0% found this document useful (0 votes)
192 views324 pages

Advanced Level Physical Chemistry

The document is a textbook titled 'Advanced Level Physical Chemistry' by A. Holderness, which serves as a revised edition of the physical section of Inorganic and Physical Chemistry. It includes updated content on atomic and molecular theories, chemical equilibria, and energetics, with a focus on modern Advanced Level syllabuses and SI units. The book also features contributions from J.N. Lazonby and is published by Heinemann Educational Books.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
192 views324 pages

Advanced Level Physical Chemistry

The document is a textbook titled 'Advanced Level Physical Chemistry' by A. Holderness, which serves as a revised edition of the physical section of Inorganic and Physical Chemistry. It includes updated content on atomic and molecular theories, chemical equilibria, and energetics, with a focus on modern Advanced Level syllabuses and SI units. The book also features contributions from J.N. Lazonby and is published by Heinemann Educational Books.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 324

A.

Holderness
Advanced Level
Physical Chen stry

TELEPEN

mM
Other books by A. Holderness

ADVANCED LEVEL INORGANIC CHEMISTRY

INTERMEDIATE ORGANIC CHEMISTRY

ORDINARY LEVEL REVISION NOTES IN CHEMISTRY

REVISION NOTES IN ADVANCED LEVEL CHEMISTRY


VOLUME ONE: ORGANIC CHEMISTRY
VOLUME TWO: INORGANIC CHEMISTRY
VOLUME THREE: PHYSICAL CHEMISTRY

and in conjunction with J. Lambert

A NEW CERTIFICATE CHEMISTRY

THE ESSENTIALS OF VOLUMETRIC ANALYSIS

THE ESSENTIALS OF QUALITATIVE ANALYSIS

GRADED PROBLEMS IN CHEMISTRY TO ORDINARY LEVEL

WORKED EXAMPLES AND PROBLEMS IN ORDINARY LEVEL CHEMISTRY

CLASSBOOK OF PROBLEMS IN CHEMISTRY TO ADVANCED LEVEL

PROBLEMS AND WORKED EXAMPLES IN CHEMISTRY TO ADVANCED


LEVEL

ORDINARY LEVEL REVISION NOTES IN CHEMISTRY


Advanced Level
Physical Chemistry
A. HOLDERNESS

A new edition of the Physical section of


Inorganic and Physical Chemistry revised by

J.N. LAZONBY, B.Sc., C.Chem., M.R.LC.


Lecturer, Department of Education, University of York
Formerly Senior Teacher and Head of the Science Department,
Archbishop Holgate’s Grammar School, York
= ae

(*
esSK
‘,

‘S,
Mo va]
4 2

\?: .
MF ~
ec / ?

HEINEMANN EDUCATIONAL BOOKS


LONDON
Heinemann Educational Books Ltd
22 Bedford Square, London WC1B 3HH

LONDON EDINBURGH MELBOURNE AUCKLAND


HONG KONG SINGAPORE KUALA LUMPUR NEW DELHI
IBADAN NAIROBI JOHANNESBURG
EXETER (NH) KINGSTON PORT OF SPAIN

ISBN 0 435 65434 9

© A. Holderness 1961, 1963, 1976


First published as part of Inorganic and Physical Chemistry, 1961
Reprinted 1962
Second Edition 1963
Reprinted 1964 (with corrections)
Reprinted 1966 (with corrections)
Reprinted 1967 (with corrections)
Third Edition (Physical Chemistry section only) 1976 .
Reprinted 1981

Text set in 10/11 pt. Monotype Times New Roman


Printed and bound in Great Britain by
The Pitman Press, Bath, Avon
Preface

Advanced Level Physical Chemistry is a new edition of the Physical section


of Inorganic and Physical Chemistry. At the time of going to press the
Inorganic section is in preparation and will be published at a later date
under the title Advanced Level Inorganic Chemistry.
In addition to attention to units and nomenclature, the revision of the
Physical section has required a regrouping, retitling, and rewriting of some
of the original chapters in order to produce a more coherent treatment
which is consistent with modern Advanced Level Syllabuses. New material
has been introduced, particularly in the parts of the book which deal with
electron arrangements, crystal structures, and energy changes.
The policy on units has been to adopt SI units entirely, except in such
cases as the use of the convenient mmHg alongside N m~? for pressure. In
making decisions about nomenclature, the recommendations of the
Association for Science Education and the Examining Boards have been
taken into account.
Thanks are due to Mr Martyn Berry, Mr J. P. Chippendale, Dr John
Dyson, and Mrs Freda Stevens for their helpful advice on the content and
editorial work on the manuscript and proofs. ee

1976
Contents

PREFACE

1 Classical Atomic and Molecular Theories


Moles and Molarity
Fundamental Particles and Their Arrangement in Atoms
The Nucleus
Electrons in the Atom
The Periodic Classification of the Elements
Bonding I: Ionic and Covalent
Crystal Structure
Oo
WH
W
LP
na
annBonding II: Intermolecular
Colloids and the Colloidal State 105
Physical Behaviour of Gases: The Kinetic Theory of Gases 113
Phase Equilibria 131
Energetics 178
Chemical Equilibria I: Molecular Equilibria 194
Chemical Equilibria Il: Ionic Equilibria 208
Chemical Equilibria III: Redox 243
eeOS
eeBh
HA
HD
“JI
NY
W
K Rates of Chemical Reactions 254
18 Electrolysis and Conductance 269
QUESTIONS 285
TABLE OF RELATIVE ATOMIC MASSES 308
INDEX 309
1. Classical Atomic and Molecular
Theories

Elements, Compounds, and Mixtures


Late in the eighteenth century, sufficient chemical knowledge had
accumulated to allow scientists to recognize elements and compounds as
different species of substances. At this period an element was defined as
follows:

An element is a substance which cannot be decomposed into simpler


substances by any known chemical process.

At the same period, the following definition was given for a chemical
compound:
A compound is a material of uniform composition throughout its bulk
and containing two or more elements ina state of chemical combination.

It was clearly recognized that, during chemical combination, the


elements concerned lost their own characteristic chemical properties and
that the compound possessed a new set of chemical properties of its own.
For example, sodium is a very reactive metallic element which attacks
water violently at ordinary temperature. Chlorine is a poisonous, greenish-
yellow, gaseous element. The ‘chemical combination’ of these two produces
sodium chloride (common salt), which is a white solid, non-poisonous, and
soluble in water without reaction with it. Scientists of the early nineteenth
century had no idea of the real nature of ‘chemical combination’, and, in
fact, its interpretation in essentially electrical terms has been possible only
in the twentieth century. This interpretation is developed to some degree
in Chapter 7.
The two most important characteristics of compounds are:

1. All pure samples of a given compound are identical in composition by


mass. (This statement disregards the occurrence of isotopy. For a
discussion of this phenomenon, see page 29.)

2. Chemical action is usually needed to separate the constituent elements


from a compound. For example, to obtain hydrogen (element) from its
2 Physical Chemistry
compound water, the chemical action of sodium on water or of heated
magnesium or iron on steam is required.
2Na(s) + 2H,O() > 2NaOH(aq) + H.(g)
Mg(s) + H,O(g) > MgO(s) + H:(g)
3Fe(s) + 4H,O0(g) = Fes O.(s) + 4H2(g)
Mixtures can be made from elements or compounds or both. In mixtures,
the constituents merely lie mingled together. Mixtures of given constituents
can usually be made widely variable in composition and the various
constituents can always be recovered by physical action alone. For
example, iron filings and sulphur can be mixed in any desired proportion
and the sulphur can be recovered by solution in carbon disulphide,
filtration, and evaporation of the solvent. This is a succession of purely
physical processes. In addition, mixing is usually a quiet process; by
contrast, chemical combination is often accompanied by large energy
changes which exhibit themselves as the evolution of heat, explosion, or
flame.

Laws of Chemical Combination by Mass


Around the beginning of the nineteenth century, two laws, which were
based entirely on the results of experiments, were known to govern
chemical combination.

1. The Law of Conservation of Mass


This states (in its chemical aspect) that:
The total products ofagiven chemical reaction have the same mass as the
total reactants involved in the reaction.
The most convincing experimental evidence for this law was supplied by
Landolt in 1906.
Solutions of pairs of chemical reagents were introduced separately into
the two arms of a Landolt tube (Figure 1.1). The tube was sealed off by

Figure 1.1, Landolt tube


Classical Atomic and Molecular Theories 3
heating and its mass was found at room temperature. The reaction was
induced by inverting the tube and allowing the solutions to mix. The mass
of the tube containing the products was then determined. Allowance had
to be made for two main possibilities of error which arose from the slight
evolution of heat during the reaction. This heat caused loss of moisture from
the surface of the tube and expansion of the tube, so that it displaced a
greater volume of air. Both these factors tended to decrease the mass of the
tube. The errors were eliminated if sufficient time was allowed for the tube
to regain its original state. Landolt investigated fifteen chemical reactions,
of which two typical cases were:

Ag.SO,(aq) + 2FeSO,(aq) — 2Ag(s) + Fe(SO.)s(aq)


SHI(aq) + HIO;(aq) — 3H,O(1) + 31,(aq)
He found that the change of mass during the reactions did not exceed one
part in ten million, which was well within the limit of experimental error.
It should be noted that, according to modern ideas, emission of energy
involves a loss of mass according to the famous Einstein equation, E =
mc?, where E is the energy, m the mass, and c the velocity of light (in
appropriate units). Consequently, an emission of energy during chemical
action can cause loss of mass and so interfere with the above law. In
practice, however, the loss of mass is so very small as to be negligible.
Even so the true conservation is one of mass and energy together.

