V Aro Hydrocarbons
V Aro Hydrocarbons
V Aro Hydrocarbons
Like all sciences, chemistry has a unique place in our pattern of understanding of the universe. It is the
science of molecules. But organic chemistry is something more. It literally creates itself as it grows.
Of course we need to study the molecules of nature both because they are interesting in their own right
and because their functions are important to our lives. Organic chemistry often studies life by making
new molecules that give information not available from the molecules actually present in living things.
This creation of new molecules has given us new materials such as plastics, new dyes to color our
clothes, new perfumes to wear, new drugs to cure diseases. Organic chemistry started as the chemistry
of life, when that was thought to be different from the chemistry in the laboratory. Then it became the
chemistry of carbon compounds, especially those found in coal. Now it is both. It is the chemistry of
the compounds of carbon along with other elements such as is found in living things and elsewhere.
So, the organic chemistry starts from the chemistry of C atom.
Organic compounds are classified into three major groups:
(1) (a) Aliphatic, open-chain, or acyclic compounds. Methane, ethane, alcohol etc.
(b) Alicyclic compounds. These are carbocyclic or ring compounds which resemble aliphatic
compounds in many ways. Example: cyclohexane, cyclobutane, etc.
(2) Aromatic compounds. These are carbocyclic or ring compounds containing at least one benzene
ring. Benzene and its derivative, naphthalene, etc.
(3) Heterocyclic compounds. These are cyclic (ring) compounds containing other elements besides
carbon in the ring. In a few cases no carbon atom is in the ring. Pyrrole, pyridine, pyparidene etc.
Organic compounds are also classified into two major groups as per the bonding in the compound:
If, in an organic compound containing two or more carbon atoms, there are only single bonds linking
any two adjacent carbon atoms, then that compound is said to be saturated, e.g., ethane, C2H6 (I),
normal propanol, C3H8 (II), acetaldehyde, C2H4 (III). On the other hand, if the compound contains at
least one pair of adjacent carbon atoms linked by a multiple bond, then that compound is said to be
unsaturated, e.g., ethylene, C2H4 (IV); this compound contains a double bond. Acetylene, C2H2 (V);
this contains a triple bond. Acraldehyde, C3H4 (IV); this contains a double bond. The double bond
between the carbon and oxygen atoms is not a sign of unsaturation.
Conjugation
The shapes of molecular orbitals by combining the atomic orbitals of the atoms involved using the
LCAO (Linear Combination of Atomic Orbitals) approach. Hybridizing the atomic orbitals first makes
1 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
this simpler. We mix the 2s orbital on each carbon atom with two of the three 2p orbitals to give three
sp2 orbitals leaving the third p orbital unchanged. Two of the sp2 orbitals overlap with the hydrogen
1s orbitals to form molecular orbitals, which will be the C–H σ bonds. The other sp2 orbital forms the
σ C–C bond by overlapping with the sp2 orbital on the other carbon. The remaining p orbital can
overlap with the p orbital on the other carbon to form a molecular orbital that represents the π bond.
Ethene is chemically more interesting than ethane because of the π bond. In fact, the π bond is the
most important feature of ethene. The C–C π orbital is the HOMO (Highest Occupied Molecular
Orbital) of the alkene, which means that the electrons in it are more available than any others to react
with something that wants electrons (an electrophile). Since this orbital is so important, we will look at
it more closely. The π orbital results from combining the two 2p orbitals of the separate carbon atoms.
Remember that when we combine two atomic orbitals we get two molecular orbitals. These result from
combining the p orbitals either in-phase or out-of-phase. The in-phase combination accounts for the
bonding molecular orbital (π), whilst the out-of-phase combination accounts for the antibonding
molecular orbital (π*). As we progress to compounds with more than one alkene, so the number of π
orbitals will increase but will remain the same as the number of π* orbitals. The π bond contains two
electrons and, since we fill up the energy level diagram from the lowest energy orbital upwards, both
these electrons go into the bonding molecular orbital. In order to have a strong π bond, the two atomic
p orbitals must be able to overlap effectively. This means they must be parallel.
In the dictionary, ‗conjugated‘ is defined, among other ways, as ‗joined together, especially in pairs‘
and ‗acting or operating as if joined‘. This does indeed fit very well with the behaviour of such
conjugated double bonds since the properties of a conjugated system are often different from those of
the component parts. We are using conjugation to describe bonds and delocalization to describe
electrons. The uninterrupted chain of p orbitals is a consequence of having alternate double and single
bonds. When two double bonds are separated by just one single bond, the two double bonds are said to
be conjugated. Conjugated double bonds have different properties from isolated double bonds, both
physically (they are often longer as we have already seen) and chemically.
