2a Oe UO Du: Ceo Ope
2a Oe UO Du: Ceo Ope
spOggUUEeuog00""""
= pee at 4-4-4 4-444
a Mefolololololololelolololol>
elaes wn
ja ee U rez
O€ :
oO » <
2a oe -
eee
Y
:
co
—0 U - = >:
= Vix . :
UO Du
ceo go ::
5 oO <= :
Ope ee S
tae @ ep aie x
OXFORD CHEMISTRY PRIMERS
Physical Chemistry Editor Founding/Organic Editor Inorganic Chemistry Editor Chemical Engineering Editor
RICHARD G. COMPTON STEPHEN G. DAVIES JOHN EVANS LYNN F. GLADDEN
University of Oxford University of Oxford University of Southampton University of Cambridge
| S. E. Thomas Organic synthesis: The roles ofboron and silicon R. H. S. Winterton Heat transfer
2 D. T. Davies Aromatic heterocyclic chemistry 51 N. C. Norman Periodicity and the s- and p-block elements
3 P.R. Jenkins Organometallic reagents in synthesis R. W. Cattrall Chemical sensors
4 M. Sainsbury Aromatic chemistry M. Bowker The Basis and applications of
5 L. M. Harwood Polar rearrangements heterogeneous catalysis
6 T. J. Donohoe Oxidation and reduction in organic synthesis M. C. Grossel Alicyclic chemistry
i J. H. Jones Amino acid and peptide synthesis Second edition J. M. Brown Molecular spectroscopy
8 C. J. Moody and G, H. Whitham Reactive intermediates G, J. Price Thermodynamics ofchemical processes
9 G. M. Hornby and J. M. Peach Foundations of organic chemistry A. G. Howard Aquatic environmental chemistry
10 R. Henderson The mechanisms ofreactions at transition metal sites A. O. S. Maczek Statistical thermodynamics
1] H. M. Cartwright Applications ofartificial G. A. Attard and C. J. Barnes Surfaces
intelligence in chemistry W. Clegg Crystal structure determination
M. Bochmann Organometallics 1; Complexes with transition M. Brouard Reaction dynamics
metal—carbon s-bonds
A. K. Brisdon Inorganic spectroscopic methods
M. Bochmann Organomertallics 2; Complexes with transition
metal—carbon p-bonds G. Proctor Stereoselectivity in organic synthesis
C. E. Housecroft Cluster molecules of the p-block elements C. M. A. Brett and A. M. O. Brett Electroanalysis
M. J. Winter Chemical bonding N. J. B. Green Quantum mechanics 2: The toolkit
R. S. Ward Bifunctional compounds D. M. P. Mingos Essentials of inorganic chemistry 2
S. K. Scott Oscillations, waves, and chaos in I. Fleming Pericyclic reactions
chemical kinetics N. S. Lawrence, J. D. Wadhawan & R. G. Compton Foundations of
T. P. Softley Atomic spectra physical chemistry: Worked examples
J, Mann Chemical aspects of biosynthesis J. R. Chipperfield Non-aqueous solvents
— B. G. Cox Modern liquid phase kinetics T. J. Mason Sonochemistry
A. Harrison Fractals in chemistry J. McCleverty Chemistry ofthe first-row transition metals
NY M. T. Weller Inorganic materials chemistry
NY
NON E. C, Constable Coordination chemistry of macrocyclic compounds
R. PR. Wayne Chemical instrumentation C. E. Housecroft The heavier d-block metals: Aspects of inorganic
D. E. Fenton Biocoordination chemistry and coordination chemistry
W. G. Richards and P. R. Scott Energy levels in atoms and P. D. Beer, P. A. Gale and D. K. Smith Supramolecular chemistry
molecules N. Kaltsoyannis and P. Scott Thefelements
M. J. Winter d-Block chemistry D. S. Sivia and S. G. Rawlings Foundations of science mathematics
D. M. P. Mingos Essentials of inorganic chemistry 1 S. Duckett and B. Gilbert Foundations ofspectroscopy
G. H. Grant and W. G. Richards Computational chemistry J. H. Atherton and K. J. Carpenter Process development:
S.A. Lee and G. E. Robinson Process development: Fine chemicals Physicochemical concepts
from grams to kilograms R. de Levie Acid—base equilibria and titrations
C. L. Willis and M. R. Wills Organic synthesis H. Maskill Structure and reactivity in organic chemistry
P. J. Hore Nuclear magnetic resonance D.S. Sivia and S. G. Rawlings Foundations ofscience
G. H. Whitham Organosulfur chemistry mathematics: Worked problems
A. C. Fisher Electrode dynamics J. Iggo NMR Spectroscopy in inorganic chemistry
G. D. Meakins Functional groups: Characteristics and P. Biggs Computers in chemistry
interconversions
D. J. Walton and P. Lorimer Polymers
A. J. Kirby Stereoelectronic effects
R. J. Davey and J. Garside From molecules to crystallizers
P. A. Cox Introduction to quantum theory and atomic structure
G, M. Hornby and J. M. Peach Foundations oforganic chemistry:
P. D. Bailey and K. M. Morgan Organonitrogen chemistry
Worked examples
C. E. Wayne and R. P. Wayne Photochemistry
M. J. T. Robinson Organic stereochemistry
C. P. Lawrence, A. Rodger, and R. G. Compton Foundations of
H. R.N. Jones Radiation heat transfer
physical chemistry
J. Saunders Top drugs—top synthetic routes
R. G. Compton and G. H. W. Sanders Electrode potentials
P. B. Whalley 7wo-phase flow and heat transfer M. J. Perkins Radical chemistry: The fundamentals
L. M. Harwood and T. D. W. Claridge /ntroduction to organic P. J. Hore, J. A. Jones, and S. Wimperis NMR: The toolkit
spectroscopy G. A. D. Ritchie and D. S. Stvia Foundations ofphysics
forchemists
C. E. Housecroft Metal-metal bonded carbonyl dimers and clusters M. J. Winter and J. Andrew Foundations ofinorganic chemistry
H. Maskill Mechanisms of organic reactions 95 J. Robertson Protecting group chemistry
P. C, Wilkins and R. G. Wilkins /norganic R. Whyman Applied organometallic chemistry and catalysis
chemistry in biology J. S. Ogden Introduction to molecular symmetry
J. H. Jones Core carbonyl chemistry 98 C. M. Dobson, A. J. Pratt, and J. A. Gerrard Bioorganic chemistry
N. J. B. Green Quantum mechanics 1; Foundations B. G. Davis and A. J. Fairbanks Carbohydrate Chemistry
I. S. Metcalfe Chemical reaction engineering: A first course
Foundations of
Organic Chemistry
Michael Hornby
Stowe School, Buckingham
Josephine Peach
Somerville College and the Dyson Pernns Laboratory, University of Oxford
OXFORD
UNIVERSITY PRESS
OXFORD
UNIVERSITY PRESS
Preface
This book has been written to bridge the gap between organic chemistry at
school and at university, to stimulate the interest of advanced level chemists,
and to provide a foundation for undergraduates starting courses in chemistry
or biochemistry. The first three chapters lay the basis of the physical
chemistry needed and lead into the discussion of the reactions from a
mechanistic point of view. We have tried to keep the majority of the main text
within the A level syllabuses, with some extensions in the margins and in the
last chapter. It is not intended as, nor can it be, a comprehensive A level text.
By concentrating on mechanism we hope to emphasize the common threads
that hold the subject together.
We are most grateful to all those who have given us valuable criticism and
advice, in particular Dr Peter Carpenter (Roedean School), Dr John Nixon
(Haberdashers’ Aske’s School, Elstree), and Dr David Smith (Winchester
College). Dr Sydney Bailey (St Peter’s College, Oxford) has been a constant
source of help and inspiration to both of us since our own undergraduate
days. Colleagues at ICI Pharmaceuticals and Agrochemicals have generously
provided information on some of the commercially important compounds.
Finally we would like to thank Mrs Brenda Armstrong (Fellows’ Secretary at
Somerville College, Oxford) for her valiant efforts at deciphering not one but
two sets of chemical handwriting.
We would like to dedicate this book to Janet and to John, and to our
mutual goddaughters, Helen and Emma.
Buckingham G. M. Ef,
Oxford J.M. P.
September 1992
Contents
Molecules
Mechanisms
Acids and bases
Reactions with nucleophiles
Reactions with electrophiles
Reactions with radical intermediates
-&
WY
&
bh
uu
NIATaking it further
Further reading
Index
1 Molecules —
1.1. Introduction
OH
C,H,CH,CONH St He CH,
0 S H,
Penicillin G cove ll Muscarine
Before we can begin to answer these questions we need to understand the The abbreviated way of drawing
ways in which atoms join together to form molecules. In later chapters we organic molecules is explained on
will explore the ways in which a molecule’s architecture affects its pp. 7 and
11.
1.2 Atoms
Chemical reactions are the result of bond breaking and bond making,
involving the most energetic, outer shell electrons of atoms. Before we can
discuss bonding we need to have some idea of how electrons are distributed
in separate atoms, particuiarly carbon atoms.
The electronic configuration of a carbon atom in its ground state, or
lowest energy state, is 1s*2s’2p*. The terms s and p describe the types of
orbital in which these electrons are found, an orbital simply showing the
probable distribution of electron density.
2 Molecules
s Atomic orbitals
or
Each s orbital can hold a pair of electrons and is spherically symmetrical.
The diagram shows the region in which an s electron is most likely to be found.
p Atomic orbitals
or In the same way the electron density in a p orbital is shown as two pears end
to end. p Orbitals are not spherically symmetrical and there are three of
them, of equal energy, mutually at right angles, in each main shell.
1.3 Bonding
o Bonds
The simplest case is hydrogen, H,. The two singly occupied s orbitals on two
hydrogen atoms combine to form one bonding orbital in Hp.
Two atomic orbitals combine to form
two molecular orbitals, but only one is
occupied and bonding. OOF ca.
Bonds like these, where two half-filled orbitals on neighbouring atoms
combine to give a bonding molecular orbital with high electron density
between the two nuclei involved, are called sigma or o bonds. This is the
normal single bond.
The energy for the promotion, about +400 kJ mol™’, is more than repaid by
the bond energy (see p. 10) from making four bonds instead of two.
Double bonds. Three of the four half-filled orbitals are used to form three
equivalent ‘hybrid’ orbitals (called sp”) leaving one half-filled p orbital to
form one other bond.
—___’
ONE Is THREE equivalent ONE p
orbital ‘hybrid’ orbitals ( sp* ) orbital
In ethene, C)H,, each carbon atom makes three o bonds, using the three
(sp”) ‘hybrid’ orbitals.
If the remaining half-filled p orbitals on the two carbon atoms are parallel,
they can combine to form a new bonding orbital: a 7 (pi) bond. The shared
electron pair is most likely to be found in the two sausage-shaped lobes on
either side of the line joining the two nuclei.
am Bonds are always associated with a o bond directly between the nuclei,
giving a ‘hamburger’ look to the electron density diagram.
in
Together the o and 7 bonds make a double bond between the two carbon
atoms.
H H
This diagram shows each shared pair of electrons as a line but makes no
distinction between the o and 7 contributions to the double bond, which is
fundamental to an understanding of the reactions of ethene (see Chapter S,
p.58). We will consider the implications of 7 bonding on the shapes of
molecules later in this chapter (p. 12). Similar 7 bonds occur in the carbonyl
group, C=O, in aldehydes, ketones, esters, and so on. The geometry is
again planar with bond angles of about 120° (see left).
Triple bonds. Two of the half-filled orbitals on carbon form two equivalent
‘hybrid’ orbitals (called sp) leaving two half-filled p orbitals to form two 7
bonds.
