionization constant of substituted benzoic acid


K0 is the ionization constant of benzoic acid
σ is a constant characteristic of the group R
In order to extend this treatment to other reactions of benzene derivatives such as ester
hydrolysis and hundreds of other related reactions an important assumption was further made
that the effect of a substituent on the reaction rate of any other reaction is proportional to its
effect on benzoic acid ionization, i.e.,
FG k IJ FG K IJ
log
Hk K
0 Any reaction
= ρ log
HK K
0 Benzoic acid
...(1.3)

Equation (1.3) can be written as:


FG k IJ
log
Hk K 0
= ρσ ...(1.4)

Equation (1.4) holds good for other reaction series as well.


where σ is the substituent constant and
ρ is the reaction constant
and k is the rate for any substituted benzene derivative undergoing the reaction
k0 is the rate of reaction for the unsubstituted compound undergoing the same reaction,
such as alkaline hydrolysis of m- and p-substituted methyl benzoates.
In this form equation (1.4) is called the Hammett equation.
For the Hammett equation to be correct, we must use the values of k and k0 under
identical experimental conditions (i.e., solvent, temperature, concentration, etc.). The Hammett
equation is correct because of the presence of linear relationship between the free energies (of
reaction or activation) for different series of reactions.
The expression for the rate of the reactions has the form shown in equations (1.5) and
(1.6) according to the theory of absolute reaction rates (see chapter 2), for the two reactions.

FG KT IJ – ∆G ≠

log k = log
H h K 2.303RT …(1.5)


KT ∆G 0
and log k0 = log – …(1.6)
h 2.303RT
≠ ≠
∆G ∆G0
or log k – log k0 = – +
2.303RT 2.303RT
≠ ≠
k ∆G ∆G 0
or log = – + …(1.7)
k0 2.303RT 2.303RT
BASIC CONCEPTS 21

Compare equations (1.4) and (1.7), we get:


≠ ≠
∆G ∆G0
– = − + ρσ
2.303RT 2.303RT
≠ ≠
or ∆G = ∆G0 – 2.303RT ρσ …(1.8)
∆G0 is the free energy of activation of the reaction of the unsubstituted compound. Similar
to equation (1.8) we can write an equation (1.9) for the free energy of another reaction of a
substituted compound characterized by a different value of the constant:

∆G ′ = ∆G0 ′ – 2.303 RT ρ′σ′


≠ ≠
…(1.9)
Equations (1.8) and (1.9) can be written, to give (1.10) and (1.11) respectively:
≠ ≠
∆G ∆G0
Divide (1.8) by ρ = – 2.303 RT σ …(1.10)
ρ ρ

∆G ′ ∆G0 ′
≠ ≠
Divide (1.9) by ρ′ = – 2.303 RT σ′ …(1.11)
ρ′ ρ′
Subtract these two equations, to obtain (1.12)

∆G 0 ′
≠ ≠
∆G ′
≠ ≠
∆G ∆G0
– = – …(1.12)
ρ ρ′ ρ ρ′
Multiply (1.12) by ρ
ρ
∆G0 ′ = ∆G0 = constant
≠ ≠ ≠
∆G – …(1.13)
ρ′
Equation (1.13) exhibits a linear relationship between the free energies of activation of the
two reactions. In other words, this relationship implies that for any reaction series the values
of the free energies of activation which correspond to a change in reacting molecules, are
linearly related to the values of the free energy of activation of any other reaction series
characterized by the variations in the same parameter.
The Hammett equation describes the influence of polar meta- or para-substituents on the
side chain reactions of benzene derivatives. This equation, however, is not applicable to the
influence of ortho substituents because they exert steric effects and to aliphatic derivatives
because the twisting of the carbon chain may also act sterically. A plot of log k/k0 against σ is
linear with a slope of ρ. The substituent constant σ is defined by equation (1.2)
k
σ = log …(1.2)
k0
where the k’s refer to the ionization constants of benzoic acids. Hammett equation is obeyed
with moderate precision for polar reactions. Equation (1.2) measures the polar effect of the
substituent relative to hydrogen and is independent of the nature of the reaction. Both the
inductive and mesomeric effects are contained in this equation. The values of the substituent
constants σ for selected groups are compiled in Table 1.4. These are based on the ionization
data of benzoic acids.