A -electron conjugated polymer chain contains alternative double bond throughout the chain. All the
carbon atoms or heteroatoms in a conjugated polymer chain are Sp2 hybridized. That means in each
atom one p-orbital (unhybridized) remains perpendicular to the polymer chain and all other p-orbitals
remain parallel to each other. So one p-orbital can laterally overlap to form -orbital with either of the
nearest p-orbital and ultimately the p-orbitals are delocalized through out the polymer chain. In case of
isolated double bond like ethylene two -electrons in atomic orbital (A.O) are localized into the
bonding -molecular orbital (M.O) and anti-bonding -orbital remains vacant. But in case of
conjugated double bond like butadiene the two bonding -orbitals with comparable energies and two
anti-bonding -orbitals with comparable energies are rearranged to two bonding and two anti-bonding
-orbitals due to delocalization. So the energy of delocalized highest occupied molecular orbital
(HOMO) is increased while that of the lowest unoccupied molecular orbital (LUMO) is decreased.
Aromaticity
Let us now return to the structure of benzene. Benzene is unusually stable for an alkene and is not
normally described as an alkene at all. For example, whereas normal alkenes readily react with
bromine to give dibromoalkane addition products, benzene reacts with bromine only with difficulty—
it needs a Lewis acid catalyst and then the product is monosubstituted benzene and not an addition
compound. Bromine reacts with benzene in a substitution reaction (a bromine atom replaces a
2 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
hydrogen atom), keeping the benzene structure intact. This ability to retain its ring structure through
all sorts of chemical reactions is one of the important differences of benzene compared to alkenes and
one that originally helped to define the class of aromatic compounds to which benzene belongs. When
cis-cyclooctene is hydrogenated, 96 kJ mol–1 of energy is released. Cyclooctatetraene releases 410 kJ
mol–1 on hydrogenation. This value is approximately four times one double bond‘s worth, as we might
expect. However, whereas the heat of hydrogenation for cyclohexene is 120 kJ mol–1, on
hydrogenating benzene, only 208 kJ mol–1 is given out, which is much less than the 360 kJ mol–1 that
we would have predicted. This is shown in the energy level diagram below.
The difference between the amount of energy we expect to get out on hydrogenation (360 kJ mol–1)
and what is observed (208 kJ mol–1) is about 150 kJ mol–1. This represents a crude measure of just
how extra stable benzene really is relative to what it would be like with three localized double bonds.
The amount of energy by which benzene is stabilized compared with the hypothetical cyclohexatriene
should be called as delocalization energy or resonance energy.
↔ ↔ ↔
Hückel’s rule
Planar, fully conjugated, monocyclic systems with (4n + 2) π electrons have a closed shell of electrons
all in bonding orbitals and are exceptionally stable. Such systems are said to be aromatic. Now the
name is applied to all the ring systems.
(i) Planarity
(ii) Complete delocalisation of the π electrons in the ring
(iii) Presence of (4n + 2) π electrons in the ring where n is an integer (n = 0, 1, 2, . . .).
This is often referred to as Hückel Rule.
So ar, we have discussed aromatic compounds in which n = 1, i.e., there are 6 π-electrons associated
with the ring. All of these compounds which do not contain a benzene ring are referred to as non-
benzenoid aromatic compounds. Polycyclic aromatic molecules contain in which two or more benzene
ring fused together. Examples are naphthalene, phenanthrene, anthracene etc. There are also non-
benzenoid aromatics in which n = 2, e.g., azulene (cf. the cyclo-octatetraene dianion, above). The 5-
membered ring has 5 and the 7-membered ring has 7 re-electrons. Because of the tendency for each
ring to acquire a closed shell of 6 re-electrons, one electron passes from the 7-ring to the 5-ring, and
now the molecule has a dipolar structure, each ring having an aromatic sextet.
4 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
Anti-aromatic and non-aromatic
Anti-aromatic molecules are cyclic, conjugated, have (4n) pi electrons, and are flat. This unusual
instability is called ―anti-aromaticity‖. For instance, epoxidation of acetylene would give the molecule
below (oxirene). Oxirene itself has never been observed, although there are tantalizing traces of its
fleeting existence. And by the way, neither has the nitrogen analogue, 1H-azirene, or the thiirene. The
fleeting existence of a 1H-azirine has been postulated in the addition of a nitrene to an acetylene,
which quickly rearranges to a (stable) 2H-azirene (below). Yes, it‘s true that the lone pair in azirene
can be put into an sp3 hybrid orbital (below). This likely reduces the anti aromatic ―penalty‖
somewhat, but the molecule is still incredibly unstable. Anti-aromaticity is the simplest way to account
for this. Cyclobutadiene looks like a simple enough molecule, but it wasn‘t actually stable at
temperatures above 35 Kelvin. Sure, it‘s a four-membered ring, and yes, it has a lot of ring strain, but
more strained molecules have been made that are actually stable at room temperature. You might also
note that like the above examples, cyclobutadiene is another example of a molecule that is cyclic,
conjugated, has 4 pi-electrons, and is flat. What‘s even more interesting is what was learned
about the geometry of cyclobutadiene. Rather than being a molecule with identical bond lengths (like
benzene), cyclobutadiene was found to have a rectangular shape, indicating that the electrons
were not delocalized. Cyclooctatetraene is anti-aromatic only if it is flat. However, the relatively
―floppy‖ structure of cyclooctatetraene allows for some flexibility. The bonds can rotate away from
flatness such that the molecule adopts a ―tub-like‖ shape, thereby avoiding the ―antiaromaticity tax‖ of
18 kcal/mol that would be paid if all the p-orbitals on the molecule were conjugated with each other.