2. The Law of Constant Composition (or Definite Proportions)


All pure samples of a given chemical compound contain the same ele-
ments in the same proportions by mass.
This experimental law was put forward by Proust in 1799. It was strongly
disputed by Berthollet, but convincing experimental evidence such as that
obtained by Stas finally led to its acceptance. In one particular case, Stas
prepared silver chloride by heating silver in chlorine, and by dissolving
silver in nitric acid and precipitating the chloride by ammonium chloride
and by hydrochloric acid. In the three products, 100 parts by mass of silver
combined with 32.8425, 32.842 and 32.8475 parts by mass of chlorine. _
In more recent years, the phenomenon of isotopy has been recognized
and is relevant to this law. Isotopy is the existence of atoms of the same
chemical properties but different relative atomic mass. For example, lead
associated with uranium minerals has a relative atomic mass of 206, but
lead associated with thorium minerals has a relative atomic mass of 208.
It is obvious that compounds of these isotopes with another element
cannot have a constant composition. This is an exceptional case. Usually,
elements encountered in chemical practice contain isotopes in fixed
proportions and so act as if their relative atomic masses are completely
constant.
4 Physical Chemistry

Berthollide compounds
Deviations from the Law of Constant Composition corresponding to
Berthollet’s ideas have been studied in solids since about 1930. In a solid,
AB,,, the ideal situation is to have every particle (ion or atom) of A placed
in its appropriate point in the solid lattice (page 86) and every such point
occupied; and similarly for B. Such a perfect arrangement can be expected
only at 0 K(kelvin), where thermal vibration of particles is absent. At other
temperatures lattice defects may occur, arising from such vibrations.
These defects may be produced in three ways: (a) by interchange of
particles of A and B, (b) by migration of particles of A or B away from
their correct locations in the lattice to interstitial (see page 90) positions
(these are called Frenkel defects), and (c) when locations are left vacant by
migration of particles of A or B to the surface of the material (called
Schottky defects).
Type (a) is very unlikely to occur in ionic compounds, e.g. KC1, because
of energy factors involved, but is found in metallic alloys. At ordinary
temperatures, ionic compounds, e.g. KCl, CaO, show no observable
lattice defects and conform perfectly to the Law of Constant Composition.
As the melting point is approached, however, defects may become more
appreciable. For example, potassium chloride at 1000K has about
0.004 per cent of lattice vacancies.
Iron(I]) sulphide, ideally FeS, has been shown to vary between FeS,o
and FeS, 34, or in reverse, Fe, 99S and Fe gS. This variation is caused by
lattice deficiency in iron, that is, Fe,.99S to Fep.ggS is the more correct
representation. This is known to be so because the material decreases in
density (by loss of Fe) as the proportion of sulphur increases. If sulphur
were added interstitially to true FeS, an increase in density would be
expected. The relative loss of iron is compensated electronically by the
conversion of Fe?* cations to Fe**, so that, in a sense, the product has
the composition of a continuously variable mixture of FeS and Fe,S3.
Similarly, iron(II) oxide has been shown to vary between FeO,9, and
FeO,19 at 1700 K. As in the case of iron(II) sulphide, this arises from the
omission of some iron from the lattice and the conversion of some Fe2* to
Fe®* to compensate.

The Atomic Theory of Dalton


Against this chemical background of the early nineteenth century, John
Dalton in 1808 was able to produce the set of postulates known collectively
as the Atomic Theory. Atomic ideas have, however, a very long history.
Dalton’s fame rests on the fact that he converted atomic ideas from
‘ineffectual thoughts casually entertained’ (Whitehead) into a set of
definite ideas which could be (a) used to explain known experimental
observations and (b) used to make predictions which could be subjected to
Classical Atomic and Molecular Theories 5
experimental test. These ideas, which constitute the classical Atomic
Theory, are:
1, All elements are made up of small particles called atoms.
An atom is defined as the smallest particle of an element which can
take part in a chemical reaction.
. Atoms cannot be created or destroyed.
. Atoms are indivisible.
. Atoms of the same element are exactly alike in every way, notably in
kWh
mass.
=p Atoms combine together in small whole numbers.
Some of these ideas have had to be restated in the light of more recent
discoveries.
The Atomic Theory clearly explains the two experimental Laws of
Chemical Combination, which were known at that time:
ke Law of Conservation of Mass (page 2). The Atomic Theory stated
that atoms cannot be created or destroyed. From this postulate it
follows that all the atoms present at the beginning of a chemical
reaction should be present at the end of it, with no gain or loss of
material. That is, the total mass of the products of a chemical reaction
should be equal to the total mass of the reagents used.
. Law of Constant Composition (page3). The Atomic Theory postulated
that all the atoms of the same element are exactly alike. From this the
conclusion can be drawn that pure samples of a given chemical com-
pound, containing atoms of the same elements in the same proportions,
must be identical in composition by mass.
Two more Laws of Chemical Combination can be deduced from the
Atomic Theory.
ai The Atomic Theory postulated that all atoms of the same element are
exactly alike and that the atoms combine together in small whole
numbers. If the second postulate is true, compounds between the
elements A and B must be of the type AB, A.B, ABz, A2Bs, ete. If all
the atoms of A are exactly alike, the amount ‘A’ (one atom of A) must
be a fixed amount. The amounts of B which combine with this fixed
amount A in these respective compounds are B, B/2, 2B, 3B/2, etc.
These are in the proportions 2B:B:4B:3B. If all the atoms of B are
alike, these must be simple, whole-number proportions. This conclusion
can be tested by analysing sets of compounds such as copper(I) and
copper(II) oxides, mercury(I) and mercury(II) chlorides, and the three
oxides of lead, and is found to be true. It is expressed in the form known
as the Law of Multiple Proportions:
If two elements, A and B, combine to form more than one compound,
the various masses of B, which combine with a fixed mass of A, are ina
simple whole-number ratio.
6 Physical Chemistry
Dalton himself was the first person to verify this law experimentally in
about 1804.
4. The second conclusion from the above two postulates describes the more
complicated relations of a number of different elements, C, D, E, F,
etc., with an element, A. If atoms combine in small whole numbers, the
compounds formed by these elements with A and with each other must
be of the simple types AC, AgC, AC, AD, ADz, CD, CDs, EFs, etc.
From this it follows that, if all atoms of the respective elements are
exactly alike, the proportions in which C, D, E, F, etc. combine with a
fixed mass of A are also the proportions in which C, D, E, F, etc.
combine with each other, or are simple multiples of those proportions.
This is known as the Law of Reciprocal Proportions:
The masses of elements C, D, E, F, etc. which combine with (or dis-
place) a fixed mass of element A, are the masses of C, D, E, F, etc. which
combine with (or displace) each other, or are simple multiples or sub-
multiples of these masses.
The following figures illustrate the law. Let the fixed mass of oxygen be
8g. The masses of other elements which combine with the 8 g of
oxygen are zinc, 32.5 g; sulphur, 8 g (in SO,); hydrogen, 1 g. The law
says that, if zinc, sulphur or hydrogen combine with (or displace) each
other, they must do so in the proportions of 32.5:8:1 or some simple
multiples of these. Actually, 32.5 g of zinc combine with 16 g of sulphur,
1 g of hydrogen combines with 16 g of sulphur, and 1 g of hydrogen is
displaced by 32.5 g of zinc, as the law requires.
The most convincing experimental verification of this law was
published by Berzelius in 1812.
Dalton’s Atomic Theory could not, of course, be directly verified
because the atoms were quite inaccessible to the measuring instruments of
the day. However, the theory provided a satisfactory explanation of the
known experimental laws and led to the prediction of others, which were
successfully tested by experiment.

Combining masses
When carrying out an investigation into how two elements are involved in
a particular reaction, quite often the easiest experimental data which.can
be obtained is the mass of one element which combines with a certain mass
of the other element. These masses are called combining masses or, when
referred directly or indirectly to the common standard of one part by mass
of hydrogen, they are called equivalent masses. (It is sometimes more
convenient to use either oxygen or chlorine as the standard when deter-
mining the equivalent mass of an element, in which case 8 parts by mass of
oxygen or 35.5 parts by mass of chlorine must be used as these are the
masses which combine with 1 part by mass of hydrogen.)
Classical Atomic and Molecular Theories 7
Relative atomic masses
Dalton used the combining masses of elements to determine values for
their relative atomic masses. He tried to compare the mass of one atom of
each element with the mass of one atom of hydrogen. (Hydrogen was
chosen as the reference element as its atom has a smaller mass than atoms of
other elements.) In order to make this comparison, it was necessary for
Dalton to assume that atoms of different elements combine in the simplest
possible ratios.
For example, when it was found that 8 g of oxygen combined with 1 g of
hydrogen, it was assumed that 1 atom of oxygen combined with 1 atom of
hydrogen and that these combining masses represented equal numbers of
atoms. Therefore, the atomic mass of oxygen relative to hydrogen was
found to be 8. Obviously, to obtain the correct value, it was necessary to
know that the formula of water is HO and that 1 g of hydrogen represents
twice as many atoms as 8 g of oxygen, from which it can be deduced that
the atomic mass of oxygen relative to hydrogen is 16.
The lack of knowledge of the ratios in which atoms combine led to the
determination of conflicting results for the relative atomic masses of
elements.
It was not until Cannizzaro (1858) realized the full significance of the
contributions of Gay-Lussac and Avogadro to the molecular theory of.
gases that this problem, for elements such as carbon, nitrogen, and
oxygen, was overcome.

Molecular Theory of Gases


The development of the molecular theory began with the following
experimental laws.

Boyle’s Law (1662)


The volume of a fixed mass ofgas is inversely proportional to its pressure,
temperature remaining constant.

Charles’ Law (1787)


The volume of a given mass of gas is directly proportional to its absolute
temperature, pressure remaining constant,

Gay-Lussac’s Law of Gaseous Volumes (1808)


When gases react, they doso in volumes which bear a simple ratio to one
another and to the volume of the product, if it is a gas, temperature
and pressure remaining constant.
This law is illustrated by experimental observations, such as that, at
constant temperature and pressure:
8 Physical Chemistry
2 volumes of hydrogen combine with 1 volume of oxygen to give 2
volumes of steam;
2 volumes of ammonia are formed from 3 volumes of hydrogen and 1
volume of nitrogen.
A notable feature of these laws is that they apply (with very minor
variations) to all gases, irrespective of whether the gases are elements or
compounds and, also, of their chemical nature, which can be very variable.
There must be some explanation of this universal similarity of gaseous
behaviour. Under the influence of the atomic ideas which pervaded
chemistry in the early nineteenth century, Berzelius (and others) put
forward the hypothesis that the factor common to all gases is the occur-
rence of equal numbers of atoms in equal volumes of all gases at constant
temperature and pressure. This idea proved unacceptable for reasons of
which the following is typical.
By experiment, 1 volume of hydrogen combines with 1 volume of
chlorine to give 2 volumes of hydrogen chloride at constant temperature
and pressure. If the idea of Berzelius is accepted, this statement can be
rendered in the form:
n atoms of hydrogen combine with n atoms of chlorine to give 2n
‘compound atoms’ of hydrogen chloride.
This simplifies to: Z
1 atom of hydrogen combines with 1 atom ofchlorine to give2 ‘compound
atoms’ of hydrogen chloride.
It is obvious that one atom of hydrogen cannot contribute to two
‘compound atoms’ of hydrogen chloride unless the hydrogen atom can be
split. But Dalton’s Atomic Theory required the atom to be indivisible so
the Berzelius idea was abandoned.
It was replaced in 1811 by the very famous and important idea usually
known as Avogadro’s Hypothesis. This states that:
Equal volumes of all gases at the same temperature and pressure contain
the same number of molecules.
It is important here to be clear about the nature of the atom and the
molecule. They are defined in the following way:
An atom of an element is the smallest particle of the element which can
take part in a chemical reaction.
Amolecule ofan element or compound is the smallest particle of it which
can exist separately.
The two are not necessarily the same particle in the case of any given gas
and, in particular, the molecule may be polyatomic and so capable of
Classical Atomic and Molecular Theories 9
splitting into individual atoms as required. This recognition of the occur-
rence of the two different particles, the atom and the molecule, removed the
difficulty which destroyed the atomic idea of Berzelius discussed above.
The following is an example of the way in which Avogadro’s Hypothesis
provides a satisfactory explanation of Gay-Lussac’s Law.
By experiment:
1 volume of hydrogen combines with 1 volume of chlorine to give 2
volumes of hydrogen chloride.
Applying Avogadro’s Hypothesis we can say:
n molecules of hydrogen combine with n molecules of chlorine to give 2n
molecules of hydrogen chloride.
Simplifying:
1 molecule of hydrogen combines with 1 molecule of chlorine to give 2
molecules of hydrogen chloride.
1 molecule of hydrogen chloride must contain 4 molecule of hydrogen;
therefore 1 molecule of hydrogen must contain at least 2 atoms of hydro-
gen.
Chemical evidence such as the fact that hydrochloric acid only forms one
series of salts (unlike sulphuric acid which forms two), indicates that
hydrogen chloride contains 1 atom of hydrogen per molecule and, there-
fore, 1 molecule of hydrogen must contain 2 atoms.
By a somewhat similar argument it can be shown that the molecule of
chlorine contains 2 atoms, and the reaction between hydrogen and
chlorine may be diagrammatically represented by Figure 1.2 at constant
temperature and pressure.

thet
8 fs pa
8 8 8
Figure 1.2.