σ constant
The constant σ is defined by equation (1.2). It is a measure of its capacity to perturb the
environment electronically. According to the standard conditions σ = 0, for the unsubstituted
22 ORGANIC REACTION MECHANISMS

benzoic acid. Also when R = H, then ρ = 1 for the ionization of benzoic acid in water. The
electron density at the reaction center either increases or decreases depending on the nature
of the substituent. It is clear from Table 1.4 that σ values for some groups are negative while
for others they are positive. A negative value for –NH2, –OH, alkyl and others implies an
increase in the electron density at the reaction center. A positive value for – CN, –NO2, halo
groups indicates electron deficiency. These values thus can be used as a measure of the degree
of electron-release or electron-withdrawal by the groups attached to the benzene ring. In other
words, they reflect the degree of change of electron density at the reaction center.
ρ constant
The slope of the Hammett plot gives ρ, the reaction constant. The reaction constant ρ measures
the susceptibility of a reaction to the electronic effects and is reaction independent. The
magnitude of ρ gives a measure of the degree to which the reaction responds to substituents.
From the magnitude and sign of ρ one can make a correct conclusion on the type of transformation
of the initial substrate in the limiting stage of the process. This information offers no possibility
for elucidating the detailed mechanism, however, it is a useful constant. A reaction involving
a positive charge in the transition state will be aided by electron-donating groups and the value
of ρ will be negative. This point may be illustrated by the following example:
.. acetone + –
RC6H4N(CH3)2 + CH3I RC6H4N (CH3)3I
35°C

The value of ρ for this reaction has been observed to be –2.56. A negative value of ρ
indicates that the reaction is aided by electron-donation and vice versa. The sign of ρ is in
accordance with an SN2 displacement involving an activated complex of the following type.
The nitrogen atom must bear a substantial positive charge in the transition state.

H
d+ d–
R C 6H 4(CH 3)2 N C I

H H

A reaction, on the other hand, involving a decrease in the positive charge or increase in
the negative charge in the transition state will be facilitated by electron-withdrawing substituents
and the value of ρ is positive. The magnitude of the value of ρ indicates that the reaction is
sensitive to the polar effects of the substituents and also provides information about the nature
of the transition state involved in the reaction. A very low value of ρ indicates that the reaction
may be insensitive to the effect of groups. Rates of many reactions have been correlated with
the Hammett equation and for many others can be predicted. It is difficult to predict ρ from
experimental conditions since it depends on a number of factors such as the solvent, the nature
of the leaving group, etc. Interposition of a methylene group between the reaction center and
the aromatic ring decreases the value of ρ because the polar effects now have to be propagated
through an additional bond.35
Certain extended equations have been proposed to the original Hammett equation. Jaffe,
for instance, investigated the additive nature of more than one substituent placed on the
aromatic ring. He found that the σ values for various groups can be added and the following
relation, [equation (1.14)] holds good:
log k/k0 = ρ ∑σ …(1.14)
BASIC CONCEPTS 23

where Σσ means the sum of the σ values of all the groups. For compounds containing more than
one benzene ring, equation (1.15) is employed to correlate the results.
log k/k0 = nρσ …(1.15)