Each of them is cyclic, conjugated, and flat – and when you count the number of pi electrons, it‘s
multiples of 4. The molecule below is called ―Pentalene‖. It has been synthesized, but is only stable
below –100 °C.
The answer is that reaction A) happens faster. Reaction B does not happen at all. But unstable it is!
The cyclopentadienyl cation is incredibly unstable and difficult to make – and it‘s not for lack of
trying. There‘s something very strange about the structure of the cyclopentadienyl cation that gives it
unusual instability.
This means that we can now draw up three categories for molecules according to the following
criteria. Aromatic molecules are cyclic, conjugated, have (4n+2) pi electrons, and are flat. Anti-
aromatic molecules are cyclic, conjugated, have (4n) pi electrons, and are flat. Non-aromatic
molecules are every other molecule that fails one of these conditions.
Problem#1: Epoxidation on ethene (aka ethylene) is done to the tune of 15 million tons, annually.
Epoxidation of acetylene to form oxirene is unknown.
5 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
Electrophilic aromatic substitution: The arenium ion mechanism
The structure of benzene and, in particular, to its delocalized π orbitals; the concentration of negative
charge above and below the plane of the ring-carbon atoms might be expected to shield the ringcarbon
atoms from the attack of nucleophilic reagents and, by contrast, to promote attack by cations, X+
electron-deficient species, i.e. by electrophilic reagents; this is indeed found to be the case.
Electrophilic aromatic substitutions are unlike nucleophilic substitutions in that the large majority
proceed by just one mechanism with respect to the substrate. In this mechanism, which we call the
arenium ion mechanism, the electrophile (which can be viewed as a Lewis acid) is attacked by the p-
electrons of the aromatic ring (behaving as a Lewis base in most cases) in the first step. This reaction
leads to formation of a new C-X bond and a new sp3 carbon in a positively charged intermediate
called an arenium ion, where X is the electrophile. The positively charged intermediate (the arenium
ion) is resonance stabilized, but not aromatic. Loss of a proton from the sp3 carbon that is ‗‗adjacent‘‘
to the positive carbon in the arenium ion, in what is effectively an E1 process, is driven by
rearomatization of the ring from the arenium ion to give the aromatic substitution product. A proton
therefore becomes the leaving group in this overall transformation, where X replaces H.
The electrophile may be a positive ion or be a molecule that has a positive dipole. If it is a positive ion,
it is attacked by the ring (a pair of electrons from the aromatic sextet is donated to the electrophile) to
give a carbocation. Ions of this type are called Wheland intermediates, σ complexes, or arenium ions.
The inherent stability associated with aromaticity is no longer present in the intermediate, but the ion
is stabilized by resonance. For this reason, the arenium ion is generally a highly reactive intermediate,
although there are cases in which it has been isolated. It might be expected that the first phase of
reaction would be interaction between the approaching electrophile and the delocalised π orbitals and,
in fact, so-called π complexes are formed. Thus toluene forms a 1:1 complex with hydrogen chloride
at - 7 8 °, the reaction is readily reversible. That no actual bond is formed between a ring-carbon atom
and the proton from HCl is confirmed by repeating the reaction with DCl; this also yields a π complex,
but its formation and decomposition does not lead to the exchange of deuterium with any of the
hydrogen atoms of the nucleus, showing that no C—D bond has been formed in the complex. In the
presence of a compound having an electron-deficient orbital, e.g. a Lewis acid such as A1C13, a
different complex is formed, however. If DCl is now employed in place of HCl, rapid exchange of
deuterium with the hydrogen atoms of the nucleus is found to take place indicating the formation of a
a complex (II) in which H+ or D+, as the case may be, has actually become bonded to a ring-carbon
atom. The positive charge is shared over the remaining five carbon atoms of the nucleus via the n
orbitals and the deuterium and hydrogen atoms are in a plane at right angles to that of the ring.
Thus the formation of former leads to no colour change and but little difference in absorption
spectrum, indicating that there has been practically no disturbance in the electron distribution in
toluene; while if AlCl3 is present the solution becomes green, will conduct electricity and the
absorption spectrum of toluene is modified, indicating the formation of a complex such as (II) as there
is no evidence that aluminium chloride forms complexes of the type, H+AlCl4-. The reaction may be
completed by A1C14- removing a proton from the σ complex. This can lead only to exchange of
hydrogen atoms when HCl is employed but to some substitution of hydrogen by deuterium with DCl,
i.e. the overall process is electrophilic substitution. In theory, as an alternative, react by removing Cl-
from A1C14- resulting in an overall electrophilic addition reaction, as happens with the π orbital of a
simple carbon-carbon double bond; but this would result in loss of tffe stabilisation conferred on the
molecule by the presence of delocalised n orbitals involving all six carbon atoms of the nucleus, so
that the product, an addition compound, would no longer be aromatic with all that that implies. By
expelling H+, i.e. by undergoing substitution rather than addition, the complete delocalised n orbitals
are regained in the product and characteristic aromatic stability recovered:
6 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
The evidence for the arenium ion mechanism is mainly of two kinds:
1. Isotope effect. Evidence for the arenium ion mechanism has also been obtained from other kinds of
isotope-effect experiments, involving substitutions of the type; where M is Si, Ge, Sn, or Pb, and R is
methyl or ethyl. In these reactions, the proton is the electrophile. If the arenium ion mechanism is
operating, then the use of D3O+ should give rise to an isotope effect, since the D–O bond would be
broken in the rate-determining step. Isotope effects of 1.55–3.05 were obtained,11 in accord with the
arenium ion mechanism.