Linear triple bonds in alkynes and nitriles are the result of the formation of
two m bonds using two mutually perpendicular p orbitals on each of the
Foundations of organic chemistry 5
triple bonded atoms, with a o bond in the middle as before. So the two
aw bonds of the triple bond are in planes at right angles to each other. The distribution of the wm electron
density is cylindrically symmetrical in
these triple bonded molecules.
H—C=C—H CH;—C=N
an alkyne a nitrile
Delocalized 7 bonds
More than two adjacent p orbitals can combine to form a set of molecular , single conventional structure does
orbitals where the electron pairs are shared by more than two atoms. These not adequately describe the methan-
form ‘delocalized’ 7 bonds. The methanoate ion HCOO , for example, can oate ion. Instead of the ‘cloudy’
be drawn as either picture, the methanoate ion can be
‘ described using the two conventional
ye structures. Neither of these contribut-
Fe or HC ing structures actually exists but the
ye ‘ true structure of the methanoate ion
lies in between them. This method of
description is called ‘resonance’ and
is denoted by double-headed arrows
but is better represented by H—CK - where between the contributing structures,
e.g.
: H H
\ \ ”
C=@ —~,- C—O
O- J
shows the probable distribution of the two delocalized pairs of electrons Resonance structures differ only in
over the three atoms. This explains the observed equal C—O bond lengths. electron distribution; all the atoms
In the case of benzene, CeH¢, there are six p orbitals making three _ stay in the same places.
delocalized bonding orbitals, each holding a pair of electrons. These are Normal C-~O and C=O bond
usually summarized as a double ring doughnut of electron density. lengths are about 0.14 and 0.12 nm
respectively. Both carbon-to-oxygen
bonds in the methanoate ion have a
length in between these values: 0.13
nm.
The o bond framework of the molecule is shown as lines. The region of high
electron density is above and below the plane of the ring. In structural
formulae (see p. 7) benzene is represented as
either oS or or
in more
The structure of benzene, and the evidence for it, will be discussed
detail in Chapter 5 (see p.67). The delocalization in benzene can be
extended to other adjacent atoms. If delocalization is extended sufficiently
6 Molecules
the compounds become able to absorb light energy in the visible region;
they become coloured.
Compounds with a system of alternate
single and double bonds are said to
have conjugated double bonds.
Dissimilar nuclei will have a differing attraction for the shared pair which
results in an unsymmetrical distribution of electron density and a permanent
slight separation of charge, a dipole. This dipole can be measured, and for a
liquid its magnitude can be related to the solvent properties. H—Cl is
actually HCl _ because the chlorine has a greater attraction for the
shared pair than does the hydrogen. This gives a small separation of charge,
see the partial charges being shown as 5+ and 5—. The bond is said to be
polarized. The presence of nonbonded pairs on only one of the atoms may
6 is used to mean ‘a small amount of’; also contribute to this dipole.
Bo Wate raeer is (acura roe’ Chlorine is said to be more electronegative than hydrogen. Some other
charge and 8— a fractional negative se ae ?
Bees electronegativities are shown in Table 1.1.
The more electronegative of the two atoms in a dipole carries the 8—. The
bigger the difference in electronegativity of the two atoms, the bigger the
Foundations of organic chemistry 7
dipole in a covalent bond between them. The molecule as a whole will have
a measurable dipole moment, unless the dipoles within it cancel each other
out by being arranged symmetrically.
H H b-
eee ae
Se) mor
cl
Ve \cl vyeet \
é- $-
clS— H
Permanent dipole No permanent dipole
Structural formulae
Molecules are drawn using structural formulae, a kind of shorthand for the
full bonding diagram. Some examples will show how they are used; look at
the yellow dyes on p. 6 and these simpler molecules below.
H
|
Methane: CH, or ae C—
O
H Il H
\ “oy Gs
axe
Propanone: CH3COCH3 or
Ei oy \ le!
H H
grouping the other atoms together in such a way as to make the structure
clear.
H
Propanonitrile Ne —C=N_ would become CH3CH,C = N
\
H
or even CoH5CN
Each corner or end represents a carbon atom and each line a bond.
Hydrogen atoms on carbon are not shown and we assume that each carbon
atom has the right number of hydrogens to give it four bonds altogether.
CH; CH; We have already shown, on p.5, the representation of the delocalized
system in benzene. So the structures on the left represent methylbenzene,
where one of the hydrogens on the ring has been replaced by a CH3; or
methyl group. The versatility of these shorthand forms can easily be
appreciated in the structures given below and the two yellow compounds on
CH; p. 6.
CH
or
NH,
O O O ¢g | N
CH; rT iI i pe
OH
Adenosine triphosphate (ATP)
H3C CH;
In the same way, all carboxylic acids possess —-COOH as the business end
of their molecules. We have already come across several more of these OXETANOCIN is a natural product
functional groups. from a bacterium which is used clinic-
ally against HIV. It includes two prim-
Cc ary alcohol groups (~CH2OH), a
\ Za
bee) ‘cae —C=N four-membered ether ring, and a
e “i primary amine.
ketone alkene nitrile
Cc
\
—NH, H as —OH
(
HOCH),
primary amine secondary alcohol
Dot diagrams
The use of such diagrams allows us to get the electron bookkeeping correct,
with the formation of the right number of bonds. The use of dots allows us
to keep track of electron pairs during the bond-making and -breaking in a
reaction (e.g. p.21). It also helps us to understand why some of the inter-
mediates of organic reactions are charged (see p.59).
H
H3C + CH; \ i
‘C—CHs -c30 You should try drawing dot diagrams
H3C CN for these two ions.
10 Molecules
Nonbonded pairs
H H If we draw a dot diagram for ammonia, we find that the outer shell of the
H:N: H:0: nitrogen atom contains a nonbonded (or lone) pair of electrons. The oxygen
H ee
atom in the water molecule similarly has two nonbonded pairs. The
ammonia water
molecular orbital picture for ammonia would have the nonbonded pair on
top.
We can simply abbreviate this to :NH3. In practice, the nonbonded pairs are
often omitted in formulae such as NH3, which can make it difficult to see
which ions or molecules will act as bases (p. 33) or nucleophiles (p. 40) until
one is familiar with them.
COCCINELLIN is a ‘defence com- Nonbonded pairs can be used to form dative covalent bonds with other
pound’ exuded by ladybirds when atoms or molecules. The ‘type reaction’ is H3N: + BF3 going to H3N—>BF3;
attacked. where — represents the dative bond between N and B using the nonbonded
The central N atom in coccinellin pair on the nitrogen. This is an underlying theme in many organic reactions
has used its nonbonded pair to make involving polar mechanisms (e.g. p.41). Although we do not draw all
a dative bond to oxygen.
nonbonded pairs onto structural formulae even when they are needed, you
will have to remember where they are. Most nucleophiles have them.
“ny
Table 1.2
Table 1.2 shows that o bonds are stronger than m bonds for C to C
connections. The 7 bond between C atoms is therefore the easier part of the
double bond to break and the chemistry of alkenes is dominated by this.
For carbon to oxygen double bonds the reverse appears to be true. This
could be one reason why there are more addition/ elimination reactions in
C=O chemistry than in C=C chemistry (see p. 49).
Foundations of organic chemistry 11
Short bonds are usually strong ones. In the case of multiple bonds, the
attraction of two shared pairs for the two positive nuclei pulls them closer
together. The effect of atomic size on bond energy can be seen in the series
C—halogen. The smaller the halogen atom, the closer the bonding nuclei
can get to each other (Table 1.3).
Table 1.3
All other things being equal, one would expect the C—I bond to be the lodobutane is indeed hydrolysed
easiest of the four to break. It is the comparative difficulty of breaking C—F faster than chlorobutane, partly
and C—Cl bonds that has made chlorofluorocarbons (CFCs) a menace to because of the weakness of its C—|
the environment. Compounds such as CCI,F, linger for years in the bond (see also p. 41).
1.5 Stereochemistry
Bond angles, molecular shapes
‘All electron pairs, being similarly charged, are mutually repulsive but
nonbonded pairs are more repulsive than others.’ All we need in order to
apply this principle to work out shapes is a knowledge of the number and
type of electron pairs around the atom of interest. Their distribution in
space will be such that they are all as far away from each other as possible.
Their mutual repulsion leads to the shape involving the maximum H
separation of electron pairs. Dot diagrams (p. 9) make good starting points.
Methane has four bonding pairs. Maximum separation is achieved by the H
four pairs taking up a tetrahedral arrangement with bond angles of 109.5°. In
the figure (right), thick lines —-_[ represent bonds that are coming out of
the page towards us, 1111: bonds that are going away from us into the page.
Lines of normal thickness — are bonds in the plane of the page.
|
wey Cca
109.5°
ig, H
The same distribution applies to ammonia (three bonding pairs, one H
nonbonded pair), though the greater repulsive effect of the nonbonded pair,
which stays closer to its parent nucleus than a bonded pair, reduces the bond
angle to 107°. The actual atoms in ammonia are arranged in a pyramid. The
angular structure of water can be rationalized in the same way with two
nonbonded pairs.
H oy H .
H:N?
Sty °
e/g ae N
H;O-
° ~ e
rae
-
H H 107° Hose!
ammonia water
12 Molecules
There are several ways in which the same set of atoms can be joined
together to form different molecules. The simplest way is to join the atoms
in a different order. Compare
OF CH)
CHEF anid, and Chin 18
methoxymethane ethanol
are also isomers, but because the molecule as a whole is flat one can simply
turn the left hand molecule over to get the right hand one. The key feature
for this isomerism is the presence of two different groups on each end of the
double bond. The conversion of one cis isomer into its trans relation is a
central] feature of human vision.
CH
all trans-retinal
11-cis-retinal
H H H CH;
: cant s th the same as \o — (
CH3 45 gies: CH, H a
14 Molecules
In fact the easy rotation about the central single carbon to carbon o bond
means that the two drawings are of the same molecule, one end being
rotated relative to the other by 120°. It is much easier to see these three
dimensional shapes with molecular models, and it is well worth having
regular access to a set of them. It could be said that Watson and Crick won
their Nobel prize for playing with models of DNA.
ALANINE
COOH '! HOOC Stereoisomerism: optical isomerism
The second type of stereoisomerism is caused by a lack of symmetry in a
Ci, wae
Ye dst Eis: SS molecule’s structure so that the molecule and its mirror image are actually
HN Neu, Hic” NH)
A B different, like a left hand and a right hand. No amount of twiddling, with
models or on paper can make the mirror structures identical, without
In diagram B you can see that we
breaking bonds; they are nonsuperimposable. An example is the amino acid
have drawn the groups the ‘wrong’
way round:
alanine (2-aminopropanoic acid).
H3C— for methyl, —CH3; If we try to rotate molecule A, hoping to get molecule B by overlapping
HOOC— for carboxylic the central C atom, the COOH and the NH) groups, we are unsuccessful
acid, -COOH. because now the methyl (CH3) group is going into the paper instead of
The groups are not different at all, coming out. Confirm this for yourself using molecular models. The isomers
but just drawn backwards to emphas-
differ in the direction in which they rotate the plane of plane polarized light;
ize the atoms which are bonded
otherwise their physical properties are the same. They are identical in
together and the mirror-image re-
lationship of A to B. The amino group chemical reactivity except in cases where shape and symmetry are critical,
in diagram A is also reversed for the most notably in reactions with other optical isomers including enzymes (see
same reasons. p. 85). In simple cases we can recognize a molecule with optical isomers by
the presence of one or more asymmetric carbon atoms, that is, a carbon
atom bearing four different substituents. A few minutes spent playing with
drawings or models will show you that the mirror image is different. On the
Only one optical isomer of DOPA,
other hand a little twiddling of H
called .-DOPA, is effective in the
treatment of Parkinson’s disease.