TABLE 1.4: Substituent Constant Values for Selected Substituents


+ +
Group σ am σ ap σ bm σ bp σ cl

– O – 0.708 – 0.519
– NMe2 – 0.211 – 0.600 0.10
– N(CH3)2 – 0.161 – 0.660 0.10
– Si(CH3)3 – 0.121 – 0.072 0.011 – 0.021
– C(CH3)3 – 0.120 – 0.197 – 0.059 – 0.256
– OCH3 – 0.115 – 0.268 0.047 – 0.778 0.25
– CH3 – 0.069 – 0.170 – 0.066 – 0.311
– C6H5 – 0.06 – 0.01 – 0.109 – 0.179 0.10
– C2H5 – 0.043 – 0.151 – 0.064 – 0.295
– OH – 0.002 – 0.357 0.25
– H 0 0 0 0 0
+d
– N2 1.7 1.8
– C(CN)e3 1.00 0.01
+
– N(CH3) 3 0.904 0.859 0.359 0.408 0.92
– NO2 0.710 0.778 0.674 0.790 0.63
– CN 0.678 0.628 0.562 0.659 0.56
– SO2CH3 0.647 0.728
– SO2NH2 0.46 0.57
– CF3 0.42 0.54
– Br 0.39 0.232 0.405 0.150 0.45
– Cl 0.373 0.227 0.399 0.114 0.47
– COOH 0.355 0.265 0.421
– CHO 0.355 0.261
– I 0.352 0.276 0.359 0.138 0.38
– F 0.337 0.062 0.352 – 0.073 0.52
– COOCH3 0.315 0.489
– COCH3 0.306 0.516 0.28
f
– C ≡ CH 0.205 0.233 0.334 0.179
g
– C ≡ C–C6H5 0.140 0.165
– C(NO2)3h 0.82

a. H H Jaffe, Chem. Rev., 53, 191 (1953). b. Y Okamoto, T Inukai and H C Brown, J. Am. Chem. Soc., 80, 4979 (1958).
c. C D Ritchie and W F Sager, Progress Phys. Org. Chem., 2, 323 (1964). d. E S Lewis and M D Johnson, J. Am. Chem.
Soc., 81, 2070 (1959). e. J K Williams, E L Martin and W A Shippard, J. Org. Chem., 31, 919 (1966). f. J A Lardgrebe
and R H Rynbrandt, J. Org. Chem., 31, 2585 (1966). g. J Kochi and G Hammond, J. Am. Chem. Soc., 75, 3452 (1953).
h. J Hine and W C Baile, Jr., J. Org. Chem., 26, 2098 (1961).

In rigid aliphatic systems, such as 4-substitutedbicyclo[2.2.2]octan-1-carboxylic acid (44),


the substituents obey the Hammett equation though with a different set of σ values, defined
24 ORGANIC REACTION MECHANISMS

by σ1. The σ1 value represents the electrical effects of a substituent bonded to an sp3 hybridized
carbon atom because the effect is now relayed through σ electrons. Values of σ1 for some groups
are also given in Table (1.4).
COOH
COOH

R
R
(44)

The Hammett equation proved to be the most successful quantitative correlation between
the structure of a compound and the equilibria or the rates of the reactions. However, deviations
from the equation are also observed in several cases. It was found that non-linear plots of the
logarithm of the rate constant against σ were obtained in reactions such as chlorination and
nitration of substituted benzenes and the reaction of benzyl halides with amines i.e., in which
an electron-donating group can mesomerically supply π-electrons and thus has a rate-accelerating
effect on the rate of the reaction. The rate constants for the solvolysis of meta-substituted
phenyldimethylcarbinyl chlorides give a linear plot against σ constants, but the para-substituents
deviate from linearity. Probably the most important reason for this deviation is considerable
resonance interaction (as shown below) between an electron-donating substituent and the
reaction center.
Cl
CH3 + CH3 CH3 CH3
CH3 C CH3 C C