2. Isolation of arenium ion intermediates. Very strong evidence for the arenium ion mechanism comes
from the isolation of arenium ions in a number of instances. For example, the folowing intermediate
was isolated as a solid with a melting point of -15°C from treatment of mesitylene with ethyl fluoride
and the catalyst BF3 at -80°C. Formation of stable π-complex with the reagent (e.g., I2, Br2, picric
acid, HCl, Ag+ etc.) is also the separable at lower temperature and to be consider as evidance of
arenium ion intermediate as it is formed after the formation of π-complex.
Energy profile diagrams. In the face of the concentration of negative charge presented to an attacking
reagent it might be expected that the substitution of benzene by the common electrophiles (i.e.
halogehation, nitration, sulphonation and the Friedel-Crafts reaction) would be extremely easy.
Though the electrophilic substitution of bezene is not difficult, that it is not easier than it is, is due to
the enfrgy barrier to be surmounted in converting the very readily formed π complex to a σ complex in
which actual bonding of the reagent to a ring-carbon atom has taken place. For in the π complex, the
aromatic nature of the nucleus (i.e. the delocalised n orbitals) is largely undisturbed, while in the σ
complex some of the characteristic stabilisation has been lost as the orbitals now only involve five
carbon atoms. Electrophilic aromatic substitution has two steps (attack of electrophile, and
deprotonation) which each have their own transition state. There is also a carbocation intermediate.
This means that we should have a ―double-humped‖ reaction energy diagram. Second, the relative
heights of the ―peaks‖ should reflect the rate-limiting step.
If the attacking species is not an ion, but a dipole, the product must have a negative charge unless part
of the dipole, with its pair of electrons, is broken off somewhere in the process, as in the conversion of
the following. Note that when the aromatic ring attacks X, Z may be lost directly to give the product.
7 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
Halogenation
In direct halogenation, low temperature and the presence of a halogen carrier favour nuclear
substitution. Chlorination and bromination may be very conveniently carried out at ordinary
temperature in the presence of an iron or aluminium amalgam catalyst; the extent of the substitution
depends on the amount of halogen used, e.g., chlorobenzene is formed when benzene is treated with
chlorine (1 molecule) in the presence of iron. If 2 molecules of chlorine are used, then a mixture of 0-
and p-dichlorobenzenes is obtained, the latter predominating. When toluene is brominated in the
presence of iron (using i molecule of bromine), a mixture of o- and p-bromotoluenes (tolyl bromides)
is obtained.
A possible mechanism for the chlorination of benzene in the presence of ferric chloride is as follows.
Nuclear chlorination can be carried out with sulphuryl chloride in the presence of a catalyst, a most
effective one being a mixture of sulphur monochloride (S2Cl2) and aluminium chloride (AlCl3); e.g.,
sulphuryl chloride (SO2Cl2) in the presence of 1 per cent, of this catalyst chlorinates benzene in the
cold. Toluene can be similarly chlorinated, and the side-chain is not attacked in the absence of organic
peroxides. Chlorination with sulphuryl chloride is stepwise, and the final product depends on the
amount of this reagent used.
Nitration
1. Aromatic nitrocompounds are almost invariably prepared by direct nitration, using one of the
following reagents:
(i) Concentrated nitric acid, density about 1-5.
(ii) Fuming nitric acid (6-12 per cent, nitrogen dioxide).
(iii) Mixed acid. This is a mixture of nitric acid (concentrated or fuming) and various amounts of
concentrated sulphuric acid (sometimes fuming sulphuric acid is used). Mixed acid is by far the most
important nitrating agent. Occasionally other acids are used, e.g., glacial acetic acid.
The active nitrating agent is the nitronium cation, N02+, is formed as follows;
Boron trifluoride is a very effective catalyst for nitration. One equivalent of catalyst must be used. The
catalytic effect of boron trifluoride is believed to be due to the formation of the nitronium ion:
The yields are better and the products purer than by the above methods. Furthermore, boron trifluoride
is particularly useful for nitrating compounds containing a negative group. Olah et al. have shown that
nitronium tetrafluoroborate (a stable compound) is a useful nitrating agent. This method of nitration
gives direct preparative proof of the electrophilic character of nitration through the nitronium cation.