Com,
Non,
+
HO NH
= COO”
should show that the presence of two identical substituents on the carbon
HO
atom rules out optical isomerism. A molecule possessing asymmetry leading
to optical isomerism is said to be chiral. Biological systems are often very
Only one isomer of MORPHINE
acts as an addictive but efficient selective about optical isomers. We can only metabolize amino acids shaped
painkiller. like B above. One optical isomer of limonene occurs in oranges, the other in
peppermint oil, and both in turpentine. The two isomers smell different!
HO
There are a number of ways in which separate molecules can attract each
other. These affect physical properties such as melting point and boiling
point as well as solubility.
Dipole—dipole attractions
The permanent dipoles (see p.6) in molecules will attract each other,
following the general principle that opposite charges attract, e.g. propanone.
H3C
6+ 8- cH,
Ga Ona)
HC De CH3
:
This is a weak attraction (only a few kJ mol’; compare with the energy of
covalent bonds in Table 1.2). This weak attraction is enough to account for
the relative boiling points of propanone and butane, which have a similar c€=0
/
number of electrons altogether. Remember that in going from liquid to gas H3C
it is only those bonds between molecules that are broken. Propanone, b.p. 56°C
The attraction of molecules with permanent dipoles (polar molecules) for permanent dipole
fully charged ions helps to explain the solubility of compounds like sodium
chloride in water. The figure shows only one water molecule per ion for
simplicity. There will actually be seve.‘al clustered around each ion. Cc CH;
ice ac
Hy
Butane, b.p. 0°C
no permanent dipole
Table 1.4
These van der Waals forces, again worth only a few kJ mol’, are generally
weaker than dipole/ dipole attractions. They are important in the mainten-
ance of the coherence and fluidity of biological membranes. Phospholipid
molecules with long hydrocarbon side chains are held together by van der
Waals forces (see left).
They can also contribute to the tertiary structure of proteins. It is likely
that at the active sites of some enzymes the catalysed reaction is essentially
taking place in an organic medium, since all the water has been excluded.
Hydrogen bonding
A special case of dipole/dipole attraction involves hydrogen bonded to a
small highly electronegative atom (fluorine, oxygen, or nitrogen). This H
atom is attracted to another N, O, or F atom; this second atom must carry a
nonbonded electron pair. This is known as hydrogen bonding and is worth
10-40 kJ mol™?.
6+ H \
5+ H 0-H
e.g. water \
Wee = 26>
pe ge
The stronger the attractive inter- Hydrogen bonding is the strongest of the weak interactions and accounts for
actions between molecules in a liquid, the high boiling points of alcohols compared with alkanes or ethers
the more difficult it is for the mole- containing a similar number of electrons. Compare ethane (b.p. —89°C)
cules to escape from each other to H3C— CH; with methanol (b.p. 64°C):
form the vapour.
ot
CH, -O—CH;
\ RANE
O—H
b-— &
It also accounts for the solubility of alcohols and carboxylic acids (of
relatively low molecular mass) in water (see p. 17).
Hydrogen bonding is essential for the complementary base pairing which
holds together the double helix of DNA and forms the biochemical basis of
genetics.
Hydrogen bonding is vital in the maintenance of secondary structure, the
a-helices, and tertiary structure in proteins. It is involved in the binding of
substrate molecules to the active sites of most enzymes, e.g. chymotrypsin
(p.17). The substrate is bound in by both hydrogen bonding and van der
Waals attraction as shown.
Foundations of organic chemistry 17
\|
1.7 Solubility O
HO
This has been mentioned already in this chapter. Substances will mix
completely, or a solute will dissolve in a solvent, as long as the sum of the
attractions between molecules after mixing is equal to or more than that
before. Where there is little change in the attractions during solution there
will be no barrier to the mixing. This is the basis of the ‘like dissolves like’
principle.
Short chain alcohols and carboxylic acids, which are hydrogen bonded
before mixing, will dissolve in water (also hydrogen bonded), with plenty of
hydrogen bonding after mixing.
Propanone, with dipole/ dipole attractions before mixing, will dissolve in CH; H §+
\ /
water because of the possibility of new hydrogen bonds after mixing (see C=0"-=---< F—@)
J+ 5- Se 26-
right). CH;
Compounds like hexane and benzene, whose molecules have no
permanent dipoles, can mix freely because both possess van der Waals
attractions before and after. But hexane, with van der Waals attractions,
will not dissolve in water because, if it did so, some of the hydrogen bonding
in the water would have to be sacrificed. Hexane molecules would physically
get in the way with no compensating new attractive forces to make up for
the loss. _ Before: hake
oe CH % & Ha y
2.1. Introduction
Types of reaction
In this book we will be considering the mechanisms of three types of
reaction: substitution, addition, and elimination.
Addition reactions mean just that; one molecule adds onto another and
nothing is lost.
€.g. CH=CH, + Br2 ae CH2BrCH2Br
Water has been eliminated from the alcohol. Remember that elimination is
the reverse of addition.
pair
Electrophiles represent the other side of the coin. They are molecules or
ions which are prepared to form a new covalent bond using a pair of CHs
electrons provided by another atom or molecule, often the organic reactant. ‘:Br * HOC
Some examples are shown in the margin. CH;
Sometimes the electrophiles, such as the three positively charged
examples, actually have an empty bonding orbital available at low enough
energy to accept the incoming pair of electrons. Uncharged electrophilic St 8
molecules are either polarized, e.g. HBr, or can be so readily, e.g. Bro. H—Br Br—— Br
The addition of a proton, H*, to an alkene will show how the first kind
operates. The 7 electron pair on the alkene is shown as asterisks (*) for
emphasis.
Nuc’ tut oc
/ 4 \ H
this pair forms
the new bond
20 Mechanisms
Ng
OGG wh es
~ : Ct 3 SBrl
‘
H shee x. oe
> Br:
The 7 electron pair forms a new bond to the hydrogen atom while the HBr
bond breaks heterolytically (see p. 18). The bromine atom of HBr becomes
a bromide ion, carrying both of the original shared pair of electrons.
i @lirs = -Br:
oe
atc
yee - +
CHs
c+
Bree CH; —— Bis
N mara
CH; CH; CH3
The pair of electrons in the breaking C—Br bond go to the bromine atom,
forming a bromide ion. The positively charged carbon is short of electrons
(electron deficient). The C—Br bond has broken heterolytically.
Curly arrows ~— can be used for bond breaking, bond making, and for
both making and breaking together.
ae
bond making PNY SH SS HN
+ as
both H,N:” “CH, Ce ee H3;N—CH; + Cl
Foundations of organic chemistry 21
In each case a curly arrow must start from an electron pair, either a bonding
or nonbonded pair, and end on an atom or in a new bond.
We use (Y for radical reactions, where it represents the movement of a
single electron.
or:
(aN ligh
s Br aE = Br°* 2. ° Br
You will frequently find such mechanisms drawn without the nonbonded
pair on the hydroxyl ion. This nonbonded pair is going to form the new
bond, a dative one, to carbon while the C—Br bond breaks heterolytically
with the bonding pair of electrons going off on the bromide ion as it leaves.
Curly arrows show electron pairs going places.
An electrophilic reaction is shown below. Notice the asterisk pairs.
The advantage of this method is that it helps us to see clearly why the right-
hand carbon atom in the carbocation carries a positive charge.
22 Mechanisms
Os
mechanisms, as for example in the reaction of methane with a chlorine
radical.
Dot version:
Fishhook version:
H H
! SN
cl WRC“ H ee ke
H H
Curly arrow and fishhook mechanisms are simple descriptions. They need
to be based on real experimental evidence, such as kinetics, isolation of
intermediates and radioactive tracer experiments. Like all good theories
they must be firmly based on fact.
For all chemical reactions there are two vital questions to be answered.
i OH
In other reactions both starting materials and products are there at the end.
The classic organic example is acid-catalysed esterification.
oe Ht 4
CH; = & + C)H;0H Ola ee + H,0
OH OC)Hs
In fact the reaction is still going on at the apparent end, but with equal
rate in each direction. There is no further change in concentration of
products or reactants once the reaction has reached this dynamic equili-
brium. The concentrations of products and reactants at this point are related
by the equilibrium constant, K,, which is constant at any particular
temperature. For this example
[CH;COOC)Hs] [H20]
K,=
[CH;COOH] [C)H;OH]
= 0.26 at 373K
energy energy
change
products
products
energy energy
This reaction is thermodynamically change
unfavourable. reactants
will give a low value of K (less than 1) and there will be a low proportion of
product in the equilibrium mixture.
In a special case there is no energy change. What do you think would be
the implications for equilibrium constant and yield of product in this case?
For many reactions in solution the enthalpy changes, AH, can act as a guide
to the energy changes and, hence, to where the equilibrium lies. But they
are only a rough guide.
2. They must presumably approach each other lined up so that the bond
breaking and making process is made easy.
3. They must, between them, possess enough energy to get the reaction
started.
Foundations of organic chemistry 25
A pen and its top can collide with each other from all directions.
BS
There is only one orientation which lets the top slip on, and they only stay
together if it is pushed on with enough energy.
[ere
—_—_—_> <—
For chemical reactions, the energy barrier is known as the activation energy,
E,. The height of the barrier determines how fast a thermodynamically
favourable reaction will go. Look at the reaction profiles below; the overall
energy change is the same in each case.
reactants
products products
l
the distribution of kinetic energy at
temperatures 7, and 75.
number of
molecules T
wo
However, every cell in our bodies can achieve this oxidation, at body
temperature, through a series of reactions, each one of which uses a specific
catalyst or enzyme. Another example is:
2H 702 a 2H20 ate Op
Both manganese(IV) oxide and the enzyme catalase in liver will make the
decomposition of hydrogen peroxide solution go faster, but catalase does it
about a million times faster than manganese(IV) oxide.
k is the rate constant for the reaction. Initial rate is proportional to [A}‘[B)?’: Rate = A[A}‘[B)’
Foundations of organic chemistry 27
Here x and y are powers that are found by experiment. They do not
necessarily bear any relationship to the balanced chemical equation at all, in
marked contrast to equilibrium constants.
Let us take as an example the reaction
CH3;COCH;
+I ——
CH)ICOCH;
+ HI
Rate = k[CH3COCH
(aq) [H™ (aq)| Reaction rates are usually expressed
as rates of change of concentrations.
The iodine does not appear in the equation at all. With all the powers in we
would write
The reaction is then said to be first order with respect to propanone, first
order with respect to hydrogen ions, and zero order with respect to iodine. It
is second order (1 + 1 +0) overall, because the sum of all the powers in the
rate equation is 2.
The proportionality constant k in the rate equation is known as the rate
constant and is a characteristic of a particular reaction at a given
temperature.
The mathematical relationship between E,, the temperature 7, and the
rate constant k for a reaction is given by the Arrhenius equation:
k= Ae "RT (R=8.3JK~'mol”*)
The exponential factor, e£a’"’, is the
Increasing T makes E,/RT smaller and therefore e~ £a/®7 gets larger, and k fraction of molecules possessing at
gets larger. least the activation energy E,. This
For many reactions it is possible to devise experiments which allow us to corresponds to the area to the right of
deduce the experimental rate equation. What does it tell us? the E, vertical line on the diagram on
The actual sequence of bond breaking and making in the reaction is likely p. 26.
to take place in a number of stages, some fast and others slow. The slowest A is the frequency factor. It can be
expressed as the product PZ, where
stage will act as a bottle-neck for the reaction and dictates the overall rate,
P is a steric factor and Z is the total
so it is called the rate-determining step. number of collisions per second of the
The substances involved in or before the rate-determining step will appear reactant molecules.
in the rate equation for the reaction. You can now see how the Arrhen-
For example, the nitration of 1,3,5-trimethylbenzene has first order ius equation relates to our pen-and-
kinetics. top analogy at the beginning of the
section.