R R R

Different σ values are, therefore, needed to correlate the reactivity of the substituent in
such reactions, Brown and co-workers36 proposed a new set of substituent constants designated
σ+ based on the solvolysis of phenyldimethylcarbinyl chloride in 90% acetone-water at 25°C
as the reference reaction. A reaction constant ρ was derived (equal to – 4.54) from a study of
the m-substituted chlorides. It was thought that resonance contribution for m-substituents
should be negligible and the usual σm values could be employed for correlation of the rate
constant of the above reaction. Therefore, σ+ values were then assigned to other substituents
R from experimental results of kR i.e., the rate constant for the substituents R for the solvolysis
of p-substituted cumyl chlorides using, equation (1.16).
FG k IJ
R
log
Hk K
0
= – 4.54σ+ …(1.16)

where k0 is the value of the rate constant for the parent compound. The value of σ+ are based
on rate and not equilibrium constants. The modified Hammett equation is thus expressed in
the following form, (equation 1.17).
k
log = ρσ+ …(1.17)
k0
BASIC CONCEPTS 25

The values of σ+ for some representative substituents are listed in Table 1.4. It is clear
from Table 1.4 that σ+p differ considerably from σp for highly electron-donating substituents
which reflect a higher degree of resonance between the substituent and the positively
charged reaction center. The correlation of rate data with σ+ values has also been obtained
(see Table 1.5) in a number of cases.
TABLE 1.5: Correlation of Reaction Rates with the Substituent Constants
Reaction Correlation ρ Interpretation Source
with

Hydrolysis of para-substit- σ – 3.6 Electron-donating groups 1


uted methyl benzoates, 95% accelerate reaction,
H2SO4, 25°C acyl-oxygen fission
Hydrolysis of 4-substituted σ + 2.6 Electron-withdrawing 2
2,6-dimethylbenzoates, groups accelerate reaction,
60 percent dioxane, acyl-oxygen fission
40% water, at 82° –174°C
Hydrolysis of 4-substituted σ+ – 3.22 Electron-donating groups 3
2,6-dimethylbenzoates, accelerate the reaction
H2SO4, at 25°C
Hydration of styrenes, σ+ – 3.44 Transition state resembles a 4
HClO4, at 25°C carbocation intermediate
Thermal rearrangement of σ+ – 0.43 Electron-release aids the 5
arylpropargyl ethers reaction
Claisen rearrangement of σ+ – 0.40 Transition state involves a 6
substituted cinnamoyl depletion of electrons
p-tolylethers, at 180°C
Benzylic hydrogen abstraction σ – 0.75 Transition state stabilized 7
from substituted toluenes in by resonance effects
benzene at 39.6°C
Benzylic hydrogen abstraction σ+ – 1.46 A positively charged 8
from substituted toluenes by transition state
trichloromethyl radical, at 150°C
Base catalyzed oxidation of σ + 1.58 Transition state is 9
benzaldehyde with negatively charged.
permanganate
Permanganate oxidation of σ+ – 4.3 Large electron deficiency in 10
furfurals, H2O at 25°C the transition state
Isomerization of cis-cinnamic σ + 1.30 Oxidation faster with 11
acids, H2SO4 at 25°C electron releasing groups
Bromination of acetanilides, σ+ – 12.1 The substituents are capable 12
Br2, HOAc-H2O at 25°C of greater resonance
interactions with the ring
Elimination reaction of 2,2, σ + 10.3 Substantial negative charge 13
2-trifluoroethanesulfonates, develops on the sulfur atom
80 percent ethanol at 25°C in the transition state

Contd....
26 ORGANIC REACTION MECHANISMS

Dissociation of conjugate σ + 0.59 p-Substituents show no 14


acids of benzyl phenyl resonance exaltation
ketones, ethanol at 25°C
Reaction of p-substituted σ – 0.49 Slight development of 15
styrenes with 9-borabicyclo positive charge on the
(3.3.1) nonane α-carbon of styrene.