8 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
Nitration (with mixed acid) thus possibly proceeds as follows:
Sulphonation
Aromatic sulphonic acids are usually prepared by direct sulphonation, since this is far more
convenient than indirect methods. The usual sulphonating agents are;
(i) Concentrated sulphuric acid (98 per cent.) ; e.g., benzene readily forms benzenesulphonic acid
(note the reversibility of the reaction).
(ii) Sulphur trioxide in an inert solvent such as sulphuric acid, i.e., oleum [fuming sulphuric acid), or
as an addition product with pyridine or dioxan. Oleum with a free sulphur trioxide content up to about
70 per cent, is particularly useful for those cases where sulphonation is difficult, e.g., the sulphonation
of compounds containing m-orienting groups in the ring (nitrocompounds, sulphonic acids, etc.).
(iii) Chlorosulphonic acid. This results in the formation of either a sulphonic acid by carrying out the
reaction in carbon tetrachloride solution (using one molecule of reagent), or a sulphonyl chloride
(using excess of reagent).
(iv) Sulphuryl chloride in the presence of aluminium chloride sulphonates aromatic compounds in the
cold to form a sulphonyl chloride.
Many mechanisms have been suggested for sulphonation with sulphuric acid or sulphur trioxide, but
none is certain. It is certain, however, that sulphonation with these reagents is reversible. The
difficulty in ascertaining the mechanism appears to be that it is not completely clear what is the active
sulphonating species. However, later experiments have shown that this mechanism is unacceptable;
the S03H+ ion is not believed to be the active species with sulphonations in oleum. These authors
conclude that sulphur trioxide is the active species. An interesting point here is that Gold et al. (1956)
have proposed the following mechanism in aqueous sulphuric acid.
The mechanisms of sulphonation given above are all electrophilic reactions. It is also possible that the
following electrophilic reaction could occur. As the experiment stands, there is no way of telling
whether this reaction occurs or not. It was, however, shown to occur by using deuterosulphuric acid,
D2S04, the final product being hexadeuterobenzene
When a compound containing an o-p-orienting group is sulphonated with sulphuric acid or oleum, the
temperature at which the reaction is carried out affects the ratio of the 0- and p-isomers. Generally,
lower temperatures favour o-substitution and higher temperatures, p-substitution; both isomers,
however, are always obtained. Lauer (1935) found that the presence of water also influenced the o-p
ratio. When concentrated sulphuric acid (98 per cent.) is used, the product is mainly a mixture of the
0- and ^-isomers, together with a small amount of the m-isomer; or vice versa. When sulphur trioxide
(gaseous or in oleum) is used, 100 per cent. o-p- or m-substitution is obtained.
9 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
Very few catalysts are known for sulphonation; the salts of mercury, silver and vanadium, and iodine,
seem the best. It has also been found that a mercury catalyst may change the orientation of
sulphonation. Chlorosulphonic acid, at low temperatures, usually gives a high yield of the o-isomer
when the compound contains an o-p-orientmg group. An example of indirect sulphonation is the
replacement of an "activated" halogen atom by sulphonic acid group, e.g., o-chloronitrobenzene reacts
with sodium sulphite to form o-nitrobenzenesulphonic acid.
Friedel-Craft’s alkylation/acylation
This reaction involves the introduction of an alkyl or acyl group into the benzene ring in the presence
of a catalyst.
The aromatic compounds may be hydrocarbons, aryl chlorides and bromides, mono- and polyhydric
phenols or their ethers, amines, aldehydes, acids, quinones, and certain derivatives of heterocyclic
compounds.
The alkylating agents may be alkyl halides, aliphatic alcohols, olefins, ethers and alkyl esters of
organic and inorganic acids. From the point of view of convenience, the alkylating agent is usually
confined to alkyl halides, alcohols and olefins.
The acylating agents may be acid chlorides or anhydrides, acids, or esters.
Many catalysts may be used, e.g., the chlorides of aluminium, iron (ferric), zinc, tin (stannic); boron
trifluoride, hydrogen fluoride, sulphuric acid, phosphoric acid, and a mixture of silica and alumina.
Acylations may also be affected in the presence of perchloric acid as catalyst. Recently it has been
found that benzene will react with an alkyl halide or acyl chloride in the absence of a catalyst, but in
this case the reaction must be carried out under pressure.
Of all the catalysts mentioned, aluminium chloride (the one originally used by Friedel and Crafts) is
the best, and gives satisfactory yields when the alkylating agent is an alkyl halide, alcohol or an olefin.
The amount of catalyst required depends on the nature of the alkylating agent used, e.g., with alkyl
halides or olefins, about 0-2-0-4 molecule of aluminium chloride (using the formula A1C13) is
necessary. With alcohols (ethers, etc.) a larger amount of catalyst (1 or more molecules) is necessary.
It appears that a trace of water is necessary to catalyse this reaction, and based on this it has been
suggested that the actual alkylating agent is the alkyl chloride which is produced as follows.