——S>>
the nitronium ion, NO,*, from nitric acid is the slow or rate-determining
step. The rate of the slow step dictates the rate of reaction. Nitration is
discussed further on p. 70.
If we understand something of the kinetics and mechanism of a reaction,
we may be able to see how to increase the rate of a slow stage to give us
more of the desired product.
What exactly is the activation energy? What is happening at the top of the
energy hump? In the substitution reaction
CH;CH,CH,CH,Br+ OH 9 —» CH;CH,CH,CH,OH + Br _
the C—Br bond begins to break at the same time as the C—OH bond begins
to form.
energy
There comes a point when the C—Br bond is weakened and the new C—
OH bond is only partly formed. At this point we are poised on an energy
reactants
maximum from which it is downhill in any direction, either back to C—Br
and OH™ or on to C—OH and Br. The energy we have invested so far is
products
the activation energy, E,. We refer to this maximum energy (downhill either
progress of reaction [=> way) state as the transition state.
Now squeeze the air into the second panel. Part-way through this process
you will suddenly feel that the balloon is poised (2). A tiny extra squeeze
sends the air on into the next panel (3); a slight relaxation lets it back into
the first (1). This in-between position (2) is like the transition state. If you
let the balloon go, the air always goes to one end or the other. But the
balloon has got to go through the ‘transition state’ if the air is to be shifted
from one end to the other.
Now we can extend this idea to a balloon with three panels, the reaction
being to get the air from the left hand panel into the one on the right. You
Foundations of organic chemistry 29
will have noticed that the reaction happens in two stages with the air-in-the-
middle (5) being an intermediate stage. Each step in squeezing the air across A catalysed reaction may involve an
will have a ‘transition state’ as before. On an energy profile it would look extra intermediate (or intermediates)
like this. in this way.
uncatalysed
transition state |
+
ve
energy intermediate
reactants
energy
intermediate
products
progress of reaction SS
Each step has its own activation energy and its own transition state. Here the
first stage is rate-determining because E,, > E,2. The intermediate itself has
a finite existence, and can sometimes be detected or even isolated in the
absence of nucleophiles such as hydroxide ions. Here is an example.
OHO ia
= OH
CLO. e:
intermediate
(carbocation)
The carbocation is the intermediate. It has been isolated as a crystalline
chlorate (VII) salt. We shall see more of these carbocations in later chap- + =
C7 cio;
ters.
Now suppose that there is more than one set of possible products. The
reaction of A+ B can go to give either C+ D or E+F as products. Which
will we get and why? Let us look at the reaction profiles.
E+F
E+F
SO3H
E+F
Chemists become quite subtle in the way they arrange for the desired
outcome, even with familiar reactions.
SUBSTITUTION OH
aqueous KOH CH3CHCH,CH;
Br
CH3;CHCH2CH3
We can prepare either the alcohol or the alkene by changing the conditions.
Look also at the two ways of brominating methylbenzene on p. 78.
3 Acids and bases
3.1 Introduction
The Brgnsted—Lowry theory states that acids are proton donors, and bases
proton acceptors. Chemists use the term ‘proton’ for the hydrogen ion, H*.
We will start by looking at acids and bases in water.
HClag) = H20q)
HzO"
(aq) + Cl (ag)
acid base
A strong acid, such as HCl, is fully dissociated into its ions in water. The
equilibrium constant for this acid dissociation is very large, and the system
settles almost entirely on the right:
HCl + H,0 H,0* + Cl
Many organic acids, such as ethanoic acid, are weak acids. The
equilibrium constants are small, much less than 1, and remarkably little of
the acid donates its proton, H*, to water in aqueous solution.
~ +
CH3COOH (aq) + H20,) CH3COO (aq) t H30(aq)
The equilibrium lies towards the left. The acid dissociation equilibrium
constant is K,.
The calculation shows that, in ethanoic acid (0.1 mol dm~°), the hydrogen ion
pH = — logso[H30* (aq)]
concentration is only 1.3 x 10-3 mol dm7?. This corresponds to a pH of 2.9.
Hydrochloric acid of the same concentration has a pH of 1.0.
Most organic bases, like ammonia itself, are weak bases. We can write
equations, similar to those for acids, to show how they act as bases in water.
ae =
CH3COOH (aq) + H20,)
CyH5NH (a9) + H20q) C2H5NH3 (aq) ap OF (aq)
no. of moles: 0.1-x \
K, is the equilibrium constant for base action in water.
= “te
CH;,COO (aq) ot H30 (ag)
[CoHsNHaqaq)]
[CH;COOH,.,)
It is often more convenient to look at bases from the other end of the
ee es ee tT ek equation, that is to consider the acid dissociation of their conjugate acids,
Owe x
the protonated forms.
Because x is very small, we can put ae +
C2HsNH3(aq) + H20q) = CoHsNHa(aq)_ + H3O0(aq)
Ko) ves VIRMO™. conjugate acid
0.1
Now use K, for the conjugate acid, C;H;NH;3°.
Thus x = 1.3x10° moldm”.
[CsHsNH3"* (aq)]
pK, = —logioK,
pKy = —logioKy
For a conjugate acid/base pair at 298K, pK, + pK, = 14. Large values of
either pK, or pK, correspond to weakness.
There is a useful consequence of this relationship. The conjugate base of
a strong acid will be a weak base.
ae *
HCl(aq) ap H20,q) H30(aq) + Cl (aq)
HCl is very strong acid; chloride ion, Cl, is a very weak base.
Foundations of organic chemistry 35
+ _
HCN is a fairly weak acid (K, = 4.9 x 107!° moldm~3). This makes cyanide
ion, CN”, a moderate base (K, = 2.0 x 107° mol dma*); Ethanol is a very
weak acid (K, = 10'*moldm~°), so ethoxide ion is strongly basic.
i = +
C2HsO0H (aq) ate H20q) ~~ C2H5O0 (aq) at H30 (aq)
3.3 Solubility
Many organic acids and bases are largely insoluble in water. Benzoic acid,
for example, is only slightly soluble in cold water. It is also a weak acid,
which means that only a small proportion of the dissolved benzoic acid will
ionize as an acid.
——— ie ot
C6HsCOOH (ag) + H20,q) er ee CeHsCOO (aq) + H30 (aq)
Addition of an alkali, like NaOH, will remove H30* from the system by SOLUBLE ASPIRIN is the sodium
neutralizing it to give water. The system will shift to the right to replace the __Salt of aspirin (see p. 1).
H30* and more undissolved benzoic acid will be able to go into solution. | Why is this more soluble than
Very shortly all the benzoic acid will dissolve in excess alkali to give sodium SP!"in itself?
benzoate in solution. OCOCH;
You might like to work through the similar argument which shows how
phenylamine, CsHsNHp, although relatively insoluble in water, dissolves COG
freely in hydrochloric acid.
We can now try to use the acidity of the acid HX to get an idea of the ability
of X~ as a leaving group in nucleophilic substitution, since both equations
have X~ on the right hand side.
cf =
HX + H,O === H30 + X
5 2 ss E
Ng ae ae —s Nu—c Fae a
Action as a base involves electron Nitriles (e.g. CH3CN), on reaction with alkali, do not lose CN™ but are
pair donation to H*. Nucleophilic re- converted to salts of the corresponding carboxylic acids (e.g. CH;COO-
actions may involve electron pair Na*). This is because HCN is a weak acid, and CN” a poor leaving group.
donation to other atoms, such as
Nucleophilic addition to C=N is faster.
carbon. Can we relate base strength
to nucleophilic reactivity? Here are
Ethers, such as C;H3;—-O—C>Hs, are resistant to nucleophilic attack
some comparisons: because C,H;O, as the base corresponding to the very weak acid ethanol,
is an extremely poor leaving group.
Base strength: You will not be surprised to find also that the best nucleophiles are often
CzHs07 > OH~ > CN7> CI the bases corresponding to the weak acids. Thus OH”, CH3;0, CN, and
Nucleophilic reactivity: CH3COO «are all good nucleophiles.
CN- > CsH,0- > OH- > CI- This is a useful guide. It cannot be more than that because we are trying
The concepts are linked but are not
to compare two situations that are very different.
the same. Nucleophilic reactivity is
measured by the rate of reaction,
whereas base strength is measured by 3.5 Acid strengths compared
the equilibrium constant, Ky.
Acid strengths, recorded as K, and pK, values in aqueous solution, are
given in Table 3.1. These are found by experiment.
Table 3.1
HX (aq) i H20q)
H30 (aq) + X(aq)
+ HX (aq) + H20q)
energy energy
change
+ =
H3O0(aq) + X(aq)
For most organic acids the energy change for dissociation in water is small
and positive. This means that the equilibrium constant, K,, is small and
certainly less than 1, as can be seen in Table 3.1. They are all weak acids.
We can now compare them and try to find reasons for their different K,
values, using the approaches listed above. As explanations they are not
fundamental but they do well as guides.
Let us start with ethanol, CH3;CH,OH, and ethanoic acid, CH;COOH.
Why should the latter be the better acid? Here the possibility of /!
delocalization in the ethanoate anion (right) makes it reasonably stable with N
respect to the parent ethanoic acid. A delocalized system involves greater O
stability, i.e. Jower energy, than expected from isolated single and double
bonds.
Such help from delocalization is not possible for the ethanol/ethoxide CH;CH, — O-
system.
In the same way phenol, CsH5OH, is a stronger acid than ethanol because
of the possibility of extended delocalization in the phenoxide ion.
Why is phenol a weaker acid than ethanoic? Here we use the idea that
delocalization is more effective for the anion which can delocalize the
charge over the most oxygen atoms. Ethanoic acid wins! In the same way
OH
\
H,0 + H7SO4 eae Oe emus + H30*
sulphuric acid is stronger than sulphurous acid because there are three
oxygens involved in the delocalization and not just two. Nitric and nitrous
acids can be compared similarly.
The pK, values for a number of organic bases are shown in Table 3.2.
Table 3.2
PHENTERMINE is a drug used to
suppress appetite. Do you think its Base pky
pKy is nearer 9 or 4?
Phenylamine Ce6H;sNH>2 9.4
NH) Ammonia NH3 4.8
Trimethylamine (CH3)3N 4.2
CH; CH3 Ethylamine CH3CH,NH, 3.4
The values for the various al/kylamines are very similar, and it seems likely
that solvent effects are important in creating the small differences that are
observed.
Phenylamine is notably weak because of the inclusion of the nonbonded
pair of electrons on the nitrogen in the delocalization. Making this pair of
electrons available for bonding to H* would involve loss of the extra
delocalization energy. Therefore phenylamine is a weak base.
Amino acids are the building blocks of proteins, compounds which have
major structural and catalytic roles in all living organisms.