1 volume 1 volume 2 volumes


6H.(g) + 6Cl.(g) => 12HC\(g)
or in simplest terms,
H.(g) - Cl,(g) oe 2HC\(g)

Thus it can be seen that the application of Avogadro’s Hypothesis has


produced a theory for this reaction which is consistent with the experiment-
al evidence.
10 Physical Chemistry
Since their molecules contain 2 atoms each, hydrogen and chlorine are
said to be diatomic, the term atomicity being defined as follows:
The atomicity of an element is the number of atoms contained in 1
molecule of the element.

The importance of Avogadro’s Hypothesis


The importance of the hypothesis lies in the fact that, since it asserts that
equal volumes of gases contain equal numbers of molecules, it enables us
to change over directly from a statement about volumes of gases to the
same statement about molecules of gases. Every time we make a statement
about one volume of a gas, we are also making a statement about a certain
number of molecules of it, and that number, by Avogadro’s Hypothesis, is
always the same, no matter what the gas may be. Consequently we can
change over at will, in any statement about gases, from volumes to
molecules and vice versa, if the temperature and pressure are constant.
This means that by applying the hypothesis to volume measurements of
gases, we can probe right to the heart of a chemical reaction, to the actual
molecules themselves. It is an enormous step to change directly from an
experimental statement like:
2 volumes of hydrogen combine with 1 volume of oxygen giving 2
volumes of steam (temperature and pressure constant)
to:

2 molecules of hydrogen combine with 1 molecule of oxygen giving 2


molecules of steam.
Another aspect of the link between volumes and molecules of gases is
that if the masses of equal volumes of two gases are compared, then it is
also a comparison of the masses of one molecule of each of the gases, i.e.
mass of 1 volume of gas A _ mass of 2 molecules of gas A
mass of 1 volume of gas B mass of n molecules of gas B
__ mass of 1 molecule of gas A
~~ mass of 1 molecule of gas B

Cannizzaro and Relative Atomic Masses


It was not until the work of Cannizzaro (1858) that the full significance of
the above statements was realized and that Avogadro’s Hypothesis was
generally accepted. Cannizzaro recognized the possibility of comparing the
molecular masses of gases by comparing their densities. Using hydrogen as
the reference element, he deduced a relationship between the density of a
Classical Atomic and Molecular Theories 11
gas, relative to hydrogen (relative vapour density), and its molecular mass
relative to the mass of one atom of hydrogen (relative molecular mass).

mass of a volume of gas


melative vapour densityse<|——— OE
mass of equal volume of hydrogen
at constant temperature and pressure.
Applying Avogadro’s Hypothesis, we can say directly:
mass of 1 molecule of gas
relative: vapour density) = ———__—___—— eee
mass of 1 molecule of hydrogen
But hydrogen is diatomic, and so
mass of 1 molecule of gas
relative vapour Gcisity = ————_—_—_——
mass of 2 atoms of hydrogen
mass of 1 molecule of gas
_____—_ =".
2 x relative vapour density = ———_____—
P z mass of 1 atom of hydrogen
= relative molecular mass of the gas
i.e. the relative molecular mass of a gas is twice its relative vapour density.
Therefore, whenever it is possible to determine experimentally the relative
vapour density of a gas (see page 125 for experimental methods), it is
possible to determine its relative molecular mass. Cannizzaro used the
relative molecular masses of gases, determined in this manner, in his very
important method for the determination of relative atomic masses. He
realized that if he selected a sufficiently large number of compounds of a
particular element, then it was likely that at least one of those compounds
would contain only one atom of the element per molecule. He then
employed the following steps in order to determine the relative atomic mass
of the element.
1. He determined the relative vapour density of each compound containing
the element.
2. He doubled the relative vapour density and thus obtained the relative
molecular mass of each compound.
3. He analysed each compound for the percentage of the element by mass.
4. Using the results from 2 and 3 he calculated the mass of the element in
one relative molecular mass of each compound.
The table on page 12 illustrates application of the method to carbon.
The figures in the last column correspond to the presence of one, two,
three or more carbon atoms per molecule. The lowest mass is 12 and the
others are multiples of 12. Now it is obvious that if the list contains any
compound containing only one carbon atom per molecule, that compound
will be the first, methane, because in this the mass of carbon is the least. If,
therefore, the molecule of methane does actually contain only one carbon
12 Physical Chemistry
atom, the relative atomic mass of carbon is 12. This process has been
applied to a very large number of carbon compounds, and the mass of
carbon in the relative molecular mass has always been found to be 12, ora
multiple of 12, but never less. From this it is concluded that the least mass
of carbon there can ever be in the relative molecular mass of one of its
compounds is 12, that this mass corresponds to the presence of one
carbon atom, and that the relative atomic mass of carbon is 12.

The following table illustrates the application of the method to carbon.

: Relative
Relative molecular pte Mass of carbon
Compound eae mass nehes if in the molecular
ried wore (=2 x vapour oaks (by mass
experiment) density) experiment)

16
Meth
ethane 8 16 75.0 £5:
5.0.x npn
700 1 ofp

Ethane 15 30 80.0 80.0 x Des 24


100

Propane 22 44 81.8 81.8 x pate 36


\ 100
Ethene ; 28

Ethyne 26
(acetylene) 13 26 92.3 92:3. % ies 24

The method can be applied to determine the relative atomic masses of


any element forming a large number of gaseous or easily vaporized
compounds. It thus became possible, for the first time, to determine
reasonably accurate values for the relative atomic masses of such im-
portant elements as oxygen, carbon, nitrogen, and sulphur. In effect, the
work of Gay-Lussac, Avogadro, and Cannizzaro had overcome the
problem of not knowing the ratios in which atoms of elements such as
these combine. It was the lack of knowledge of these ratios which had
caused so much trouble to Dalton and others in the early part of the
nineteenth century (see page 7).
When comparing the combining ratios of different elements it was
found to be useful to employ hydrogen as the reference element. Thus if an
element of symbol M combined with hydrogen to give a compound of
formula MH,, then 7 was known as the combining power or valency of the
element. If the relative atomic mass of M is A, it follows that in this
Classical Atomic and Molecular Theories 13
compound A g of M must combine with n g of hydrogen (since the relative
atomic mass of hydrogen is, by definition, 1). Therefore 1 g of hydrogen
must combine with A/ng of M and this value is, by definition, the
equivalent mass of M. This leads to the relationship
relative atomic mass = equivalent mass x valency
or
A=Alnxn
Thus if the valency of an element can be determined, the above relationship
can be used to convert the equivalent mass into the relative atomic mass.
However the relationship must be applied with caution, as certain elements
can exhibit more than one combining power. Therefore it is essential that
the combining power used is the same as that exhibited by the element in
the compound from which its combining mass is determined.

Other Methods for the Determination of Relative Atomic Masses


1. Using Dulong and Petit’s Law
The above relationship was used, along with the experimental law of Dulong
and Petit (1820), to determine the relative atomic masses of certain
metallic elements. Dulong and Petit observed that a relationship existed
between the relative atomic masses and specific heat capacities of most
solid metallic elements. The Dulong and Petit’s Law, stated in a modern
form, is:
For solid elements near room temperature, relative atomic mass X
specific heat capacity is approximately constant at 27 J g-' Kt.
This law has been used to determine relative atomic masses of metals in
the following way.
Consider indium as a newly discovered metal. Its equivalent mass was
found to be 37.8; its specific heat capacity (by experiment) was 0.24
J g-1 K~1. By Dulong and Petit’s Law,
relative atomic mass x 0.24 = 27 (approx.)
; ; on
1.¢. relative atomic mass = BY 110 (approx.)

But relative atomic mass = equivalent mass X valency

.. valency = ae = 3 (must be an integer)


37.8
accurate relative atomic mass = equivalent mass x valency
= 37 .8)%3
= 113.4
14 Physical Chemistry
Any divergence of the valency from an integer arises from the fact that the
law of Dulong and Petit is only approximate and can only give an approxi-
mate value for a relative atomic mass.

2. Using the Law of Isomorphism


The phenomenon of isomorphism has proved useful for determining
relative atomic masses and, more particularly, for correcting them. As
first put forward in 1819 by Mitscherlich, the Law of Isomorphism stated:
Substances of similar chemical character, having equal numbers of atoms
combined in a similar way, show identity of crystalline form.
It is now known that this statement is too precise. No two substances have
exactly the same crystalline form. In cases of general similarity of crystalline
form, there are always minute angular differences. The law is, however,
quite useful if it is taken in the looser form of an assertion such as:
Substances of corresponding chemical composition often show similarity
in crystalline form.
That is, such substances are often isomorphous, showing at least two of the
following characteristics: a general similarity of crystalline shape, forma-
tion of mixed crystals, formation of overgrowths on each others’ crystals.
Isomorphism has proved useful in cases such as the following. In the
early nineteenth century silver was taken to be divalent, its'sulphide was
taken to be Ags, and its relative atomic mass to be 216 (H = 1). Dumas
showed copper(I) sulphide, Cu,S, and silver sulphide to be isomorphous.
The formula of silver sulphide was then corrected to Ag.S and the relative
atomic mass of silver to 108.
The atomic mass of selenium was determined from the fact that
potassium sulphate and potassium selenate are isomorphous. By analogy
with the sulphate, K,SO,, the selenate must be represented as K.SeQ,.
The percentage of selenium in the selenate was found to be 35.75 per cent.
From this,
Se
Kisco, x 100= 35.75

Inserting the relative atomic masses K = 39 and O = 16, we have


Se
————~ xX 100 = 35.
Se + 142 * aa
Solution of this equation yields the result Se = 79.