The orientation of the products in the alkylation of an alkylbenzene with aluminium chloride as
catalyst depends on the temperature of the reaction. The reason for this is not certain. One possibility
is that the 0- and pcompounds are the kinetically controlled products, whereas the m-compound is the
thermodynamically controUed product. The Friedel-Crafts reaction is reversible, and consequently at
high temperature the product formed will be the thermodynamically controlled one. On the other hand,
boron trifluoride, which is very useful with alcohols and olefins (as is hydrogen fluoride), the main
product of disubstitution is the p-derivative; sulphuric acid also produces mainly the p-derivative.
Boron trifluoride does not catalyse alkylations with alkyl chlorides or bromides, but the reaction can
be carried out in the presence of water or alcohol. On the other hand, alkylation with alkyl fluorides is
readily carried out with boron trifluoride as catalyst.
10 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
The Friedel-Crafts reaction is often the most useful method of introducing an alkyl group into the
benzene ring. The ease of alkylation with an alkyl halide depends on the nature of the alkyl radical and
the halogen atom. The ease of alkylation for a given halogen atom is tertiary
halide>secondary>primary; and for a given alkyl radical, alkyl fluoride>chloride>bromide>iodide
(this is the reverse order of the usual reactivity of alkyl halides). The Friedel-Crafts reaction is usually
carried out in the presence of a solvent, but if one of the reactants is a liquid hydrocarbon, e.g.,
benzene, this may be used as solvent as well. The solvents generally used are nitrobenzene, light
petrol, or carbon disulphide, and the solvent sometimes affects the orientation. At the end of the
reaction the complex is usually decomposed by ice-cold concentrated hydrochloric acid. The Friedel-
Crafts alkylation reaction has certain drawbacks;
(i) Problem of rearrangement. The structure of the alkyl group plays a part in the alkylation, e.g., it is
easy to introduce a methyl, ethyl or isopropyl group, but usually difficult to introduce a n-propyl or n-
butyl group, since these tend to rearrange to the iso-radicals. isoButyl halides very readily give a tert-
butyl substitution product. If the reaction is carried out in the cold with benzene and w-propyl
chloride, the n-propyl radical is introduced. If, however, the reaction is carried out at higher
temperatures, the isopropyl radical is mainly introduced. The catalyst itself also affects isomerisation,
e.g., n-alcohols usually alkylate at low temperatures without rearrangement taking place when
aluminium chloride is used, but rearrangement occurs when boron trifluoride or sulphuric acid is used,
a primary alcohol giving rise to a secondary alkyl radical, and a secondary alcohol to a tertiary alkyl
radical. A very good means of introducing the w-propyl radical is to use cyclopropane as the
alkylating agent.
(ii) Problem of over-alkylation or polyalkylation. It is not always possible to stop the reaction at the
required stage, i.e., there is always a tendency to over-alkylate. Over-alkylation, due to the presence of
the first alkyl group in the ring increasing the ease of further alkylation, may be partly prevented by
using a large excess of the hydrocarbon.
iii) Problem of reversibility. The Friedel-Crafts reaction is reversible, i.e., an alkyl group may be
removed, especially at high temperatures. This renders the structure of the product uncertain in a
number of cases.
(iv) Problem of deactivation. The presence of a negative group (m-orienting group) in the ring hinders
or inhibits the Friedel-Crafts reaction, e.g., nitrobenzene, C6H5-N02, and acetophenone,
C6H5COCH3, do not undergo the Friedel-Crafts reaction. On the other hand, if a strongly activating
group (o-p-orienting group) is present in either of the above two compounds, reaction can take place,
e.g., o-nitroanisole reacts with isopropanol in the presence of hydrogen fluoride to form 2-nitro-4-
isopropylanisole. This hindering effect can be used to advantage for preparing a monoalkylated
benzene (free from dialkylated-product).
(v) Problem of catalyst deactivation. For phenols and acids it is better to use boron trifluoride than
aluminium chloride, since the latter forms aluminium salts, thereby necessitating the use of a large
excess of aluminium chloride
The generally accepted mechanism of acylation, until recently, was an ionic one involving the
intermediate formation of an acyl cation. According to Brown et al., this ionic mechanism cannot
operate in the case of toluene, which acylates in the p-position. These authors suggest a substitution
mechanism involving a " larger " attacking reagent to account for the steric requirements. This reagent
11 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
may be the complex R-COX-AlX2 itself or a solvated complex. Thus the following mechanism has
been proposed:
Reactive aromatic hydrocarbons are acylated by both processes simultaneously, but less activated
hydrocarbons proceed through more of the bimolecularmechanism, and benzene itself almost
exclusively through the bimolecular mechanism.
Any group Z that has an electron-donating field effect (+R, +I, Z will have a –ve charge or a δ- dipole
in most cases) should stabilize all three ions (relative to substrate), since electron donation to a
positive center is stabilizing. Formation of a stabilized ion should be faster than benzene (which
generates substrate), or activating. On the other hand, electron-withdrawing groups (-R, -I, Z will have
a +ve charge or a δ+ dipole in most cases) will increase the positive charge on the ring (like charges
repel), and destabilize the arenium ion. Formation of a destabilized ion should be slower, or
deactivating. On the basis of these discussions, we can distinguish following types of groups.