+ = aE - + =
H3NCH,COO H,;NCHCOO H3;3NCHCOO
CH; CH2SCH3
glycine alanine methionine
At a pH value exactly equal to the pK,, 4.78 for ethanoic acid, the base/ This argument can be demonstrated
acid ratio will be 1. If the pH drops by 2 units, caused by added acid, 99% easily using the log form of the equil-
will be in the —-COOH form. Likewise, if the pH rises by 2 units above the ibrium expression, known as the
Henderson equation.
pK,, to 6.78 in this case, 99% will be in the -COO™ form. So at pH 7 we
CH3COO,
can write itas —-COO-. pH = pK, + logio [cH3CO0 ool
The nitrogen end can be given the same treatment and you should work |CHyCOOH
cag)
through it for methylamine, CH3NH)y. It will be easier to consider it starting
At a pH 2 units below the pK, we have
from the conjugate acid, CH;NH3*, whose pK, value is 10.6. You will
discover that at pH values below 8.5, the compound is mostly in its
pK, = 2 =
protonated form, CH3NH;". This is therefore the predominant form at pH
ch
Now look at the amino acids shown on p. 38. If we start, theoretically, [CH3COOH
4g)
with one —NH) and one —COOH group as in H2N—CH,—COOH, we can
show that in aqueous solution around pH 7.0 they will exist as the doubly
charged structures, e.g. H3N*—CH,—COO, as shown. This is also the Thus logy {CtHICOO cea} =9
form that they take in the crystalline state. |CH3COOH (aa)]
We can now follow what happens to glycine, for example, as we change
the pH of its solution:
and so [CH3C00 aa] = a
[CH3COOH
(4)| 100
H3N*CH,COOH H3N*CH,COO”- H,NCH,COO-
You should work through the calcu-
low pH pH7 high pH lations for pH 6.78 in the same way.
acidic solution neutral alkaline solution.
ie te =
| |
low pH eS eS high pH
valine
You should try drawing similar diagrams for the three amino acids in the + e
margin. H3N~ Gy -COO
At physiological pH, around 7, any free acid or amino ends in a protein l
will be charged. These charges are vital in maintaining the three CH,
dimensional structure of the protein. They can also be important in binding
the substrate to the enzyme, and in the catalytic action which follows. Quite
small changes in pH can alter the distribution of charge in the protein,
usually disastrously. Most of your enzymes only work well over a narrow pH
range, e.g. around pH 8 in the gut and mouth and around pH 2 in the
OH
tyrosine
stomach.
4 Reactions with nucleophiles
4.1 Introduction: nucleophiles
Nucleophiles can react with species which have either positive charge or low
electron density. This is the basis of many reactions, which begin with the
transfer of electron density from the more electron-rich atom (in the
nucleophile) to the more electron-deficient atom (in the electrophile). An
example is the reaction of ammonia with bromoethane.
H3N¢ ae H,C — br S007 HyN— ae Br
/
H3C CH,
Nucleophiles and bases
Nucleophiles are also bases (see pp.33 and 35) because they react with
protons, Ht. Ammonia can also act as a base:
i. g
HN? HCO = NI
Basicity and nucleophilicity are linked but are not the same.
=~ Hy s,H =
HO 72s Cra — HO-¢ + Br The first arrow shows a pair of
electrons from the oxygen making a
H3C CH,
covalent bond to carbon. The carbon
Remember that curly arrows start at a pair of electrons: a nonbonded pair already has eight electrons (a pair in
each of four bonds) in its outer shell;
for OH™~ and the C—Br bond pair in bromoethane. We are dealing with
so if a new pair is brought in, a pair
electron pairs so these are heterolytic reactions (see p. 18). must be lost at the same time. The
This is a single-stage reaction going via a transition state (TS) (see p. 28). pair the carbon loses to the Br is
shown by the second arrow.
The shape of the transition state TS is a trigonal bipyramid (see p. 12). The
transition state TS is unstable and is at an energy maximum.
TS
A reaction is more likely to go if its
transition state is of relatively low
energy.
ener
ee CH3CH Br
+ HO
CH3;CH,0OH
an Bde
General mechanism
Alkoxide and phenoxide ions also
All the reactions in this and the following five sections follow similar
react with haloalkanes.
mechanisms. If we write Nu: or Nu: for the nucleophile and X for the
Draw the mechanism for the forma-
tion of 2,4-D leaving group, the general mechanisms are:
H
a CcN
Nui osc &X —— Neate" ab
Oo COOH
or
Cl
from
Nui SCL — NO Oe sua
Gl The nucleophile’s nonbonded pair forms the new Nu—C bond at the same
time as the leaving group (X) goes off with the C—X bonding pair. We can
OH
now apply this general mechanism to many similar reactions.
Cl Hy
7 Cy Other haloalkanes with hydroxide ion or water
and Cl COOH
Many other haloalkanes react with aqueous hydroxide ion to give alcohols
Cl
in the same way as bromoethane. For example, iodomethane gives
methanol and 2-chloropropane gives propan—2-0l.
in the presence of NaOH. 2,4-D is a
selective weedkiller for broadleaved H
= H3C,< 3CH3 =
weeds in grassland. HO7 CrCl —- HO-C til
pe \
HC CH;
Overall: NaOH + (CH3)2CHCI — (CH3)2CHOH + NaCl
The reactions of these haloalkanes with pure water are much slower than
their reactions with aqueous NaOH or KOH. Water with its non-bonded
pairs of electrons on oxygen is also a nucleophile, but a weaker one than the
negatively charged hydroxyl ion (see p. 35). «
Foundations of organic chemistry 43
Cyanide ion, CN~
Cyanide ion reacts with haloalkanes in the same way as OH:
H H
Bisa SH 3 Here again the cyanide ion is drawn
NG Za Ce Bi eae NG + Br the other way round as NC~ to put
TX \
H3C CH; the nonbonded pair on the carbon
atom of the cyanide ion near the 8+ C
Overall: KCN + CH3CH,Br a CH3CH2CN + KBr of the haloalkane.
A new C—C bond is made in the product, which is a nitrile. These nitriles
are particularly interesting because they can be elaborated into other
organic compounds. An example using iodomethane is shown below.
LiAlH, CH;CH,NH,
=
KCN
CH31 ———————» (CH,CN
nucleophilic
substitution In an organic hydrolysis reaction,
conc. acid O
hydrolysi
ydrolysis CH; ea
water is involved both as reagent and
as solvent; the organic molecule is
OH split.
Overall: CH3CH,2NH>) +
aa = Draw the mechanism for the prepara-
CH3CH,Br = CH3CH2NH7CH»CH; Br
tion of GLYPHOSATE
+ a
(compare NH3 + HCl ae INISY EN
H
If you want the product amine and not the salt, you can either treat the
HOGG No ee Pe
product with a strong base to liberate the weaker base (the amine) from its
salt
+ ~
+ +
NH; + H ——
———_ NH,
H H
Thus an amine cannot be made B
E
Se Pee
as gia a eo
_—
\
i + H,O
2
GIScl Gl
- He a BH faCl -
Cl ~~~ C=0—P—Cl a Cle + O=P Tel
page| \ \
H3C Cl CH; Cl
/
HoC—CH, —» H»C—CH,
7
But the product is also a nucleophile and can attack another molecule of the
cyclic ether, and so on until a polymer is built up:
HO—CH2—CH,—O —CH,—CH2—O—CHz . . . The repeating unit is This cyclic ether can also be polymer-
CH,OCH), an ether, so the polymer is called a polyether. Polyethers are __ ized using an acid catalyst. Draw the
used for making surfactants. mechanism for this.
In all the reactions so far, the nucleophilic substitution has taken place in a
single stage and the new bond to carbon is made at the same time as the
leaving group’s bond is broken.
ees ee A Aa ‘A ;
Nu-*C X — Nu---C---X — = Nu-—-C + X
transition state
planar, trigonal
carbocation
Compare this two-stage reaction profile with the single-stage one (p. 42).
energy
You can guess that the halocompounds which react by this two-stage
mechanism will be the ones which can ionize to give low-energy, stable
carbocations. One example is (CH3)3CBr, a crowded halide which gives a
cation that is stabilized by inductive effects. A second example is
(C6Hs)3CBr, which is also crowded and whose cation can be
stabilized by the benzene rings.
x 4
(CgHs)3CBr — > (CgHs)3C + Br
H3C CH;
Overall: KOH + (CH3).CHBr — > (CH3)2CHOH + KBr
i)H
Overall: (CH3)2CO —_— (CH3)2CHOH
then ii) H"(ag)
with great care
H,: ef
$40 ay $,0H CN
NC— - H TN —> Ne— S ar
CH; CH;
H
i
Overall: CH;CHO + HCN — Cha oT OH
CN
Notice that the cyanide ion used in the addition is regenerated at the end; it
is a catalyst and strictly it is the HCN which is used up.
The best reagents for this HCN addition are a mixture of HCN and KCN,
formed by addition of cold sulphuric acid to KCN. HCN is a weak acid (see
p. 33); by itself it produces a very low concentration of cyanide ions and so
the reaction rate would be slow. Excess KCN is needed to increase the
reaction rate by raising the cyanide ion concentration. Alternatively, a small
amount of a base, such as KOH, can be added to react with HCN to
increase the concentration of CN.
Note that these are overall additi»ns of HCN. Contrast the conditions
used with those for cyanide ion substitutions on p. 43.
Uses. The HCN addition product, with its new C—C bond, offers scope for
further synthesis (see p. 43).
H3C
\_-OH-
H3C
hydrolysis
Sees
aa . EL ~COOH
H3C
\ KCN
+ H,SO Nee Ort
C=O pee (Oz
/ 7 CN
H H
reduction \ OH
H
Foundations of organic chemistry 49
: (a)
Si elimination Rome mS OH
N=c x
+ H,0 nec = \ Predict the structure of the product of
CH,
4
Overall: CH;
NH) CH nee
HN + o=c ——~ HN CH3 + H,0
ae
‘ NO) CH; NO,
P 4 6 2
5 3 2 :
4 4
NO, NO,
50. Reactions with nucleophiles
Cl Cl
Uses. These reactions have all been used to make crystalline derivatives of
aldehydes and ketones; now they are used more often in syntheses. We have
exchanged a C=O for a C=NX group and made new C to N bonds. Many
This is part of the synthesis of
nitrogen-containing compounds (such as dichlobenil) can be made using this
DICHLOBENIL. CN
type of reaction.
@ Cl
H3C H3C ne 3C eS
C=O C=0 C=O C=O C=O
/ / /f
H3C H OH OCH, Cl
(ketone) (aldehyde) (carboxylic (ester) (acyl
acid) chloride)
We can write a general structure CH3;COX for all of these compounds, where
only X changes. The main distinction between the left- and right-hand sets is
that for the right-hand set X is a possible leaving group (see p.35). These
can therefore undergo addition—elimination reactions re-forming the C=O
group.
The initial stage of the reaction is the same for all: the nucleophile adds to
the electron-deficient carbon of the °*C=O°" to give a tetrahedral
Foundations of organic chemistry 51
intermediate. For the acid derivatives this can now go on directly to re-form An example is the conversion of an
the C=O group by eliminating X (often as X~ ion): e.g. ethyl ester into a methyl ester (a
transesterification reaction).
HC HC oe HC
=
Nie C=O
\ one
addition Nieuw
20 ; ‘4
elimination
3 CH;
‘c=o eg CO
aa oO
YA
xX
AMA G X
\ Nu
/ OGjH:
intermediate
Note that the observed flat structure of the amide group, X = NHnp, with its
angles of about 120° around carbon and nitrogen, would be predicted from
this delocalization. This flat amide or peptide structure is very important in
maintaining the shapes of proteins and enzymes.
Now look at the balance of inductive effects and delocalization on the Esters and acids also have
reactivity of CH;COX.
4
a flat = system
1. Acyl halides (also called acid halides), e.g. X = Cl (CH3COCI). Here the o—
Let’s start off by looking at the most familiar reaction of esters, ester
hydrolysis, and its reverse, esterification.
52 Reactions with nucleophiles
Esters
HO Ge= — + HO-C \ A
OCH,CH; OCH,CH;
(ii) Intermediate A re-forms the C=O bond with elimination of alkoxide
ion, CH3CH,0 .
Hs oO
oy) if
CH3 =
This is a very good way to hydrolyse esters, and it is used in the hydrolysis of
natural esters (fats) for soap production.
Esters can also be hydrolysed in aqueous acidic solution. The C=O group
is first protonated to make the carbonyl carbon even more electron-
H3C CH; deficient; this is necessary because water is a weaker nucleophile than OH.