Source: (1) B M Wepster et al., Rec., Trav., Chim., 88, 301 (1969). (2) H L Goering et al., J. Am. Chem. Soc., 76, 787
(1954). (3) M L Bender and M C Chen, J. Am. Chem. Soc., 85, 37 (1963). (4) W M Schubert et al., J. Am. Chem. Soc.,
86, 4729 (1969). (5) M Harfenist and E Thom., J Org. Chem., 37, 841 (1972). (6) W N White and W K Fife, J. Am. Chem.
Soc., 83, 3846 (1961). (7) R D Gillion and B F Ward. Jr., J. Am. Chem. Soc., 87, 3944 (1965). (8) E S Huyser, J. Am. Chem.
Soc., 82, 394 (1960). (9) K. B. Wiberg and F. Freeman, J, Org. Chem., 65, 573 (2000). (10) D S Noyce and H S Avarbock,
J. Am. Chem. Soc., 84, 1644 (1962). (11) F Freeman et al., J. Org. Chem., 35, 982 (1970). (12) H C Brown and M Dubeck,
J. Am. Chem. Soc., 82, 1939 (1960). (13) R K Grossland, Jr. et al., J. Am. Chem. Soc., 93, 4217 (1971). (14) A Fischer,
et al., J. Am. Chem. Soc., 83, 4208 (1961). (15) L C Vishwakarma and A Fry, J. Org. Chem., 45, 5306 (1980).

It is noticed from Table 1.5 that the reaction involving carbocations usually give large
negative values of ρ and the reactions are facilitated by electron-release.
The Hammett plot (between pkAH versus σ) is also curved for the dissociation of substituted
phenols. Those substituents which have electron-withdrawing properties mesomerically do not
lie on the straight line. In other words certain groups have an added acid-enhancing effect on
phenol than they have on the ionization of benzoic acids. Therefore, for such reactions in
which the reaction center is conjugated with p-substituents which are electron-withdrawing a
new constant σ– was defined based on the following reference reaction, equation (1.18).

OH O

+
+ H2O + H3O ...(1.18)
R R

Values of σ for certain groups are given in Table 1.6.



TABLE 1.6: σ Values for Some Substituents

Substituent σ–

– C6H5 0.08
– C ≡ CH 0.52
– COOH 0.78
– COOC2H5 0.74
– COCH3 0.84
– CHO 1.04
– CN 0.88
– NO2 1.27

Source: O Exner, “A Critical Compilation of Substituent Constants”, Chap X of Correlation Analysis in Chemistry: Recent
Advances, (Eds.) N B Chapman and J Slater, Plenum Press, N.Y. (1978).

The σ– values for groups like –NO2, – CN and – COCH3 are larger than σ. Thus a better
correlation of the pkAH of anilinium cations is obtained with σ– rather than σ.
BASIC CONCEPTS 27