3. Using a mass spectrometer


A mass spectrometer provides the most accurate method for determining
relative atomic masses and its use has now superseded all other methods.
Classical Atomic and Molecular Theories 15
After some initial experiments by Thompson, the first mass spectrometer
was constructed by Aston in 1919. Aston’s mass spectrograph and more
modern versions of the instrument are discussed in some detail on page 30.

A summary of the Development of the Methods for the


Determination of Relative Atomic Masses

Laws of
Chemical
Combination
(pre-1808)

Dalton’s
Atomic Gay-Lussac’s
Theory Law (1808)
(1808) ih
Avogadro’s
Hypothesis
Simplest Combining Laws of (1811)
formulae masses Dulong and Petit,
Mitscherlich
(1819)

Dalton’s
atomic
masses

Cannizzaro-
relative
molecular
masses
(1858)

Aston’s
mass
spectograph
egies (1919)
RELATIVE ATOMIC MASSES

Lists of relative atomic masses are given on page 308.


16 Physical Chemistry
Standards for Relative Atomic Masses
The accepted reference element for the determination of relative atomic
masses has varied. In the early days, hydrogen, being the element with
atoms of least mass, seemed to be the obvious choice. Towards the end of
the nineteenth century, the methods used for determining relative atomic
masses largely depended on finding the combining masses of elements.
Therefore, as more elements combine directly with oxygen than with
hydrogen, it was decided to fix the atomic mass of oxygen at 16 and to use
this as the standard.
Finally, with the realization, during the first half of this century, that a
lot of elements consist of mixtures of atoms of slightly different masses
(isotopes, see page 29), it became necessary to select a particular isotope
as the standard. In 1961 it was internationally agreed that the isotope of
carbon, known as !7C, should have an atomic mass of 12 and that this
isotope should always be used as the standard. #2C was chosen as it is the
most convenient standard, from an experimental point of view, to use
when determining relative atomic masses with a mass spectrometer.
2. Moles and Molarity

The availability of a reliable set of relative atomic masses has enabled


chemists to use reacting masses to establish formulae for compounds and
equations for reactions. Also, once an equation is known, it is possible
to predict reacting masses and the masses of products. In their modern
form, these calculations involve the unit amount of substance known as a
mole.
The approximate relative atomic mass of oxygen is 16 on the H = 1
scale, which means that each atom of oxygen is 16 times heavier than 1
atom of hydrogen (ignoring, for'the moment, the existence of isotopes).
It follows from this that any masses of these elements, which are in the
ratio of 16:1, must contain equal numbers of atoms. For example, 16 g of
oxygen must contain the same number of atoms as | g of hydrogen. By
similar arguments, it is possible to see that the relative atomic mass of any
element, expressed in grams (a g-atom), will contain the same number of
atoms as | g of hydrogen.
This constant number of atoms is called the Avogadro Constant and is
given the symbol L. The amount of an element containing the Avogadro
Constant number of atoms is called 1 mole of atoms of the element.
In addition to atoms, chemical reactions can involve molecules of
elements, molecules of compounds, ions, or electrons. It is therefore
useful to extend the definition of a mole to cover these other species. For
example, 1 mole of oxygen molecules (O,) contains 2 moles of oxygen
atoms and has a mass of 32 g.
The modern definition of a mole on the ??C = 12 scale is:
1 mole is the amount of a substance containing the same number of
particles (atoms, molecules, ions etc.) as there are atoms present in 12 g
of the isotope of carbon, ?2C,

This number, which is the accepted value for the Avogadro Constant, is
6.02 x 107%, e.g. 6.02 x 107° molecules of carbon dioxide (CO,) contain
6.02 x 1022 atoms of carbon (1 mole-mass 12 g) and 2 x 6.02 x 10°
atoms of oxygen (2 moles—mass 32 g). Therefore 1 mole of carbon dioxide
has a mass of 12 + 32 = 44g.
The mass of 1 mole of a substance is thus the formula mass of the
substance.
18 Physical Chemistry
The following table gives a few simple examples of the relationships
between masses and numbers of moles.

Mass of Mass of substance Number of moles


ets 1 mole involved involved

Na (atoms) 23 g 57.5 g 2.5


Cl, (molecules) 73 g 7.3 g 0.1
H,O (molecules) 18 g 0.9 g 0.05
NaCl (formula units) 58.5 g 11.7 2 0.2
CO2- (ions) 60 g 1.8 g 0.03

When referring to 1 mole of a compound which exists in the form of a


giant lattice (see page 86), it is misleading to refer to 1 mole of molecules,
as individual molecules (represented, for example, by the formula NaCl)
do not exist. A more satisfactory term is 1 mole of ‘formula units’.
A large proportion of chemical reactions involve aqueous solutions of
reagents and it is useful to express the concentration of these solutions in
terms of the number of moles of solute dissolved in 1 dm? of solution.
This is known as the molarity of the solution.
The simple examples in the following table will serve to illustrate the
relationships between mass dissolved, volume of solution, and molarity of
solution.

Foraite Mass of Mass of solute Volume of Molarity of


1 mole dissolved solution solution

4.0
NaOH 40 g 40g 1000 cm® 0 = 0.1M

HSO, %%¢ 49g 250 em? ee ee Hi


15.9 1
Na,CO; 106g 15.9 g 2000 cm® aa ~ are 0.075 M

Due to the direct relationship between moles and numbers of particles


(atoms, molecules, orions), the mole concept is of fundamental importance.
It is obviously of more interest to know the numbers of particles involved in
a reaction, rather than just the reacting masses. The following worked
examples illustrate some of the more common uses of the mole in quan-
titative chemistry.
Moles and Molarity 19
Bs ee the Formula of a Compound from its Composition
Y
Hydrated magnesium sulphate contains 9.8 per cent magnesium, 13.0
per cent sulphur, 26.0 per cent oxygen, and 51.2 per cent water of crystal-
lization. (Relative atomic masses, Mg = 24, S = 32,0 = 16, H = 1.)

Mg S) O H,O
Percentage by mass 9.8 13.0 26.0 mf io
9.8 13.0 26.0 Sie
Number of moles 74 37, a Ga cica

=0.408 =0406 =1.63 = 2.84

Divide each by the 0.408 0.406 1.63 2.84


smallest 0.406 0.406 0.406 0.406
— — — 4 =

The formula of hydrated magnesium sulphate is MgSO,,7H,O.


This calculation gives the empirical formula of the compound, i.e. the
simplest formula which expresses its composition by mass. In some cases
the molecular formula of a compound is not the same as its empirical
formula. This problem mainly arises with organic compounds. To decide
this point it is necessary to find the relative molecular mass of the
compound. For example, the empirical formula of ethane can be shown to
be CH, and, as its relative molecular mass is found to be 30, its molecular
formula is C.Hg.

To Determine the Equation for a Reaction


By titration, 25.0 cm® of a 0.1 M solution of sodium hydroxide was found
to react completely with 28.2 cm® of a solution of sulphamic acid (NH,SO3)
containing 8.54g¢ of the acid per dm® of solution. (Relative atomic
masses, N = 14, H = 1, S = 32, O = 16.)
To find the number of moles of NaOH in 25.0 cm® of solution:
1000 cm? of solution contains 0.1 moles

25.0 cm? of solution contains 0.1 x acs it


= 0. praee
1000»
To find the number of moles of sulphamic acid in 28. dem? ofthesollition’s,
The mass of1 mole of the acid is 14 + 3 + 32 + 48 = 97g

1000 cm® of the acid contains 37Ba


mlhe Lee
20 Physical Chemistry

28.2 cm® contains 8.54


97
28.2
1000
= 0.00248 moles
From this it can be seen that 1 mole of sulphamic acid will react completely
with 1 mole of sodium hydroxide and that the left-hand side of the
equation is
NH,SO,(aq) + NaOH(aq) >
The products are likely to be a salt plus water and the complete equation
for the reaction is probably
NH,SO.(aq) + NaOH(aq) > NaNH.SO,(aq) + H,0()

Gravimetric Calculation From an Equation


Calculate the mass of nitric acid necessary to react with 10 g of lead(II)
oxide and the mass of lead(II) nitrate formed in the process. (Relative
atomic masses, Pb = 207, N = 14,O = 16, H = 1.)
The equation for the reaction is

PbO(s) + 2HNO,(aq) > Pb(NOs)2(aq) + H20()


From the equation,
—". 1 mole of lead(II) oxide reacts with 2 moles of nitric acid!
The mass of 1 mole of lead(II) oxide = 207 + 16 = 223 g
The mass of 1 mole of nitric acid = 1 + 144+ 48 = 63g
Therefore 223 g of lead(II) oxide reacts with 2 x 63 g of nitric acid
and 10 g of lead(II) oxide reacts with 2 x 63 x a g= 5.65 ¢

i). 1 mole of lead(I]) oxide produces 1 mole of lead(II) nitrate.


The mass of 1 mole of lead(ID) nitrate = 207 + 28 + 96 = 331 g
Therefore 223 g of lead(II) oxide produces 331 g of lead(II) nitrate
and 10 g of lead(II) oxide produces 331 x on g=148¢g

The examples in this chapter illustrate the fundamental importance of


the mole as a unit amount of substance. As previously stated, chemists are
interested in the numbers of particles involved in reactions, and their way
of counting particles is to convert the masses of solids, or the volumes of
solutions of known concentration, or the volumes of gases at known
temperatures and pressures, into numbers of moles.
3. Fundamental Particles and
Their Arrangement in Atoms

Dalton regarded atoms as ultimate particles of matter, discrete and


indivisible, but by the end of the nineteenth century this idea had been
abandoned. It is now believed that atoms are made up of certain
fundamental particles. In order to account for the properties of matter
which are of interest to a chemist, it is only necessary to consider three
fundamental particles — the electron, the proton, and the neutron.

The Electron
In the late nineteenth century a great deal of work was carried out on the
effects of electrical discharges through gases at very low pressures of the
order of 1 mmHg (133 N m~?) or less (see Figure 3.1). One of the results
of this work was the discovery of ‘cathode rays’. These rays were observed
as a faintly luminous beam which caused a fluorescence when it hit the
glass wall of the discharge tube.

cathode rays

cathode anode

Figure 3.1. Production of cathode rays in a discharge tube

Cathode rays were investigated and found to have the following properties:
1. They emerge from the cathode at right angles, travel in straight lines,
and cast a shadow of an object placed in their path.
2. They are deflected by an electrostatic field away from the negative
plate, proving that the rays carry a negative charge.
3. They are deflected by magnetic fields.
4. They can cause a small paddle wheel to rotate by exerting a mechanical
pressure.
From a consideration of these facts, it was concluded that cathode rays
consist of a stream of negatively charged particles, to which the name
electrons was given. In 1897, J. J. Thomson, by measuring the angles
n9) Physical Chemistry
through which the rays were deflected by known magnetic and electric
fields, succeeded in determining a value for the ratio of charge to mass
(e/m) for the electron. It was found that the ratio e/m was constant
irrespective of the nature of the cathode, the nature of the residual gas, or
the voltages employed. Thus it was considered that the electron is a
fundamental particle common to all atoms.
About 1910, Millikan determined the charge on the electron by observing
oil drops which were charged by the capture of gaseous positive ions.
These ions were produced by a beam of X-rays (see page 26) which removed
electrons from air molecules. By recording the movement of the charged
oil drops in a known electric field, he was able to establish that the smallest
charge on an oil drop was 1.6 x 10~*® coulomb and that all other charges
were simple multiples of this value. This charge, 1.6 x 10~7° coulomb, was
taken to be the charge on the electron, and together with the e/m value it
showed that the mass of an electron is about 1/1840 of the mass of a
hydrogen atom.
The electron, however, cannot be regarded as a particle without
qualification, as it can exhibit properties of a wave-like character; for
example, a stream of electrons can be diffracted by crystals, which act as
diffraction gratings (see page 64).