It is an ester. Draw the structures of See if you can complete the mechanism for this.
the acid and the alcohol from which it
might be synthesized.
HC N H3C
% 4 +
ye = O. H —_— pe =OH
OCH,CH, OCH,CH3
Overall: H’ + CH;COOCH,CH; + H;0 as
CH,COOH + CH,CH,OH + H*
The proton is regenerated at the end so this is an acid-catalysed hydrolysis.
Since it is a reversible reaction, an equilibrium, we can also use it to prepare
esters. We still need the acid catalyst, in order to reach equilibrium faster,
but we must now start off with as /ittle water as possible so that as much ester
as possible is present at equilibrium.
Thus, if we choose the conditions carefully we can use the acid-catalysed
reaction for either ester hydrolysis or esterification.
Foundations of organic chemistry 53
ESTER HYDROLYSIS
+
Ge
=H3C 45 => 4g, elimination ees
uf ~
eel —o= eens H—-c* Hee + OCH,CH3
\
OCH,CH, OCH,CH; e
H3C H3C _ ee
i) 4H a Nc =o ve
addition H—-C $0
4 protonation H—-C = 4 + OH =
/ \ +H,0 \
H H H
Notice that NaBHy, which is less reactive than LiAlH,, does not normally
reduce either esters or acids but will reduce aldehydes and ketones.
Amides
Amides and peptides contain the same planar functional group (see p. 51). i
Many important natural products (proteins, enzymes, hormones, antibiotics) eee
and synthetic polymers (nylon) contain this group. Amides and peptides are
generally less reactive to nucleophiles than esters, but like esters they can be |
hydrolysed in either acidic or alkaline solution by nucleophilic addition— amide or peptide link
elimination mechanisms. q f
Overall: CH3CONH) + H20 + H —s CH;COOH + NH,
Hooc Sy
H
N COOH
O 4O NH 3CH3
H3C ae A
d O ats 2CH3NH = H3C — . +
a
O a ‘ NHCH3 =oz oe
4
CH3 CH;
You should be able to see why two molar equivalents of amine are used in
the second example.
Acid anhydrides can be made from acyl chlorides and anhydrous
carboxylate salts. Write out this mechanism, too.
Foundations of organic chemistry 55
Carboxylic acids
Nucleophiles are also bases and often react with carboxylic acids to form
salts. For example,
2 +
CH3;COOH + NaOH ——= CH;COO Na + H,O
This limits the reactivity of carboxylic acids with nucleophiles, because the \
O
resulting carboxylate anion is much less reactive to nucleophiles than esters
Predict the structure of the product
or acid anhydrides.
formed when this anhydride reacts
= at
CH3COOH at H,NCH»CH3 — > CH;COO H3NCH2CH3 with ammonia, and write the mechan-
ism.
Under forcing conditions amides can be made from carboxylic acids, e.g. in Captan is used against apple and
the polymerization to give nylon-6. pear scab. It reacts with other com-
pounds in the fungus to produce toxic
H,N(CH2)sCOOH gives ---- NH(CH2)sCONH(CH))sCONH(CH));CO ---- CSClo.
Carboxylic acids react with inorganic acid chlorides such as SOCI, or PCI,
to give organic acid chlorides; compare this with the conversion of alcohols
into haloalkanes (see p. 44).
dr
CH,COOH + SOC, —» CH;COCI + SO, + HCI
Aldehydes and ketones differ from the acid derivatives because they have
H, alkyl, or phenyl instead of a possible leaving group, X. They all undergo
addition first to give a tetrahedral intermediate. Now the best leaving group
in the case of the aldehydes and ketones is the nucleophile (—Nu) and its
loss simply reverses the addition. For example, for propanone with cold,
dilute OH “:
56 Reactions with nucleophiles
H3C
1G 30
3 se
HO Cl
Foundations of organic chemistry 57
Presumably these intermediates are of high energy and are difficult to form,
because the Group III halides are essentially inert to OH~ under normal
conditions. Notice how the reactivity changes sharply according to whether
the Cl is directly attached to a benzene ring (Group III) or attached to a
carbon next to a benzene ring (Group I). _
The reactivities of the three groups, with variations according to precise
structure, are summarized in Table 4.1.
Table 4.1
Group II I Ill
Type Acyl halides Alkyl halides Aryl halides
Reaction withaq.OH™ Violent, almost Fast No reaction
explosive
Reaction with H,O0 Very fast Slow No reaction
5 Reactions with electrophiles
5.1 Introduction
Notice that the stable cations Na* Cations, e.g. H* from HCl, NO,* from HNO3. These cations can readily
and K* are not included because accept a pair of electrons to form new covalent bonds.
they do not easily accept extra elec-
trons to form covalent bonds. Neutral molecules, e.g. Brz, HBr. Some of these are easily polarized, such
as bromine, ®*Br—Br®~; the 8+ part is electrophilic. Others are
permanently polarized, e.g. hydrogen bromide (see p. 6), °*H—Br®.
We could also include as electrophiles all the organic molecules which react
with the nucleophilic reagents in Chapter 4, e.g. the polarized haloalkanes
and carbonyl compounds. To prevent repetition, we will follow the usual
convention in organic chemistry and classify the inorganic and small organic
species as the reagents. This can lead us into trouble when neither reactant
fits into the ‘reagent’ category!
The electron-rich organic molecules in this chapter all have formal carbon—
carbon double bonds. The carbon-carbon double bond can be considered as
a o bond and a w bond. The electrons of the 7 bond are in a m7 molecular
orbital, made by combining two p atomic orbitals. This 7 orbital has its
average electron density further from the positive carbon nuclei than in a
single o bond. The electrons are more easily available to electrophiles.
The carbon-carbon double bond is weaker than two single bonds and so
electrophiles (generalized as E*) normally add to C=X. In the addition to
C=C, electrons are transferred from the double bond to E*, creating
positively charged intermediates.
7 ,
fe ) NS E ‘ss ~
E*
intermediate
A new C-—E bond is formed leaving the other carbon atom as an electron-
deficient carbocation, with only six electrons in its outer shell. This may be
clearer from a dot diagram taking ethene as our example.
Foundations of organic chemistry 59
H, nisl H OH
a G ‘ Cc" os Hi 2s@y 3. Ga
H re H A EH
E eu %
8 electron C 6 electron C
The electrons of the 7 bond both go to form the new C—E bond, so this is a
heterolytic fission of the bond. The reaction can then be completed by
addition of an anion, X~.
+
+ Es
ECH) = CH) te x —_ ECH, ae CH)X
Ethene
When hydrobromic acid and ethene are mixed, oily drops of bromoethane
are formed.
The electrophile is the proton, from the dissociation of HBr, a strong acid.
ot “
HBrvaq) Hoag) + Br(aq) If the HBr |is not dissociated, the
polarized H —Br molecule can act
The proton adds to the carbon-carbon double bond to form a new C—H as the electrophile to give the same
bond. The carbocation intermediate then adds the Br’ to give bromoethane, carbocation intermediate.
in which both C atoms now have eight outer shell electrons. For dot
diagrams this is:
vt
N
9) | Ose
ee,
Fin om
EG 0
H
ay
if H ee =
Ht carbocation intermediate wo—m
a
Or, using curly arrows (remember that the arrow begins at the electron pair
and goes to make a bond to the electron-deficient atom):
H H ie) | OR +
\ 7
pe~ a 4 Pa ‘N re \ pice
w emg
=
Guta
proeeeats
OD
H a H H H H HH
H ai
intermediate
60 Reactions with electrophiles
Propene
The initial addition of H* to propene can give two different carbocation
intermediates, one primary (P) and one secondary (S).
H H Z H
Either Sn eee eee ct
Primary carbocations have only one HC ) H Mm
C substituent attached to the C*, ea (P)
secondary have two, and tertiary (e.g.
(CH3)3C*) have three.
H H H Ay
ot ic N SS re=e4
7 vA \
H,C A H H3C H
H* (S)
The secondary carbocation (S) is more stable (of lower energy) than the
primary one (P) and. is formed faster in this first, rate-determining step of
the reaction. We assume that the transition state leading to the secondary
More stable (lower energy) intermed- carbocation is of lower energy than the corresponding one leading to the
iates are likely to be preceded by primary carbocation. Let’s consider two possible reasons for this: stereo-
lower energy transition states, since chemical and inductive.
the structure of a transition state is
probably similar to the structure of the Stereochemical effects. For a stable three-coordinate carbocation, the ideal
intermediate. angles around C* are about 120° (three electron pairs, as far away from
each other as possible). This keeps large substituents farther apart than the
109° (tetrahedral) bond angle of a 4-coordinate carbon. Thus the change
from four- to three-coordination favours the more stereochemically
crowded carbon; tertiary > secondary > primary for stability of carbo-
cations.
H3C G Hy
+
stability fe
|
CH;
|
CH,
|
H
In general, the more stable the intermediate, the lower its energy and the
more likely it is to be formed. Therefore the addition of HBr to propene
goes mostly via the more stable secondary carbocation (S) to give 2-
bromopropane, rather than via the less stable primary carbocation (P) to
give 1-bromopropane. In fact, the product is nearly all 2-bromopropane.
Br Draw the mechanism for the addition
CH=CH
= ae = nb zee of molecular HBr (not ionized) to
H3C* i : H3C~ ~CH, fear CH, propene.
H+ (S)
For propene and HCl, the 2-halopropane is again the major product.
If you want to predict which haloalkane will be formed by polar addition Try this idea for the addition of HBr to
1—pentene and to F3CCH=CH>.
of hydrogen halide to an unsymmetrical alkene (like propene), draw the
possible carbocation intermediates. Work out which you think is the more
stable; this intermediate should be formed more easily and lead to the major Markovnikov’s rule, which states that
product. ‘the more positive Y portion of the
reagent Y—Z goes to the carbon
atom of the C=C which already bears
the more H atoms’, is an empirical
5.3 Reactions of alkenes with sulphuric acid: hydration rule governing these additions, and
was used before the mechanisms
Ethene were understood.
Concentrated sulphuric acid (a strong acid) will protonate the double bond
of ethene just as HCl and HBr do.
H,SO AD =
CH,>=CH, 2s CH;—CH, HSO,
carbocation
intermediate
H H
H\ 7
H “
Hat C—C+ / J HSO,
H = ‘
Hy% Cc—- Vo
OSO3H
C=C — a Gp
7 ) \ My re de MH
H H H H H ?
y a hydrogensulphate
Coso3H ester
: OSO3H Tag OH
C=C, — H
is C,.,
\ "
a ’ H
H H hydrolysis H
‘onc. H2SO,
Overall: CH,»=CH, + H2O Peers ae) (CHR CHOU
Propene
The acid-catalysed hydration of propene gives propan-2-ol. First,
protonation gives the more stable, secondary carbocation (see p. 60) which
then forms the secondary alcohol. The mechanism is similar to the addition
of HBr.
H3C H3C OSO3H OH
GH=ca,
= 25 ‘tH — Cea aeee
— ag BS
e : H,C— 4 CH, H3C ~ yy CH3
H. a H7SO4
(0s
The hydration of ethene and of propene are of great industrial
importance, as the two alcohols are used on an enormous scale as solvents
and as chemical feedstock.
Ethene
Br Br
at x Cl
Cl /
HC as CH) (3)
Be
The formation of compounds (2) and (3) is seen as evidence for the cationic
intermediate. What other product might be formed if sodium ethanoate
(CH3;CO,~ Na*) were present in the bromine solution?
Propene
In the addition of bromine to propene, the intermediate cation is not
symmetrical and more positive charge is on the secondary carbon atom than
on the primary carbon atom (carbocation stability: tertiary > secondary >
primary, see p. 60).