12 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
1. Groups that contain an unshared pair of electrons on the atom connected to the ring are strongly
activating, e.g., NR2, NHR, NH2, and OH, which are strongly activating, as is O-. The groups like
NHCOR, OCOR, SR etc. are mildly actevating by weak electron pair donation. The groups that
contain an unshared pair, the ortho and para ions are more stable than either the meta ion or the
unsubstituted ion and they are ortho para orienting.
2. The halogens deactivate the aromatic ring to substitution (the rate of reaction is slower than that of
benzene), and this effect may arise from the unique energy level of the halogen lone-pair orbital,
which is higher than the adjacent π-molecular orbital of benzene. Fluorine is the least deactivating (as
+R=-I), and fluorobenzenes usually show a reactivity approximating that of benzene itself. The other
three halogens make the ortho and para ions more stable than the meta, but less stable than the
unsubstituted arenium ion means deactivating but ortho, para orienting.
3. Groups that lack an unshared pair on the atom but with –I effect connected to the ring and that are
activating as well as ortho–para directing. In this category are alkyl groups, aryl groups, and the COO-
group etc. With a pair of electrons in benzene substituent from the aromatic sextet playing the part
played by the unshared pair. Without having resonance interctaion, the effect of negatively charged
groups like COO- is easily explained by the electron withdrawing field effect. The effect of alkyl
groups can be explained in the same way, but, in addition, we can also draw canonical forms of
hyperconjugation, even though there is no unshared pair.
4. Groups that lack an unshared pair on the atom connected to the ring and that are -I, in approximate
order of decreasing deactivating ability, NR3+, NO2, CF3, CN, SO3H, CHO, COR, COOH, COOR,
CONH2, CCl3, NH3+, SR2+, PR3+ (having positive charge on the group) etc. The field-effect
explanation predicts that these should all be meta directing and deactivating, except NH3+.
5. The NH3+ group is an anomaly, since this group directs para about as much as or a little more than
it directs meta. The NH2Me+, NHMe2+, and NMe3+ groups all give more meta than para
substitution, the percentage of para product decreasing with the increasing number of methyl groups.
Ortho/para ratio.
1. It is usually difficult to predict how much of the product will be the ortho isomer and how much the
para isomer for ortho-para orenting groups. On a purely statistical basis there would be 67% ortho and
33% para, since there are two ortho positions and only one para. However, the phenonium ion which
arises from protonation of benzene, has the approximate charge distribution shown in the figure. If
other effects are absent, this would mean that >33% para and <67% ortho substitution would be found.
For example, chlorination of toluene gives an ortho/para ratio anywhere from 62:38 to 34:66.
2. The meta-directing groups, which destabilize a positive charge, give ortho/para ratios >67:3350 (of
course the total amount of ortho and para substitution with these groups is small, but the ratios are
generally >67:33).
3. Another important factor is the steric effect as some groups are so large that they direct almost
entirely para. An example may be seen in the nitration, under the same conditions, of toluene and tert-
butylbenzene. The former gave 58% of the ortho compound and 37% of the para, while the more
bulky tert-butyl group gave 16% of the ortho product and 73% of the para.
4. That a steric factor is not the only one at work, however, is seen in the nitration of fluoro-, chloro-,
bromo and iodobenzenes where the percentage of o-isomer obtained increases as we go along the
series, despite the increase in size of the substituent. This is due to the fact that the electron-
withdrawing inductive effect influences the adjacent o-positions much more powerfully than the ihore
distant p-position. The inductive effecrtfecreases considerably on going from fluoro- to Iodo benzene
(the biggest change being seen in going from fluoro- to chlorobenzene) resulting in easier attack at the
o-positions despite the increasing size of the group already present.
5. When the ortho–para-directing group is one with an unshared pair (this of course applies to most of
them), there is another effect that increases the amount of para product at the expense of the ortho. A
13 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
comparison of the intermediates involved shows that canonical form with a para-quinoid structure are
more stable than an ortho-quinoid structure. Therefore para intermediate contributes more to the
hybrid and increases its stability compared to the ortho intermediate.
6. The o/p-ratio is also a good deal influenced by the actual conditions under which substitution is
carried out, and there are a number of anomalies that have not yet been adequately explained.
If two substituents are already present in a benzene nucleus, the position of entry of a third can, in a
number of cases, be forecast with fair accuracy.
(i) When both groups belong to class I, the directive power of each group is in the following order
(almost as per their electron donating nature). If the p-position is unoccupied, then generally this
position is entered preferably to the o-, i.e., more of the p-isomer is formed than the o-,
(ii) When both groups belong to class II, then it is difficult to introduce a third group, and the directive
power of each group is as follows. The directive powers given above are due to Holleman, but study of
more recent literature appears to show lack of agreement on this problem. However, it appears the
order of w-directing groups is probably the reverse of that given above, i.e., it appears to be follows.