REFERENCES

1. S W Benson, J. Chem. Educ., 42, 502 (1965).


2. G Pimental and A McClellan, The Hydrogen Bond, W H Freeman & Co., San Francisco (1959).
3. I D Sadekov et al., Russ. Chem. Rev., 39, 179 (1970).
4. J H Blanch, J. Chem. Soc., (B), 1937 (1966).
5. L L Ferstanding, J. Am. Chem. Soc., 84, 3553 (1962).
6. A McClellan, J. Am. Chem. Educ., 44, 547 (1967).
7. M J S Dewar, Hyperconjugation, The Ronald Press Co., New York (1962).
8. P Von R Schleyer and C W Woodworth, J. Am. Chem. Soc., 90, 6528 (1968).
9. J A Pople and M Gordon, J. Am. Chem. Soc., 89, 4253 (1967).
10. M J S Dewar and A P Marchand, J. Am. Chem. Soc., 88, 354 (1966); and previous papers.
11. R Golden and L M Stock, J. Am. Chem. Soc., 88, 5928 (1966); L M Stock, J. Chem. Educ., 49,
400 (1972).
12. B H Klanderman and W C Perkins, J. Org. Chem., 34, 630 (1969).
13. M S Newman, (Ed.) Steric Effects in Organic Chemistry, John Wiley and Sons, Inc., New York
(1963).
14. P von R Schleyer and E Wilkott, Tetrahedron Lett., 2845 (1967); P von R Schleyer et al., J. Am.
Chem. Soc., 88, 4475 (166); P von R Schleyer, J. Am. Chem Soc., 86, 1854, 18566 (1964).
15. H C Brown and W J Hammer, J. Am. Chem. Soc., 89, 6378 (1967).
16. E W Fawcett, Chem. Rev., 47, 219 (1950). Also see G Kobrich, Angew. Chem. Int. Edn., (Engl.)
12, 4644 (1973).
17. J R Wiseman and W A Pletcher, J. Am. Chem. Soc., 92, 956 (1970); J R Wiseman, J. Am. Chem.
Soc., 89, 5967 (1967).
18. J A Marshall and H F Faubl, J. Am. Chem. Soc., 89, 5965 (1967).
19. C F H Allen et al., J. Org. Chem. 27, 1447 (1962).
20. H Pracejus, Ber., 92, 988 (1959); H K Hall and A E Shakeil, J. Org. Chem., 45, 5325 (1980).
21. W Theilacker and E Wagner, Ann., 604, 125 (1963).
22. C C Price and S Oae, Sulfur Bonding, the Ronald Press Co., New York (1962).
23. R J Garratt, Aromaticity, McGraw Hill Book Co. (UK) Ltd., London (1971).
24. G M Badger, Aromatic Character and Aromaticity, Cambridge University Press, London (1969);
H F Vulpin, Russ. Chem. Rev., 29, 129 (1960).
25. T J Katz, J. Am. Chem. Soc., 82, 3784 (1960); H L Strauss et al., J. Am. Chem. Soc., 85, 2360
(1963).
26. E Vogel et al., Angew Chem., 76, 786 (1964); E Vogel and W A Boll, Angew. Chem., 76, 784
(1964).
27. F Sondheimer, Acc. Chem. Res., 5, 81 (1972).
28. F Sondheimer and Y Gaoni, J. Am. Chem. Soc., 82, 5765 (1960).
29. Y Gaoni and F Sondheimer, J. Am. Chem. Soc., 86, 521 (1964).
30. J P Snyder (Ed.), Non-benzenoid Aromatics, Academic Press, New York (1969).
31. R Breslow, J. Pure Appl Chem., 28, 111 (1971); R Breslow, J. Am. Chem. Soc., 89, 4383 (1967).
32. D J Raber et al., J. Org. Chem., 53, 2117 (1988).
33. C D Johnson, “The Hammett Equation”, Cambridge University Press, London (1973).
34. A Fisher et al., J. Am. Chem. Soc., 83, 4208 (1961).
35. C W McGraw et al., J. Am. Chem. Soc., 77, 3037 (1955). For a review see L M Stock and H C
Brown, Adv. Phys. Org. Chem., 11, 35 (1963).
28 ORGANIC REACTION MECHANISMS

REVIEW PROBLEMS

1.1. Offer reasons for the following observations:


(a) Nitrobenzene has a larger dipole moment than aniline and phenol but is much less soluble
in water.
(b) Tropolone does not form 2, 4-dinitrophenylhydrazone.
(c) Cyclobutadiene is unstable.
(d) Ionization constant (Ka) of acetic acid is 2.7 × 10–6 in 50 percent methanol-water but that of
triethylacetic acid is 0.36 × 10–6.
(e) The σ value for F is larger than for Cl.
(f) Phosphorus pentachloride has zero dipole moment whereas phosphorus trichloride has a
definite value.
(g) Methoxy group (– OCH3) has σp value of – 0.268.
(h) Deuteration of methyl group of acetophenone lowers the basicity of the molecule.
(i) Treatment of cycloheptatriene with potassium metal dispersed in benzene does not yield the
corresponding cycloheptatrienide anion.
1.2. Predict the approximate magnitude and sign of ρ for the following reactions and draw structures
for the transition state in support of your answer.
OCH3 OCH3