H3C, 5. ee A
CH,CH = CH, —— H=—C-C ane Br
44, ~ -<
Srl
pf
Br-=—Br
La
The nucleophiles attack the more positive carbon atom (6+).
pe ae Nu
H3C,,= \ =
fo hae —e aes
Ca, CH, Br
o+
SG c CH tc.
HC’ ~Cc* ‘7? = 4c
— Chr
CHs CH CH;
i
sD co2fN 2CH
HCH C.* H,C
H
The structures of these alkene polymers can be predicted using the stability
of the cationic intermediates, in the same way as we predicted the
orientation of addition of HBr to propene (see p. 60). Draw the mechanism
and the product for the cationic polymerization of HyC=C(CH3)p.
The polymers no longer have alkene C=C bonds, so they are resistant to
attack by electrophiles like acids and KMnO,. They are generally inert, like
alkanes, as you would expect from their structural formulae. The reason
why so many of these plastics are not biodegradable is that their molecules
are not accepted as substrates by fungal or bacterial enzymes.
H) H)
H,C NN 4 CH)
n Selec ee ‘c=c’%
if \ / N
H3C H H3C H n
H H O= © H H
2. Deduce the structures of the pro- ozonide
ducts from treating LIMONENE with
first excess O3 and then Zn + H2O. Ozonolysis has been used to identify the substituents on each end of a C=C
bond; the two carbonyl compounds produced could be identified (for
example) from the melting points of their 2,4-dinitrophenylhydrazones (see
H3C p.49). In modern organic chemistry, ozonolysis is more likely to be used to
prepare aldehydes or ketones from alkenes.
Potassium manganate(VII)
The C—H bonds are left out of the last three diagrams for clarity, as is usually
done for complex organic molecules. Don’t forget that if no substituent is
shown on a benzene ring, you must assume that there is a hydrogen atom
attached to each carbon, making C—C—H bond angles of 120°.
The actual shape of the benzene molecule is a flat, regular hexagon (from
the X-ray crystal structure analysis of the solid). The carbon-carbon bond
lengths are all the same (0.139 nm), a value in between normal double
(0.134 nm) and single (0.154 nm) bonds.
If each carbon atom in benzene makes single bonds to each carbon
neighbour and a third single bond to hydrogen, there will be one unused
outer shell electron left in a p orbital on each carbon (diagram A).
These six atomic orbitals can combine to form a new set of molecular
orbitals in which all six ring carbon atoms are involved. It is this lower-
energy arrangement of the remaining six electrons in their special set of
molecular orbitals which gives benzene (and other ‘aromatic’ compounds)
their characteristic properties, such as bond lengths, stability, and resistance
to reaction compared with alkenes. The six electrons are delocalized over
the six-carbon ring, with the electron density greatest above and below the
ring in a ‘double doughnut’ shape (diagram B). This structure is much more
stable than the theoretical ‘cyclohexatriene’ structure which would have
alternating double and single carbon-carbon bonds. Let’s look at the
evidence for this extra stability, compared with alkenes.
1. Thermochemistry. The enthalpy change on hydrogenation or on Remember that six atomic orbitals
combustion of benzene is less exothermic than ‘expected’ in terms of a combine to make six molecular
structure with three alkene C=C bonds. orbitals, of which three are bonding
orbitals and are occupied (see margin
on p. 3).
(i) Hydrogenation
CéHeoy + 3Hag) = CeHi2)
AH = —209.2 kJ mol?
‘expected’ value is —360.4 kJ mol”!
Cyclohexane
(ii) Combustion
Draw your own enthalpy diagram for CeHeay + 15/2 Or(g) = 6CO 2g) ate 3H20
the combustion. AH = —3313.8 kJ mol!
‘expected’ value is —3473.4 kJ mol!
These figures suggest that benzene is much more stable (of /ower energy)
than the theoretical ‘cyclohexatriene’ model, by about 150-160 kJ mol™?.
The delocalization energy contributes to this stabilization. Breaking the
bonds in benzene is more endothermic than it would be for cyclohexatriene;
the difference is used as a measure of the delocalization energy.
2. Low reactivity of benzene. If benzene is more stable than the alkenes,
we expect it to be less reactive (more energy needed for reaction). This is
true for a wide range of reactions. Here are some examples.
The two standard ways are the ‘nondelocalized’ and ‘hexagon + circle’
diagrams: on :
Foundations of organic chemistry 69
either or
First of all, neither way is ‘right’ but both are useful. A ‘nondelocalized’ Only one 1,2 disubstituted CeH4Xo
diagram implies separate double and single carbon-carbon bonds; this is not &xiSts; that is
so, as all the carbon-carbon bond lengths are identical in benzene.
Nevertheless, these diagrams are useful for drawing mechanisms, and we x
will use them for mechanisms. In the hexagon + circle diagram, the circle
represents the six delocalized electrons in their set of molecular orbitals.
This shows the hexagonal symmetry of benzene but it is more difficult to use
when drawing mechanisms. Professional chemists often use the hexagon +
circle for speed but change to nondelocalized diagrams when drawing
mechanisms.
and
5.8 Electrophilic substitution of benzene
18)
sO x
—_ ae
substitution addition
product product
The stability of the benzene system pulls these reactions towards overall
substitution rather than addition. The three electrophilic substitution
reactions we will look at in more detail are nitration (in which NO replaces
H), bromination (Br replaces H), and chlorination (Cl replaces H).
70 Reactions with electrophiles
Nitration
The detailed equilibria are: The H,SO, does not appear in the equation. It acts as an acid catalyst and
H,SO, HSO4 the reaction is very slow without it. The electrophile is the electron-deficient
a2 SSS se ‘ ion NO,* (O=N =O). This is formed by mixing a very strong acid (e.g.
HNO} H»NO} H,SO,) with nitric acid.
HONO; === ehs088 NO, Overallin, 2H5SO,. .4.EINO - NO, 4. 4:0 + 2HSG,
We will now draw the mechanism of the substitution using a non-
H,O H,0* delocalized diagram for benzene. This makes it easy to compare this
+3 a he mechanism with the electrophilic addition to alkenes.
H)SO, HSO, +
NO,
ey >
addition
wunH
+
a
ary +
elimination
A
Compare:
Sseen Me E mae E
pene a ] py l
Ee addition Pe addition x7 ne
B
Intermediate cations A and B are formed by the initial addition of E*, but
notice that the positive charge in cation A can be delocalized using the two
remaining conjugated double bonds. This can be drawn as a ‘horseshoe’ to
Cation A:
show the delocalized electrons. Note that the ‘horseshoe’ delocalization
involves carbons 2 to 6 inclusive but not carbon I (which is saturated and
tetrahedral). The reaction is completed by loss of a proton to re-form the
fully delocalized benzene system.
+
mH 3 %
or ae
The structure of the nitro-group can Es NO> NO, NO;
be drawn as
cP" | x oy
The aromatic amines are very useful synthetic intermediates because a wide
variety of new functional groups can be introduced via diazotization (see
2.23): PARATHION is an aromatic nitro-
The explosive TNT (trinitrotoluene, or 2,4,6-trinitromethylbenzene) is COMPOUNd which isa broad spectrum
made by (careful!) nitration of methylbenzene. lnsecticie it meanics acctyicnolie
and acts as an inactivator of the
CHs enzyme acetylcholine esterase.
NO)
Polynitrocompounds are thermodynamically unstable chiefly because they
produce the very stable nitrogen molecule, N2, when they decompose.
NO,
Bromination
Br* has only six electrons in its outer shell and will now react with the
benzene.
+ Br -
Br Br
ee LE un H Ht
—_ eS
mS SF:
72 Reactions with electrophiles
iy 4 Albee Se er Ales
All these reactions are usually carried out away from bright light to avoid
radical reactions (see p. 78).
Chlorination
Dy 5
=
£) white precipitate
+ 3Cly a OH
(CH3)2CH CH(CH3)2
NH) ee
HNO, + HCI A
ee Cl” + 2H,0
0 to 5°
a diazonium salt
The —N.* group is weakly electrophilic and can react with very electron-
rich aromatic compounds such as phenols. This is the basis of the azo dye
industry, because the ‘aromatic ring-N=N-aromatic ring’ products are
brightly coloured, fast azo dyes. They are widely used in the printing
industry.
O Nat OH
N
fe N mame NII + NaCl Draw the two geometric isomers of
a
the dye.
Cl
Work out the mechanism of this last reaction (it’s an electrophilic substitution
reaction on the phenoxide ion).
The —N,* of the diazonium salt also provides the best leaving group of
all, stable molecular nitrogen, Nz. Diazonium salts are also useful
compounds for further synthesis because many different groups can be
substituted for the N2*, e.g. cyanide.
UeA
N
N ; ven CN
Cl eekge +ON5. KCl
CuCN catalyst
74 Reactions with electrophiles
SOS
is
E+ E
delocalized
cationic intermediate
E
re S| “s Ht
re
CH, COOH
+ 30 — a5 H20
light —
Cle El ee CaCl
CH, tcl ae CH CI
H H
or Tiere — , H:C° + Figen Gla
H H
The new methyl radical "CH; can then take a chlorine atom from Cl, to
make CH3Cl and another chlorine radical.
Ch t5CH, oe Cl._+, CHC
or Ge — Cl° + CI—CH,
This chlorine radical can go on to react with another molecule of methane,
and so on. Notice that in each of these propagation steps there is an odd
number of electrons on both sides of the equation.
Cl Cl
is the combination of two radicals. cl a ‘CH; —
Paraquat is used for cleaning cereal
stubble before direct drilling, as part
of ‘minimum cultivation’ to retain soil fe ‘Ch. — + H,C—CH,
structure. Paraquat is inactivated by
the soil.
Putting all the steps together:
Ny light s
Initation: Cl, — —2C!
Propagation: CHC) =e TCH re!
light
OVERALL: Cly + CHy ee CH;C! + HCl
Foundations of organic chemistry 77
We can also show this as a radical flow diagram. The central circle contains the chain-
carrying radicals in the propagation
CH;Cl ie CH, steps. The side arrows show which
starting materials go in and which
products go out at each propagation
step.
Ge "CH3 HCI
This is the major overall reaction if there are equimolar amounts of Cl and
CHg,. If there is more Cly, the reaction can go on to give CH,Cl,, CHCl,
and eventually CCl, as well as HCl. The propagation steps for CH>Cl are:
~e
A
‘a °
re — HC. + H-C
Cl
ae Pi oes
H»C—Cl —> H,C—Cl
° \
Write the corresponding steps for CHCl; and CCl,. Remember that the
termination steps are any combination of chain-carrying radicals. By
adjusting the relative amounts of CH, and Clo, you can make mostly CH3Cl,
CH2Cl,, CHCl;, or CCly. Draw radical flow diagrams for these, too.
Bromination of methane
Conclusions
CH, CH;
Br
NOLVADEX is used to treat breast
cancer.
Using structural formulae, draw the . :
equation for the reaction of Nolvadex 6.3 Reduction: catalytic hydrogenation
with hydrogen and a palladium cata-
lyst at room temperature. Alkenes can be reduced to alkanes by the addition of hydrogen in the
presence of a finely-divided transition metal catalyst, often Ni, Pd, or Pt.
CH;CH,
Xs Pd
€ CH;CH=CH, + H2 —- CH3CH2CH3
I] room temp.
Pd
C.H;sCH = CH a H) ee CeH5CH2CH3
room temp.
The mechanism of these reactions is not fully understood. Both alkene and
hydrogen are probably absorbed on the catalyst surface, and the H)
molecule is probably split into hydrogen atoms. The reaction is thus
CH3—N. heterogeneous (taking place in more than one phase, here gas + solid) and
CH; homolytic (H—H goes to 2H , bonding electrons shared evenly).