The Proton
As early as 1886 it was realized that as negative particles were observed
moving away from the cathode in a discharge tube, it was likely that
positive particles were moving towards the cathode. These were observed
and investigated by allowing them to pass through a perforated cathode.
The arrangement required to produce a beam of positive ‘rays’ is shown in
Figure 3.2.

positive rays cathode rays

anode
+

perforated
cathode —
Figure 3.2. Production of positive rays in a discharge tube

The e/m ratio for positive rays was determined by Thomson in 1912 and
it was found that the values obtained did depend on the residual gas. It was
concluded from this that the rays were positive ions formed when the
highly energetic electrons travelling from the cathode collided with the
molecules of gas left in the discharge tube. These collisions resulted in the
breakdown of some molecules into atoms and the removal of one or more
electrons to form positive ions.
Fundamental Particles and Their Arrangement in Atoms 23
The largest e/m value was obtained with hydrogen as the residual gas
which indicated that the hydrogen ion had the smallest mass. The masses
of ions from other gases were found to be approximately whole number
multiples of the mass of the hydrogen ion. Thus the hydrogen ion, formed
by the loss of one electron from a hydrogen atom, H — e~ — H+, was
thought to be a fundamental particle present in all atoms and it was called
a proton.

The Neutron
The existence of a neutral fundamental particle was predicted by Ruther-
ford in 1920, but it was not until 1932 that the neutron was discovered.
The experimental work which led to the discovery involved the use of
radioactive materials (see page 37). A radioactive substance contains atoms
which are unstable and emit radiation spontaneously. One type of radiation
consists of alpha-particles. Rutherford showed that alpha-particles
consist of helium ions, He?*, of relative mass 4.
In 1932 it was discovered that several of the lighter atoms, and in
particular the atoms of beryllium and boron, give out a very penetrating
radiation if they are subjected to the action of alpha-particles emitted from
polonium. Chadwick put forward the view that the radiation consisted of
fast-moving particles comparable in mass with hydrogen atoms and
possessing no electrical charge. This last factor accounts for their
penetrating power, as they are little affected by either positively or negative-
ly charged particles in the atoms which they encounter. The name neutron
was given to this type of particle.
The origin of the neutron in the above process appears to be that the
alpha-particles of relative mass 4 are captured by boron atoms of relative
mass 11. These are converted to nitrogen of relative mass 14 with emission
of a neutron of relative mass 1. Using modern nomenclature (see page 30)
this nuclear reaction may be represented by the equation
$He + 11B > 14N + Gn
The characteristics of the three fundamental particles are summarized in
the following table:

Approx. relative Relative charge


mass

Electron eS —1
1840
Proton 1 +1
Neutron 1 no charge
SS
ES
24 Physical Chemistry

The Distribution of Particles in the Atom


As early as 1906, Rutherford noticed that if alpha-particles pass first
through a very thin sheet of metal (about one-millionth of a centimetre
thick) and then on to a photographic plate, the effect produced on the
plate is not sharply defined but is diffuse at the edges. This was attributed
to ‘scattering’ of the alpha-particles by some action of the metallic atoms in
the sheet through which they had passed. Three years later Geiger and
Marsden, under the guidance of Rutherford, studied this scattering more
carefully and showed that while most of the alpha-particles were only
slightly affected by a sheet of gold of thickness 6 x 107° cm, about one
particle in eight thousand was scattered through an angle of 90° or more
from its original direction. Rutherford later commented that, considering
the high velocity and mass of the alpha-particle and the thickness of the
metal used, this result was ‘about as credible as if you fired a fifteen-inch
shell at a piece of tissue paper and it came back and hit you’.
In 1911, on the basis of Geiger and Marsden’s work, Rutherford put
forward his nuclear theory of the atom. He showed that it was extremely
improbable that the large deflections of alpha-particles could arise from a
succession of small deflections in the same direction. These large deflections
require the assumption of a region in the atom which fulfils the following
conditions:

1. It must contain a positive charge because the positive alpha-particles


are repelled by it.
2. It must be relatively massive and highly charged because the deflection
of the alpha-particles is so great.
3. It must occupy a very small space because most alpha-particles are
little affected and very few are repelled to any marked extent.

nuclei of metal atoms


+

a ce

alpha-particles 2 ei MGERLGAis

3 +

Figure 3.3 Scattering of alpha-particles by atomic nuclei


Fundamental Particles and Their Arrangement in Atoms 2
This is the postulate of the atomic nucleus, very small relative to the whole
bulk of the atom, massive and positively charged. The general idea is
illustrated in Figure 3.3. In the figure, alpha-particle 2 suffers slight scatter-
ing by distant effects of the nuclei. Particle 3 is one of the small minority
which suffer a large deflection by close encounter with a nucleus.
Rutherford was able to derive an equation relating the scattering to the
thickness of the metal, the charge on the nuclei of the metallic atoms, and
the velocity of the alpha-particles, but the war of 1914 intervened and it
was not until its close that the equation could be accurately tested.

The Wilson Cloud-chamber


In 1912, Wilson, using his cloud-chamber, obtained evidence which
supported Rutherford’s nuclear theory. The cloud-chamber consisted of a
cylindrical glass container in which was air saturated with water vapour
(or, better, with water vapour and alcohol vapour). The base of this
cylinder (in later models) consisted of a thin rubber diaphragm fixed at its
edges and kept taut by air pressure on the side away from the cylinder. By
automatic devices it was arranged that a valve could operate to reduce the
air pressure supporting the diaphragm so that rapid expansion occurred in
the cloud chamber while, immediately afterwards, alpha-particles were
allowed to enter the chamber. They ionized the gases through which they
passed and the ions acted as centres of condensation for drops of liquid.
By strong lateral illumination, two cameras (at right-angles to each other)
then photographed these drops, which appeared as a continuous trail on
the photographic plate. In this way, the tracks of alpha-particles could be
accurately catalogued and studied in three dimensions.
Wilson found that the great majority of such tracks were straight lines,
but a few showed forking into a longer and a shorter arm (see Figure 3.4).

Figure 3.4. Tracks of alpha-particles

This supported Rutherford’s nuclear theory. For the straight tracks


were those of the great majority of alpha-particles which passed through
the gases, ionizing atoms by striking out electrons, but encountering no
26 Physical Chemistry
nuclei because of the minute size of the nuclei. The small minority of
alpha-particles showing forked tracks underwent close encounter with a
nucleus and were deflected into a changed track (long arm) while the
ionized atom (or molecule) of the gas also showed an ionizing track
(short arm).

Moseley’s Experiments and Atomic Number


The work of von Laue at Zurich and the Braggs at Leeds had recently
shown that crystals could act as diffraction gratings for X-rays. Moseley,
working in Rutherford’s laboratory at Manchester University in 1913,
devised an apparatus in which cathode rays (i.e. electrons in very rapid
motion) were allowed to fall on an anti-cathode containing the element
under investigation. The element then produced its characteristic X-ray
spectrum, containing (generally) two strong lines. Moseley’s first set of
results was obtained for the elements calcium to copper in the Periodic
Table of Elements. He was able to show that the square roots of the
characteristic X-ray frequencies of these elements were proportional to the
atomic numbers of the elements, i.e. their ordinal number in the Periodic
Table. The effect is called the ‘Moseley stair-case’ (see Figure 3.5).

—s wavelength

Figure 3.5. X-ray spectra

In Moseley’s words, ‘we have here a proof that there is in the atom a
fundamental quantity that increases by regular steps as we pass from one
element to the next. This quantity can only be the charge on the atomic
nucleus.’ He went on to argue that, on average, atomic masses increase by
about two units from element to element in the Periodic Table. So, the
number of charges on the nucleus being about half the relative atomic
mass, it is very probable that in passing from one element to the next,
one proton is added to the nucleus. Moseley was able to assert on these
grounds that all possible elements up to aluminium were known and that,
Fundamental Particles and Their Arrangement in Atoms 21
between aluminium and gold, only four elements remained to be
discovered.
After the war Chadwick (1920) returned to the alpha-particle scattering
experiments and was able to show, with a probable error of between
1 and 2 per cent, that the nuclei of copper, silver, and platinum carried
positive charges of 29.3, 46.3, and 77.4 units respectively. The atomic
numbers of these elements in the Periodic Table are 29, 47, and 78, leaving
little doubt that the number of protons in the nucleus of these atoms is the
same as the ordinal number of the element in the Periodic Table. Similar
results were later obtained by various workers for aluminium, magnesium,
and gold.
This number-the ordinal number of the element in the Periodic
Table, the number of protons in the atomic nucleus and, since the atom is
electrically neutral, the number of electrons outside the nucleus — is called
the atomic number of the element and is usually denoted by the symbol Z.
For lighter elements, the atomic number of an element is about half its
relative atomic mass, as shown by the following table.