AlCl3
(a) + CH3COCl
C2H4Cl2, 25°C

COCH3

+ –
N(CH3)2 N (CH3)3I

CH3OH, 65°C
(b) + CH3I

CHO CH(OH)CN

90% ethanol, 20°C


(c) + HCN

HS NH
CN C

– C2H5OH
(d) + H2S + OH
Heat

COOH COOCH3
+
H
(e) + CH3OH
25°C

CH(OH)CH2COOH CH CHCOOH
+
H , H2O
(f)
45°C
BASIC CONCEPTS 29

CH2NHCOCH3 CH2N(Cl)COCH3

CH3COOH
(g) + Cl2
180°C

CH2F CH2OH

– –
(h) + OH + F

1.3. The following three nitrophenols have the melting points indicated below each compound. Discuss
this progressive change in their melting points.
OH OH OH
NO2

NO2

NO2
mp 45°C 96°C 114°C
1.4. Write the products of the following reactions. If no reaction takes place write NR.
COOH
COOH
O 200°C
(a) HOOC (b) O
COOH 245°C

O
HOOC O

COOH COOH
O
(c) O 260°C (d) 145°C

1.5. (a) Which compound of the following pairs is a stronger acid?


+ +
(CH3)3 N COOH or (CH3)3 N CH2 COOH

COOH
COOH
or
COOH

COOH

(C6H5)3CH or triptycene

(b) Which compound of the following pair is a stronger base?

or CH 3 CH 3
N CH 3 N CH 3

CH 3 CH 3

NH2 NH2
CH 3 CH 3
or
CH 3 CH 3
NO2 NO2
30 ORGANIC REACTION MECHANISMS

(c) Explain the difference in the magnitude of dipole moments of the following compounds:

(CH 3)2N NO 2 (CH 3)2N CF 3 N(CH 3)2

µ = 6.89 D µ = 4.62 D µ = 1.50 D

(d) Which of the following compounds do you expect to volatalize easily?


OH
NH NH
CH or CH
N N
(e) Explain the difference between the dipole moments of the following two isomeric compounds:
O2 N O 2N

NO 2 H
C C C C
H H H NO 2

µ = 7.38 D µ = 0.50 D
1.6. The value of the rate constant k for the solvolysis of an unsubstituted organic halide is
1.6 × 10–4 s–1 at 40°C and ρ is + 3.78. Calculate the rate constant of the reaction for a
p-bromosubstituted compound if σp– Br is 0.232.
1.7. Could the following two enantiomers be isolated?
Br Br

HOOC COOH and HOOC COOH

Br Br
1.8. Benzenesulfonyl chloride and substituted benzoate anions react in CH3OH at 25°C. The reaction
follows a second order kinetics and the following rate constants were obtained.
O O

x x
C6H5SO2Cl + O C C6H5SO2 O C

X σx 103 k2/l/mol/sec

p – OCH3 – 0.28 6.58


p – CH3 – 1.170 5.75
m – CH3 – 0.06 5.33
H 0 5.18
p – Cl 0.22 4.27
p – Br 0.23 4.04
m – Br 0.39 3.80
m – NO2 0.71 2.91
p – NO2 0.78 2.78

Calculate ρ for this reaction. Compare this value for the same reaction when substituted acetate
anion is replaced by substituted anilines (ρ = – 2.15).
[ J. Org. Chem., 45, 4966 (1980)]
BASIC CONCEPTS 31

1.9. Answer the following and give reasons.


(a) Would you expect compound (A) to show a reduced reactivity towards acetylation?
O

H
(A)

(b) Which of the two (B) or (C) will have slower rate of isomerization?
C 6H 5 O C 6H 5 O
H H
C C 6H 5 or C C 6H 5
C 6H 5 C 6H 5

(B) (C)
(c) Which effect should determine the acidity of the acids (D) and (E) ? Which is more acidic?
COOH COOH

C C

Cl C C

and
Cl

(D) (E)

1.10. Explain the difference in pKa values of the following acids:


CH 3
CH 3

N COOH N COOH N COOH

pKa 4.40 4.60 5.10

1.11. The dipole moment of the following compound has been estimated to be 5.6D. Explain.

S-ar putea să vă placă și