Foundations of organic chemistry 79
The world’s major sources of hydrocarbons are coal and oil. Both contain a
wide range of hydrocarbons. Some of the most desirable hydrocarbons have
quite low relative molecular masses. The less useful, higher relative
molecular mass hydrocarbons can be ‘cracked’—that is, converted by Plant oils are possible renewable
heating into hydrocarbons of lower relative molecular mass. The heat fuels for the future. The methyl ester
energy causes homolytic breakage of the C—C bonds to give radicals, which of rape-seed oil has been tested as
react further to give alkanes and alkenes. an alternative tractor fuel, and palm
oil has been suggested as a future
car fuel.
heat °
e.g. CH3(CH, Cy - CHy(CH,)4CH3 —» 2 CH;(CH>),CH2
CH jh
CH3(CH2)4CH2 H Oe Saas CH3(CH?)4CH3 13 CH
\(CH,)3;CH, (CH,),CH:
heat
Overall: CH3(CH,)9CH; 9——= ~~CH3(CH),CH; + —CH3(CH,);CH=CH,
Ethene
Radicals will add to carbon-carbon double bonds to make new radicals. If
R’ represents a radical from the initiation step:
80 Reactions with radical intermediates
saan nearemcarei y
brie
ois?
Liem
R
The new radical can add to another alkene molecule, and so on to make a
polymer. For ethene these steps are:
H) J H> H> 2
ROH, ACA Cth WF aba RCH ACN CK CH gto,
H> H) Hy Hy Hy
R
Propene
Propene is not a symmetrical alkene and R’ could add to either end of the
double bond to give two different radicals, one primary (P) and one
secondary (S):
7 ne R° x
(P) (S)
Usually reaction goes via the more stable radical, which is preceded by the
lower-energy transition state. The stabilities of alkyl radicals are in the
order
tertiary > se ondary > primary
This is the same order as for carbocations. Both species are 3-coordinate
and electron-deficient, so some of the same influences may be operating
here, e.g. stereochemical and inductive effects (see p. 60). Polymerization
of propene goes through secondary radicals like (S) and so gives a regular
head-to-tail polymer. Here is the overall equation: you should write the
detailed mechanism.
R CH,
\
Poly(propene) is used to make ropes n CH,CH=CH, — —+ CH—CH,>—
and in light engineering. polypropene, or polypropylene
Foundations of organic chemistry 81
Phenylethene
When a radical adds to phenylethene, the more stable radical intermediate
is the one where the carbon with the odd electron also carries the benzene
ring. This radical can be stabilized by delocalization over the ring.
H,
CH i. 26
HO4N AL °*R Hes ES R
—_—_—_
HORA
CH, ee e Ec ae
Adee HC<
e
psc. Ae
S Gc
— etc
H)
c
Overall: nCg.H;CH=CH, — Puc }
| Expanded polystyrene, used for
C6Hs thermal insulation and packaging, is
a honeycomb made using gaseous
The polymer produced is ‘polystyrene’. pentane during the polymerization.
Here solvation is not possible and radical reactions often dominate, for
example in the atmosphere. The formation of photochemical ‘smog’ and the
formation and destruction of the ozone layer are all complex radical
processes.
? Taking it further
7.1 Introduction
You are now familiar with a good deal of factual functional group
chemistry, and when you have worked through the first six chapters of this
book you will have met many of the basic ideas that underpin organic
chemistry. These ideas help us to relate one aspect of the subject to another.
They allow us to thread the facts together, like beads on a necklace, holding
them together to make a satisfactory pattern. Such patterns should be
recognizable and memorable so that they can be extended and developed in
new and interesting ways.
Underpinning these patterns is a solid base of fact, that is, good, repeatable
experimental data. We need theories, but all good theories are firmly based
on facts. The mechanisms in Chapters 4 to 6 are descriptions of what
happens; in very few cases have we given you any of the experimental data Mechanisms cannot be ‘proved’; they
upon which they are based or explained the interpretation of the data (but are descriptions based on the best
see p.63). ‘How do you know that?’ and ‘How do you show that this available data. As new reliable data
appear, mechanisms are constantly
statement is true?’ are questions you will be asking more often in the future.
modified and updated.
It is in answering these questions that the armoury of techniques of the
modern organic chemist comes into play, from chemical kinetics to nuclear
magnetic resonance (NMR) spectroscopy and X-ray crystallography.
Ho copy = foe es Br |
\
This flight path has important stereochemical consequences if the halo-
alkane is optically active, because it throws the molecule ‘inside out’.
Stereochemistry of addition to carbonyl. As the second example, let’s look
at addition to the carbonyl group (see p. 47). We believe that nucleophiles
either \
attack the carbonyl carbon atom from a direction at right angles to the plane
of the carbonyl group and from slightly behind. This is where the lowest-
energy unoccupied orbital is located. Some of the experimental evidence
for this comes from collecting observations of many different X-ray crystal
structure analyses of compounds which include both a carbonyl group and a
nucleophile close together, and then looking for the most popular relative
positions. These results agree with modern theoretical calculations and with
the current molecular orbital picture of the carbonyl group. Now we can use
this idea to predict the outcome of a new reaction, or to help us to
understand the stereochemistry of other reactions.
Foundations of organic chemistry 85
Enzymes can break the symmetry of simple molecules
When 1-deuterioethanal is reduced by NaBHy, an equimolar (1: 1) mixture
of the two optical isomers of 1-deuterioethanol is produced.
It is equally likely that the H™ will approach the top or bottom face of the
molecule.
a
H
H3C Mago H3C 4 4
=O ==> c
O
1:1 ratio
ec ME
Make yourself some molecular
H tn. csLi H (© tay Aa models of these molecules to help
you to understand the stereo-
7 H chemistry of these reactions.
H
OH
H;C.., enzyme H3Cy,,, 7
D= C— SS D= Ss only
H
The deuterioethanal is bound to the enzyme surface, so that the top and
bottom faces are not identical; the symmetry has been broken. The reducing
H is delivered from a nearby part of the enzyme surface. enzyme
surface
OH
Freya.
D&= q
H
86 Taking it further
Having looked at the extensions of some theories, now let’s look at the
application of some familiar reactions on new molecules.
O O CH 2 CH 3
II i H) ]
~ OCH2CH; N9-"No- NcHcHy
2:
O O
II Il
Ce Cy.
OCH,CH; aq.NaOH OCH,CH;
HOCH,CH;
————
+ + NaCl + H,O
dry HCI H3N H)N
Novocaine is a newer, related local anaesthetic. Work out the last stages
for a synthesis of Novocaine.
NH, O Luminol
NO, O NO, O
I Il
Cc e
és OH NH)NH) at 220° Se
11 — oe |
ee _NH
Il ll
Oo
Ce NH Na ,S 70. ~ NH HO
(iii), oe
a29204
_NH oo NH
HI It
O
Luminol
Cl
(CH3)3CBr + NaOH —— is (CH3)yC=CH, + NaBr + HzO
The biochemistry may alter too, even with quite subtle changes. The two
molecules 2,4-dichlorophenoxyethanoic acid (2,4-D) and 2,4-dichlorophen-
Cl oxybutanoic acid (2,4-DB) look very similar chemically but they are very
different to a plant.
2,4-D 2,4-D stimulates a fatal level of growth in broadleaved plants and is an
excellent weedkiller for narrow-leaved crops like wheat. 2,4-DB itself is
Hy Hy
inactive as a weedkiller, but some broadleaved plants metabolize it by
oxidation to 2,4-D—with fatal effects. Thus 2,4-DB is even more selective
O°. "ey “G00H
than 2,4-D.
cl $
K NH; 4 in
4 \ enzyme H>C NH3 enzyme CH 2
Be ees YS 2 pre ee 4 "
tio coo HxC 2 ~COO™ 3 CH)
Methionine A
H H
\ /
D ne) ND) H Sa
a a m
\. ABN \ Na D D What chemical reaction(s) could you
- Sie or f ie enzyme a use to differentiate between CH2CD2
H Coy D Cia ratio and CHDCHD? Which of the two
geometrical isomers of CHDCHD has
COOm COO 3 x C=C / H the larger dipole moment?
/ SS
H D
Taxol Baccatin IV
Indicator Sa Indicator H
(alkaline solution) (acidic solution)
ORANGE RED
CH, CH;
_ 4%
N NYN Z Ns a
Ne
N
03S N CH3 0,8 N CH3
x
Calcium indicators will do the same with calcium ions, Ca*+. The most
sensitive indicators are fluorescent and show a different fluorescence colour
depending on whether they are bound to Ca’* or not. These indicators are
complex molecules with two overlapping parts: one to bind the Ca**+ and
one to give the fluorescence, e.g. INDO-1.
o0c” “y cae
ie 00C'«
Yas -N “coo ay te ide
' : ‘ : binding
; ' oO! O
aN EQ ‘ae eee ee S ewer:
INDO is a large organic molecule,
and yet it is soluble in water: why?
And to which atoms in INDO do you CH;
think the calcium ion binds? LC Pisin
skt
COO | INDO-1
fluorescence
Calcium ions are very important in
cells such as nerve cells and in the New and better indicators are being synthesized for Ca** as well as for
retina of the eye. Calcium ions are other biologically important ions such as Mg** and Zn?*, which will be very
also involved in the opening and valuable for medical research. The collaboration between the synthetic
closing of the stomata in leaves. ' organic chemists and the physiologists is crucial for progress.
7.5 Conclusion
Almost every section in this book could have been developed further, with
more sophisticated ideas and examples. From time to time you will have
glimpsed the depths and sensed their challenge, a challenge which we hope
you now feel ready to take up.
Further Reading
General A level chemistry: Atkins, P. W., Clugston, M. J., Frazer, M.
J., and Jones, R. A. Y. (1988). Chemistry: principles and applications.
Longman, London.
Undergraduate organic chemistry: Kemp, D. S. and Vellaccio, F.
(1980). Organic Chemistry. Worth, New York.
Biological Aspects: Stryer, L. (1988). Biochemistry. 3rd edn, Freeman,
New York.
Research topics in Chapter 7:
(a) Ethene: Adlington, R. M., Baldwin, J. E., and Rawlings, B. J.
(1983). J. Chem. Soc. Chem. Comm., 290.
(b) Taxol: Potier, P. (1992). Chem. Soc. Rev. 21, 113.
(c) INDO: Grynkiewicz, G., Poenie, M., and Tsien, R. Y. (1985). J.
Biol. Chem. 260, 3440.
Index
SERIES EDITORS
STEPHEN G. DAVIES _ This series of short texts provides accessible accounts of a range of essential topics in
chemistry and chemical engineering, Written with the needs of the student in mind,
RICHARD G. O COMPTON
the Oxford Chemistry Primers offer just the right level of detail for undergraduate
[OHUN EVANS. Study, and will be invaluable as a source of material commonly presented in lecture
courses yet not adequately covered in existing texts. All the basic principles and facts
LVI GUADOEN ina particular area are presented in a clear and straightforward style, to produce
concise yet comprehensive accounts of topics covered in both core and
specialist courses.
Advanced school students and beginning undergraduates will find this book a read-
able and stimulating summary of the fundamentals of organic chemistry. The first
three chapters introduce some basic physical chemistry, and laythe groundwork
for the mechanistic organic chemistry covered later in the book. The importance of
bonding and mechanism are stressed throughout, and students are encouraged to
apply their chemical knowledge in new and unfamiliar situations in order to develop
and sustain their interest, A wide range of examples including natural products and
pharmaceuticals is included, with the final chapter exploring some new develop-
ments and providing an introduction to current research.
Sz
be
¥
ei
si
al
19855 Hy