Relative atomic ;
Atomic number
mass

Oxygen 16 8
Sodium 23 11
Aluminium 27 13

It was to account for this difference that Rutherford in: 1920 predicted the
existence of a neutral fundamental particle. As previously described
(page 23) this particle, the neutron, was discovered in 1932 and was found
to have a relative mass almost equal to that of a proton. Thus, the nucleus
of an atom of relative atomic mass A and atomic number Z contains Z
protons and (A — Z) neutrons.
The general result so far may be summarized in the following way. All
atoms contain a nucleus which is very small, both absolutely and relative
to the size of the complete atom. The nucleus is made up of protons equal
in number to the ordinal number of the element in the Periodic Table, and
of neutrons (except the hydrogen nucleus which is a single proton). The
nucleus is positively charged by the protons present. For the lighter atoms
the numbers of protons and neutrons are about equal, but in the heavier
atoms the number of neutrons increases more rapidly than the number of
protons. For example, if the relative atomic mass of uranium is taken as
238, its atoms must contain 146 neutrons and 92 protons. A few typical
nuclei are given in the table on page 28.
28 Physical Chemistry

: Nucleus Relative atomic


Atomic number Protons Neutrons mass

Fluorine 9 9 10 19
Sodium 11 11 12 23
Aluminium 13 13 14 27

Atoms are neutral and so the number of electrons around the nucleus
must be equal to the number of protons in the nucleus. The more detailed
arrangement of the electrons around the nucleus will be discussed in
Chapter 5.
Since the proton and the neutron both have a relative mass very close to
unity, it would be expected that the relative atomic masses of all elements
should be integral (or very nearly so). This is in fact the case for a large
number of elements. The exceptions, like chlorine (relative atomic mass
35.5), arise from the fact that the naturally occurring element consists of a
mixture of atoms of different masses. Some chlorine atoms have a relative
atomic mass of 35 and some, 37. This phenomenon, called isotopy, which
arises from differing numbers of neutrons in the two nuclei, will be dis-
cussed in the next chapter.
4. The Nucleus | abies ae

Isotopy rds
Atomic nuclei are made up of protons and neutrons. The protons must be
accompanied by an equal number of extra-nuclear electronsto produce an
electrically neutral atom, but a neutron contributes no charge and requires
no such balancing by an electron. It is, therefore, theoretically possible for
two or more atoms to exist possessing the same number of protons on the
nucleus, the same number and arrangement of electrons outside the
nucleus, and, therefore, the same chemical properties (see Chapter 5), but
having different numbers of neutrons in the nucleus and therefore different
relative atomic mass.* Such cases are very common and the phenomenon
is known as isotopy. It may be formally defined as follows:
Isotopy is the occurrence of atoms with different relative atomic mass
but the same atomic number, and chemical properties which are identical
(or nearly so).
The idea of isotopy may be illustrated by the element chlorine, which
has two principal isotopes, ®°Cl and °"Cl.

Soy SGt

Nucleus: Protons 17 35 Protons 17 37


Neutrons 18 Neutrons 20
Atomic number Ai. Atomic number 17
Electrons 17/ Electrons 17

The identical atomic numbers and electron arrangements produce the


same chemical properties in the two isotopes, but the difference of two
neutrons in the nucleus produces a difference of two units of relative
atomic mass. The ordinary element, chlorine, has a relative atomic mass
of 35.45, showing that it contains about 75 per cent of ?$Cl and 25 per cent
of 37Cl by mass.

* Or mass numbers (see below).


30 Physical Chemistry
The numerical convention used in connection with isotopes is the
following. Consider the expression ¢X. X is the symbol representing one
atom of the isotopic element; a and b are figures written in front of the
symbol. The upper figure, a, is called the mass number of |the isotope and
is the sum of the number of protons and the number of neutrons in the
nucleus of the atom. The lower figure, b, is called the atomic number of the
element, X, and is the number of protons in the nucleus.
The relative atomic mass of the naturally occurring element is the
weighted mean of the mass numbers of the isotopes present, i.e. a mean
value which takes into account the relative abundance of each isotope in
the naturally occurring element.
Most elements exhibit isotopy. A few do not, having only a single type
of atom, notable examples being fluorine, sodium and aluminium. The
record number appears in tin, which has ten isotopes of mass numbers from
112 to 124 (atomic number 50). Many elements have more than six isotopic
forms. It is a notable fact that elements of odd atomic number never show
more than two isotopes, while elements of even atomic number show much
more extensive isotopy, e.g.

Element Atomic number Isotopes

Ag 47 Two (107, 109) .


Cd 48 Eight (106-116)
In 49 Two (113, 115)
Sn 50 Ten (112-124)
Sb 51 Two (121, 123)
Te 52 Eight (120-130)

Evidence for the existence of isotopes


1. Evidence from the mass spectrometer
In 1919 Aston developed his mass spectrograph with which he analysed the
positive rays from neon and found that there were two ions present with
relative masses of 20 and 22 respectively. The modern mass spectrometer
has been developed from Aston’s mass spectrograph and its essential
components are indicated in Figure 4.1.
The apparatus must be evacuated in order to avoid interference by the
air molecules. A vaporized sample of the element is introduced at A and is
subjected to a beam of electrons from the hot filament B. The electron
beam removes one or more electrons from the atoms, thus producing
positive ions. The positive ions are accelerated to the same velocity
through a slit by an electric field, C, and then deflected by a magnetic
The Nucleus 31

Figure 4.1. The essential components of a modern mass spectrometer

field, D. The angle of deflection of each ion depends on its ratio of charge
to mass. If all the ions have single positive charges, the angles of deflection
will be directly related to their mass numbers.
The strength of the magnetic field is varied so that each type of ion, in
turn, is detected by a collector, E. When the ions strike the collector they
produce an electric current, the magnitude of which is proportional to the
abundance of the isotope which gives rise to the ions. The detector
currents are recorded as a series of lines or peaks which is called a mass
spectrum. Each line or peak represents one isotope and the height of the
line or peak is proportional to the relative abundance of the isotope.

height
peak

80» 181.2+82 «837° 84.85, 86 87


mass number

Figure 4.2. The mass spectrum of krypton

The mass spectrum of naturally occurring krypton, Figure 4.2, provides


the information shown in the table on page 32.
32 Physical Chemistry
i

Percentage abundance, Relative mass of


Mass tune tse (taking total line length isotopes in 100
isotopes detected as 100 per cent) atoms of mixture

80 3.0 240
82 21:5 943
83 nD 945.5
84 57.0 4788
86 t720 1462

Total 8387.5

8387.5
The mean value per atom is therefore == $3.67

This weighted mean is the relative atomic mass of the naturally occurring
element which, by this method, can be determined to a very high degree of
accuracy.

2. Evidence of isotopy from diffusion


When experiments by Thomson, which were forerunners of the mass
spectrograph, showed the probability of isotopy in neon (?0Ne and ?2Ne)
Aston attempted their separation by diffusion experiments, based on
Graham’s Law. This states that:
The rates of diffusion of gases are inversely proportional to the square
roots oftheir densities (under comparable conditions).
The lighter neon isotope should, therefore, have a more rapid diffusion
rate than the heavier, in the proportion of 4/22:+/20. Aston was able to
separate neon (obtained from air) into two fractions differing in mass
number by more than 0.13 unit, which is greater than any possible
experimental error. Similar diffusion experiments on hydrogen chloride
(chief constituents H3°Cl and H87Cl) have produced similar results.
A similar diffusion technique is used to separate the fissionable isotope
of uranium, 733U, from isotope of mass number 238. The gaseous com-
pound UF¢ is employed in the diffusions, and after separation is heated
to generate the pure 235-isotope, which is used as a nuclear fuel.

3. Evidence from lead associated with radioactivity


Radioactivity is a phenomenon resulting from spontaneous changes in the
nuclei of the heaviest atoms (page 37). There are two main radioactive
The Nucleus 33
series, starting from uranium and thorium, respectively. Lead is the
end-product of each and is generated by changes briefly summarized as:

Uranium by toss of 3 alpha Radium which by lossofs Lead


———_—— SS
(238) particles gives (226) alpha-particles gives (206)

Thorium by loss of 6 alpha-particles gives Lead


(232) * (208)

Each alpha-particle (He**) emitted represents a loss of four units of


atomic mass so that, as the above figures indicate, lead associated in
nature with uranium minerals would be expected to have a mass number
of 206, while lead from thorium minerals would be expected to have a mass
number of 208. Actual results obtained experimentally are:
Uranium minerals Mass number of lead
Cleveite (Norway) 206.08
Pitchblende (W. Africa) 206.05

Thorium minerals
Thorite (Ceylon) 207.8
Thorite (Norway) 207.9
These samples of lead differ in mass number by almost the two required
units. They are indistinguishable chemically, having the same atomic
number, 82, but differ in that the heavier isotope has two extra neutrons
in its nucleus.
Such cases in nature are very rare. In a general way the relative atomic
masses of elements are constant, which shows that the isotopic mixtures
are constant in composition. Some differences, usually very slight, have
been reported, however. For example, atmospheric oxygen has been
shown to contain 3 per cent more of the isotope 180 than oxygen from the
waters of Lake Michigan.

Limitations of isotopy
It will perhaps have been noted that the variation of mass number in
isotopes of a given element is not very great, being rarely more than eight
units. The reason for this is that the addition of neutrons to a given nucleus
soon produces instability and this sets a limit to isotopy. Consequently, the
order of relative atomic masses for the elements is the same as the order of
atomic numbers in most cases. The important exceptions are Ar—K,
Co—Ni, Te—I, in which the order of relative atomic mass reverses the
order of atomic number. This similarity of order for most elements has
been important in the development of chemical classification because it
enabled Mendeléeff to devise the Periodic Table on the basis of relative
atomic mass though its true basis is atomic number.
34 Physical Chemistry
Isotopy of hydrogen
Hydrogen is known to exist in three isotopic forms, though only one of
them occurs in considerable quantity in ordinary hydrogen gas. The main
constituent of hydrogen is made up of atoms containing one proton as
nucleus and one electron. This isotope is denoted by the symbol +H, and
is called protium.
Ordinary hydrogen also contains a small proportion (about 1 in 4500) of
an isotope having, as nucleus, one proton and one neutron, with one
electron outside the nucleus. This isotope’s atom has about twice the mass
of the lighter atom. The chemical properties of the heavier isotope resemble
those of +H (though its reaction rates are different), because of the similar
electron structure. The effect of the added neutron is, however, relatively
great in such simple atoms and the compounds of the heavy isotope,
notably the oxide, differ appreciably from the compounds of protium. For
this reason, it has been considered desirable to give the heavy isotope
an individual name, deuterium, and a separate symbol, D.
The third isotope, tritium, symbol T, possesses a nucleus consisting of
one proton and two neutrons and has one electron outside this nucleus. It
can be produced by interaction between neutrons of low energy and the
isotope of lithium, §Li. The products are tritium and helium.
These three isotopes of hydrogen are represented diagrammatically as
follows:
x x . x

O oe og
iH
1
7D Tk:
© = nuclear proton
@ = nuclear neutron
x = electron

Production of deuterium oxide, ‘heavy water’, D,O, and deuterium, Dz


During electrolysis of ordinary water, which contains about 0.02 per cent
of deuterium oxide, hydrogen and deuterium may both discharge at the
cathode.
2H *(aq) + 2e~ > H,(g) 2D* (aq) + 2e- > D.(g)
There is, however, a preferential discharge of the light isotope, with the
result that deuterium oxide concentrates in the residue. This circumstance
is utilized in the following way. A 0.5 M solution of potassium hydroxide
in water is electrolysed using nickel electrodes till the concentration of the
alkali reaches about 5 M. The potassium hydroxide is then neutralized by
passage of carbon dioxide and the water is distilled off. The processes are
repeated on the distillate and, after seven such stages, reasonably pure
deuterium oxide is obtained. The gas evolved in the later stages contains
The Nucleus 35
increasing proportions of deuterium and is burnt with oxygen, the product
being returned to the earlier cells. Deuterium itself can be obtained at the
cathode by electrolysis of the deuterium oxide with potassium hydroxide as
electrolyte and nickel electrodes.
This electrolytic process for producing deuterium oxide has one great
disadvantage—the enormous consumption of electric power, which is in
the region of 40 000 ampere-hours per gramme of deuterium oxide.