(iii) When the two groups direct differently, i.e., belong to classes I and II, then class I takes
precedence. Furthermore, if the orientations reinforce each other, the third group enters almost entirely
one position. Thus if an o/p- and a m-directive substituent are present, as in m-nitrotoluene, we should
expect nitration to take place at the positions indicated by arrows. That is o- and p- to the activating
substituent, Me, but not m- to the deactivating substituent, NO2. This is borne out in practice, i.e.
where an o/p- and a m-directive substituent are in competition the latter can often be looked upon as
merely occupying a position in the nucleus; though any possible steric effects it may exert must also
be taken into account in deciding which positions, out of several alternatives, are likely to be most
readily attacked. With two suitably situated o/p-directive substituents, however, actual competition
does take place. It is not always possible accurately to forecast the outcome, but normally those groups
that exert their effects via unshared electron pairs are more potent than those operating via inductive or
hyperconjugative effects, possibly due to the added electromeric effect exerted on approach of the
electnaphile. Thus nitration of acet-p-toluidide leads to virtually no attack at all taking place o- to Me.
When the groups oppose each other, predictions may be more difficult. In a case such as where two
a) If a strong activating group competes with a weaker one or with a deactivating group, the former
controls. Thus o-cresol gives substitution mainly ortho and para to the hydroxyl group and not to the
methyl. For this purpose we can arrange the groups in the following order: NH2, OH, NR2, O- > OR,
OCOR, NHCOR > R, Ar > halogen > meta-directing groups.
b) All other things being equal, a third group is least likely to enter between two groups in the meta
relationship. This is the result of steric hindrance and increases in importance with the size of the
groups on the ring and with the size of the attacking species.
c) When a meta-directing group is meta to an ortho–para-directing group, the incoming group
primarily goes ortho to the meta-directing group rather than para. For example, chlorination of 3-
chloro-nitrobenzene gives mostly 3,6 dichloro-nitrobenzene. The importance of this effect is
underscored by the fact that the middle product, which is in violation of the preceding rule, is formed
in smaller amounts, but the last one is not formed at all. This is called the ortho effect, and many such
examples are known. Another is the nitration of p-bromotoluene, which gives 2,3-dinitro-4-
bromotoluene. In this case, once the first nitro group came in, the second was directed ortho to it rather
than para, even though this means that the group has to come in between two groups in the meta
position. There is no good explanation yet for the ortho effect, though possibly there is intramolecular
assistance from the meta-directing group.
14 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi
8. In fused ring systems, the positions are not equivalent and there is usually a preferred orientation,
even in the unsubstituted hydrocarbon. The preferred positions may often be predicted as for benzene
rings. Thus it is possible to draw more canonical forms for the arenium ion when attack by
naphthalene leads to attachment of the electrophile at the α position than when attack by naphthalene
leads to attachment of the electrophile at the β position. Therefore, the a position is the preferred site
of attachment, though, as previously mentioned, the isomer formed by substitution at the β-position is
thermodynamically more stable and is the product if the reaction is reversible and equilibrium is
reached. Because of the more extensive delocalization of charges in the corresponding arenium ions,
naphthalene is more reactive than benzene and substitution is faster at both positions. Similarly,
anthracene, phenanthrene, and other fused polycyclic aromatic hydrocarbons are also substituted faster
than benzene. Heterocyclic compounds, too, have nonequivalent positions, and the principles are
similar, in terms of mechanism, and rate data is available. Furan, thiophene, and pyrrole are chiefly
substituted at the 2 position, and all are substituted faster than benzene. Pyrrole is particularly reactive,
with a reactivity approximating that of aniline or the phenoxide ion. For pyridine, it is not the free base
that must attack the electrophile, but the conjugate acid (the pyridinium ion), making the reactivity
much less than that of benzene, being similar to that of nitrobenzene. The 3 position is most reactive in
electrophilic substitution reactions of pyridine. However, groups can be introduced into the 4 position
of a pyridine ring indirectly, by performing the reaction on the corresponding pyridine N-oxide. When
fused ring systems contain substituents, successful predictions can often be made by using a
combination of the above principles. Thus, ring A of 2-methylnaphthalene is activated by the methyl
group; ring B is not (though the presence of a substituent in a fused ring system affects all the rings,
the effect is generally greatest on the ring to which it is attached). We therefore expect substitution in
ring A. The methyl group activates positions 1 and 3, which are ortho to itself, but not position 4,
which is meta to it. However, substitution at the 3 position gives rise to an arenium ion for which it is
impossible to write a low-energy canonical form in which ring B has a complete sextet. For fused
heterocyclic systems too, we can often make predictions based on the above principles, though many
exceptions are known. Thus, indole is chiefly substituted in the pyrrole ring (at position 3) and reacts
faster than benzene, while quinoline generally reacts in the benzene ring, at the 5 and 8 positions, and
slower than benzene, though faster than pyridine.
15 Prepared by Dr. P. Kar, Asst. Prof., Dept. Chem., BIT Mesra, Ranchi