Properties of deuterium, D,
In a general way, the chemical properties of deuterium resemble those of
hydrogen (+H) but deuterium usually reacts at a different rate. Its explosion
with oxygen is less violent. Quite a number of deutero-analogues of
ordinary hydrogen compounds have been prepared. By passing benzene
vapour and deuterium over nickel at 400 K the hydrogen atoms of benzene
are replaced by deuterium atoms.
C.H.(g) + 3D.(g) = CeDo(g) + 3H2(g)
Similar replacement in carboxylic acids has also been observed.
Deuterium, hydrogen and water also engage in the reversible reaction:
H.(g) + D,0(g) = H2O0(g) + D.(g)
Under similar conditions the reverse reaction is about three times as rapid
as the forward reaction, so that the deuterium concentration in the water
tends to rise.
By the reaction between deuterium oxide and calcium carbide, dideutero-
ethyne has been prepared and used in further syntheses.
CaC,(s) + 2D,0(1) > C2D.(g) + Ca(OD).(s)

Properties of deuterium oxide, ‘heavy water’, D,O


This liquid is colourless like ordinary water, but differs appreciably from
it as the following table shows.

Water Deuterium oxide

Boiling point 373.0 K 374.4 K


Freezing point 273.0 K 276.8 K
Maximum density 1.0 gcm~8 1.108 g cm~®

Deuterium oxide exhibits the same behaviour as water in showing a


maximum density at a certain temperature and lower densities both
above and below this temperature. For ordinary water the temperature of
maximum density is 277 K and for deuterium oxide 284 K.
36 Physical Chemistry
Separation of isotopes
Separation of isotopes is usually difficult because of the almost complete
identity of their chemical behaviour. The most successful separation
methods are physical.

1. Separation by positive-ray methods


This has already been discussed in principle on page 30. The mass
spectrograph separates isotopes and causes them to register the separate
lines of the mass spectrum. By actual collection of the particles, the
isotopes may be obtained separated. This was done on a minute scale for
the isotopes §Li and 3Li in 1934 and was later applied to the two uranium
isotopes, 725U and 738U.

2. Separation by diffusion
This was discussed earlier (page 32) in connection with the isotopes of
neon, of chlorine (combined in hydrogen chloride), and of uranium
(using UF.).

3. Evaporation methods
Lighter atoms tend to evaporate more readily than heavier atoms from
a mixture of the two. In similar conditions, the rate of evaporation is
inversely proportional to the square-root of thesmass of the atom. Using
this principle, Brénsted and Hevesy allowed mercury to evaporate at very
low pressure and at 320 K. The vapour was condensed on a glass surface
cooled by liquid air so that the mercury solidified. After a time the solid
distillate was removed, melted, and refractionated in a similar way.
Repetition of the process produced mercury differing by nearly 0.2 unit in
atomic mass from ordinary mercury. That is, a partial separation of
isotopes was accomplished.

4. Separation by thermal diffusion


Enskog (1911) showed that, if a mixture of gaseous molecules is enclosed in
a tube with its ends at different temperatures, heavier molecules tend to
concentrate at the cooler end. Clusius and Dickel applied this observation
to isotopes. They employed a vertical tube over 30 m long, with its sides
cooled by water and an electrically heated wire running down the middle.
In this way, convection currents were made to assist the thermal separation
and the isotopes ?7Cl and $7Cl were almost completely separated.

Mass defect, binding energy, and nuclear stability


The mass spectrometer can be used to determine the relative masses of
nuclei to such a high degree of accuracy that it has been possible to detect
The Nucleus 37
that the relative mass of a nucleus is always slightly less than the sum of
the relative masses of its constituents. This loss in mass represents the
binding energy of the nucleus (energy and mass being related by the
equation, E = mc’?, page 3).
The nature of the forces holding the nucleus together is not fully
understood, but an interesting observation is that only neutron-proton
ratios which lie within fairly narrow limits give rise to stable nuclei (see

130
120

of
number
neutrons

O 10 20 30 40 50 60 70 80 90 100
number of protons

Figure 4.3.

Figure 4.3). Just outside these limits, nuclei may exist, but they spontane-
ously disintegrate until stable nuclei are formed. This spontaneous
disintegration of unstable nuclei gives rise to radioactivity.

Radioactivity
Radioactivity was first noticed by Becquerel (1896) in uranium salts. Rays
emitted by these salts were found to affect a photographic plate in the
same way as light, to ionize gases, causing the discharge of a gold-leaf
electroscope, and to produce phosphorescence in certain materials, e.g.
zinc sulphide. Marie and Pierre Curie, working in Paris about 1900,
detected more intense radioactivity in a new element (named by them
polonium), and later they isolated another very powerfully radioactive
element, which they named radium.
38 Physical Chemistry
Kinds of radiation
The radiation emitted by uranium was found to be composite and of three
types, originally named alpha-, beta- and gamma-rays before the true
nature of each was recognised.
The alpha-rays were shown to undergo deflection in an electrostatic field
away from the positively charged plate and so must carry positive charge
themselves. They are now known to be helium ions, He?*, of relative
mass 4.
The beta-rays underwent a deflection opposite to that of alpha-particles
in the electric field and so must carry negative charge. By determination of
the ratio charge/mass, the particles were shown to be electrons. Beta-
radiation is simply an electron stream.
The gamma-radiation was unaffected by electrical fields and was found
to consist of electromagnetic waves of very high frequency, i.e. very short
wavelength (about 5 x 10774 cm).

Results of radioactive changes


The general principles of radioactive change are the following:
Alpha-particle emission. When a nucleus emits an alpha-particle, it
loses two protons and two neutrons, i.e. four units of mass. The lost
neutrons do not affect the chemical nature of the element. The two lost
protons reduce the atomic number of the element by 2 and cause the
product to take up a position two groups Jower in the Periodic Table.
Thus, ?28Ra in Group II loses an alpha-particle and produces 222Rn in
Group 0.
228@Ra— 222222Rn + $He
Beta-particle emission. Beta-disintegration can be considered as equivalent
to the splitting of a neutron in the nucleus into an electron, which is then
emitted as a beta-particle, and a proton, which the nucleus retains. The
electron is so light that loss of it leaves the relative atomic mass of the
residue virtually unchanged. The extra proton increases the atomic
number of the element by one unit and it takes up a position one group
higher in the Periodic Table. For example, the radioactive isotope of lead,
233Pb, in Group IV, loses a beta-particle (electron) to give a radioactive
isotope of bismuth, 249Bi, in Group V.
210 210;
a2Pb— *2gBi+ _{e
Radioactive changes involve the nucleus alone and are hence called
nuclear reactions. As can be seen above, nuclear reactions can be represent-
ed by equations. These equations, unlike equations for chemical reactions,
involve the formation of isotopes of different elements. The equations are
balanced by ensuring that the total relative masses (upper figures) on each
The Nucleus 39
side of the equation are equal and that the total relative charges (lower
figures) on each side of the equation are equal.

Decay series
When the product of a radioactive disintegration is itself unstable, then a
further radioactive disintegration will occur. Thus a whole series of
spontaneous radioactive disintegrations can occur (see the thorium
decay series below), each step involving the loss of either an alpha- or a
beta-particle or very occasionally both, until eventually a stable isotope is
formed.
Thorium decay series
40 Physical Chemistry
It is now recognized that there are several series of radioactive changes,
named from the elements in which they now originate in nature, e.g. the
actinium series, the uranium series (which contains radium), and the
thorium series. All these series show many and complicated changes,
ending in some isotope of lead (atomic number 82), e.g. the thorium series
produces 228Pb and the uranium series, ?88Pb. Another series, based upon
neptunium, has bismuth as its end-product.

Half-life
Radioactive change follows an exponential law, that is, the rate of decay of
a given material is directly proportional, at any instant, to the amount of
material present. This means that, kinetically, it is a first-order reaction
and it can be expressed by the formula (see page 258)
1 N 0
kt = ln—
W

where WN, is the original number of atoms of the isotope,


N is the number of atoms of the isotope left after time ¢,
kis the radioactive decay constant for the isotope.
The time for half the nuclei to disintegrate (i.e. when N = 4N,) is
represented by t, and is known as the half-life of the Botops. Substituting
in the above equation,
In
i, = ie

Hence the half-life of a particular isotope is constant and is independent of


the amount of starting material. It is therefore customary to characterize
radioactive isotopes by quoting their half-lives. For example, the half-life
of a particular isotope of radium is 1602 years, i.e. 1 g of radium held in
1978 will have decreased in amount to 0.5 g by about the year 3580. Half-
lives range from fractions of a second to millions of years.

Nuclear Energy
The energy changes associated with nuclear reactions are very much
larger than those associated with chemical reactions. For example, when
radium loses an alpha-particle to form radon, 4.2 x 10" joules are
evolved per mole of radium.
226Ra —> 222Rn + 4He

The possibility of utilizing nuclear energy changes in a humanly


controlled way has existed since Rutherford brought about the first
The Nucleus 41
artificial transmutation of one element into another. This was accom-
plished (1919) by the action of swift alpha-particles on nitrogen gas. In
this process the alpha-particle was captured by the nucleus of 14N (a gain
of 4 units of mass) with subsequent emission of a proton (loss of one unit
of mass) and the production of the isotope of oxygen, 130.
4He + 4N—> 120 + 1H
The chief process which has so far yielded energy from nuclear sources is
the process of nuclear fission. Its discovery in 1939 is usually associated
with the German nuclear scientist, Hahn. He found that the nucleus of the
uranium isotope, 285U, could absorb a neutron and then break up into two
roughly equal parts with mass numbers lying between 72 and 162. Neutrons
are also emitted. During this fission there is a loss in mass of about 0.2
units per mole of 22U. This mass is converted into energy according to the
equation E = mc?, where E is the energy in joules, m is the mass in kg, and
c is the velocity of light in m s~?.
E=02.x 107 °83.x7 10°)
= 1.8 #101? J mol"?
So the heat change associated with the nuclear fission of ?85U is of the order
of 2 x 10!°kJ mol-+. Relative to energy changes during ordinary chemical
reactions, such as the burning of carbon in which about 4 x 10? kJ
mol-? are evolved, this output of energy is stupendous.

Atomic bomb
This process of nuclear fission of 285U can be made into a “chain-reaction’
in suitable conditions. This is because, while fission occurs as a result of

Figure 4.4. Chain-reaction


42 Physical Chemistry
neutron absorption by the uranium nucleus, the process also emits
neutrons. These are available for absorption for new fissions, and so on.
If the mass of uranium is small, the neutrons may escape from it at such a
rate that the chain-reaction cannot be maintained but, above a certain
critical mass, the uranium absorbs the neutrons so as to produce very
rapid nuclear fission of at least a very considerable proportion of it. The
energy liberated then produces a terrific explosion, with a temperature of
several million degrees Kelvin for a short period after its occurrence.
The essential ideas relating to the production of the atomic bomb were
(a) the isolation of a sufficient amount of the isotope 788U from natural
uranium, (b) the invention of a device which would keep the uranium apart
in quantities below the critical mass until the decisive moment and then
shoot them rapidly together, and (c) the provision of a ‘tamping’ device to
reflect escaping neutrons back into the fissionable material. A source of
neutrons (such as beryllium activated by alpha-particles) is also required to
initiate the fission. The isotope of plutonium, ?39Pu, will produce a similar
chain-reaction by fission and was used in the second atomic bomb of 1945.

Hydrogen bomb
The nuclear reactions occurring in this bomb involve the fusion of hydrogen
isotopes into helium. These reactions are activated by the heat from a
smaller uranium or plutonium fission bomb. It is possible to construct
more powerful fusion bombs as the restriction of critical mass, which is
important in the case of a fission bomb, does not apply in this' case. Energy
is released in this case because a helium atom has a mass slightly less than
that of the hydrogen isotopes which combine to form it.
Nuclear fusions only occur at very high temperatures and hence they
are known as thermonuclear reactions. Stellar energy is due to thermo-
nuclear reactions which produce temperatures of several millions of
degrees. The sun is estimated to contain enough hydrogen to continue
radiating at its present rate for several thousand million years.

Nuclear reactors
The problem of adapting nuclear energy for useful purposes is that of
releasing the same energy as in the atomic bomb, but steadily over a
relatively long period and at a temperature convenient for making use of
the energy.
286U forms only 0.7 per cent of natural uranium and is difficult and
costly to purify. It is much more economical if natural uranium, containing
99.3 per cent of 288U, can be used. This isotope can absorb fast neutrons
but not slow neutrons. ?83U can absorb either of the two types and undergo
fission. Fast neutrons can be slowed down, sufficiently to affect 285U only,
by passage through graphite in which they collide with the carbon nuclei
The Nucleus 43
and lose energy. They are then rejected by 238U but are available for 22815,
fission, which will occur rapidly enough to maintain the chain-reaction.
Calder Hall, which came into operation in 1956, was the first nuclear
reactor to produce power on an industrial scale. In this type of reactor
uranium fuel rods in metal containers are arranged between blocks of
pure graphite. The reaction is controlled by lowering boron steel rods into
this atomic pile. The control rods absorb neutrons and hence slow down the
reaction. For safety, the pile is surrounded by several feet of concrete to
absorb stray radiation. A circulation of carbon dioxide at about 7
atmospheres pressure absorbs the liberated heat energy. The gas leaves the
reactor at a temperature in the region of 600 K. The heat is conveyed to
boilers which generate steam and operate alternators for the production
of electrical power.
A more recent development has been the reactor at Dounreay. The fuel
used in this case is 282Pu which will undergo fission when bombarded with
fast neutrons, hence it is not necessary to include the bulky graphite
moderator. Another advantage of the Dounreay-reattor is'that, by allowing
some of the neutrons from the plutonium fission to bombard 7338U, more
plutonium can be formed. Thus, by theseries of nuclear reactions given
below, this reactor produces its own fuel and is Known as a breeder reactor,
238 1 239TT _, 239 0
a2U + gn— 792U> 733Np + -{e
239 239 / 0
oa NP — 79¢Pu + _{e
Attempts are being made to control nuclear fusion which, if it proves
possible, will have the advantage that,the fuel, isotopes of hydrogen, is
more readily available than the heavy metals needed for the fission
reactors. ; v

Artificial Elements
The most complex atomic nucleus known on earth up to about 1940 was
that of uranium (atomic number 92). Since that time more complex
nuclei have been produced. The first two of these, neptunium and pluto-
nium, as indicated in the equations, have been built up from 7g5U.
By using a variety of projectiles (even as heavy as 1§O) and targets, this
series of elements has been built up as far as kurchatovium which has an
atomic number of 104.

‘Artificial’ production of radioactive isotopes


The development of methods for the transmutation of elements has also
proved useful in that it is now possible to prepare radioactive isotopes
which do not exist in nature. This means that when selecting a radioactive
isotope for a particular purpose in industry, medicine or some type of
research, there is now a much larger variety of isotopes to.choose from.
44 Physical Chemistry
For example, if sodium chloride is subjected to the action of an intense
concentration of neutrons in a uranium pile for a few days, a certain
proportion of the 2%Na will absorb one neutron per atom to form a
radioactive sodium isotope, 24Na. In similar conditions, sulphur, 98S, will
absorb one neutron per atom, emit a proton, and so finish as a radioactive
isotope of phosphorus, $2P.
38S + on — 3gP + ip

Uses of radioactive isotopes


The uses of radioactive isotopes may be categorized as follows:
1. those in which the effects of radiations are used,
2. those in which the fact that the radiations are easily detected are put to
use.

The first category may be subdivided into:

(a) Curative uses


Cancerous growths may sometimes be eradicated by exposure to gamma-
rays from radioactive materials. Radium has been used for this purpose
but was scarce and expensive. More recently radioactive cobalt has been
used and is a much cheaper and more easily available material. It can be
made by irradiating ordinary cobalt, containing 32Co, with neutrons to
produce 89Co. This radioactive cobalt is placed near the centre of a lead
sphere Hae absorbs unwanted radiation. The wanted radiation is
allowed to escape down a very narrow hole in the lead and directed at the
tissue to be treated.
Radioactive phosphorus has been used in the treatment of leukaemia
and radioactive iodine in disease of the thyroid gland.

(b) Sterilization of medical equipment


Radiation is also used for the sterilization of such items as disposable
hypodermic syringes. The syringes are sterilized after they have been
packed and they are not unpacked until immediately before they are to be
used.

The second category may be subdivided into:

(i) Tracer techniques


The radioactive isotope of an element does not differ chemically from the
‘ordinary’ atoms, therefore the radioisotope may be used as a marker,
tracing what occurs in some process under review. A biological example is
The Nucleus 45
the tracing of the course of photosynthesis by using carbon dioxide
containing the radioactive ‘$C instead of the ordinary 12C.
In organic chemistry the labelling of a particular atom in a compound,
by preparing the compound with a radioactive isotope in that position, can
be used to trace the reaction path and hence help to elucidate the mechan-
ism of the reaction. An example in inorganic chemistry is the determination
of the solubility of a sparingly soluble salt. A saturated solution is prepared
using the salt containing a known proportion of radioactive isotope. The
concentration of this solution is determined from the intensity of the
radiation. The analysis of such a dilute solution would not be possible by
more conventional means.
In industry the tracer technique can be used to detect the leaks in water
mains; or, when the grade of oil is changed in a pipeline, radioactive
material is added to the first of the new flow and is readily detected at the
receiving end.

(ii) Absorption of radiation


The thickness of a sheet of metal can be continuously checked bymeasuring
the absorption of radiation which is directed through it. A similar technique
can be used to detect that containers are full of a particular product.

(iii) Dating
Carbon-14 dating depends upon the fact that the percentage of #C in the
atmosphere, and hence in all living things, is almost constant as it is
continuously regenerated by the action of cosmic radiation on nitrogen.
From the moment death occurs the 14C will not be replenished from the
atmosphere and the 14C which is present will decay at a definite rate. The
half-life of this isotope is 5600 years, and by measuring the proportion of
14C remaining it is possible to put an approximate date on the time of
death.

(iv) Radiography
Faults in welded metal plates and pipes can be detected by a radiography
examination. The advantage of this method is that the source of radiation
can be very small compared to the equipment required to produce the
more conventional X-rays.
5. Electrons in the Atom

The atom is an electrically neutral structure. It is, therefore, obvious that


it must contain electrons equal in number to the protons in the nucleus.
The question is, how are the electrons distributed? An indication of the
arrangement of the electrons around the nucleus is provided by atomic
emission spectra.

Atomic Emission Spectra


When atoms of an element are heated or subjected to an electrical dis-
charge, it is observed that the element emits electromagnetic radiation. This
is caused by the atoms of the element absorbing some energy and then
emitting it in the form of electromagnetic radiation. When this radiation is
passed through a spectrometer, it is observed that, unlike sunlight, it is not
a continuous spectrum. The spectrum obtained consists ofa number of
definite lines where each line corresponds to a definite wavelength of
radiation. In the visible part of the spectrum this is observed as a number
of lines, each of different colour.
The fact that it is a line spectrum and not continuous indicates that only
certain energy absorptions and emissions are possible. This can only be
accounted for by the existence of definite electron energy levels within an
atom, each absorption being due to the movement (jumping) of an
electron from one energy level to a higher one. When the electron reverts |
to a lower energy level, energy is emitted in the form of electromagnetic

higher
er ENS Se Se ee energy level

input emission of radiation


of aie
normal
SANG SS Sa Seen eee ——@—— energy level
Figure 5.1. Diagrammatic representation of absorption and emission of energy by
an electron
Electrons in the Atom 47
radiation and this radiation will have a definite wavelength corresponding
to the definite energy drop. The discrete amounts of energy which can be
absorbed or emitted by an atom are called quanta (see Figure 5.1).
The energy and frequency of the radiation are related by Planck’s
equation:
E=h

where £ is the energy, » is the frequency, and / is Planck’s constant.

Hydrogen spectrum
Each element has a characteristic pattern of lines in its atomic emission
spectrum. Hydrogen, having only one electron per atom, is the simplest
example to consider. Its spectrum consists of several series of lines all of
which can be accounted for by postulating the existence of energy levels as
indicated in Figure 5.2. The possible energy levels are denoted by in-
creasing values of n. The lowest energy level, or ground state, is described
as the m = 1 state.

n=co
n=T7
n=6
n=5
n=4

n=3

n=2

energy levels

n=1

hydrogen
spectrum

Paschen series Balmer series Lyman series

Figure 5.2. The energy ‘drops’ which produce part of the hydrogen atomic emission
spectrum
48 Physical Chemistry
Each series is produced by the reverting of electrons to a particular
energy level from all the energy levels above it. Thus the Lyman series is
due to electrons reverting back to the lowest energy level, n = 1. The
Balmer series, which occurs in the visible part of the spectrum, is due to
electrons reverting to the n = 2 state; the Paschen series to the n = 3
state, and so on.

Convergence frequency and ionization energy


The spaces between the successive lines in each series gradually become
smaller as the frequency increases, indicating that the higher energy levels
in the atom gradually become closer together until eventually they merge.
If an electron can be made to jump beyond this highest energy level, it will
not return and the atom will have ionized. The energy required to remove
an electron from its ground state (n = 1) to beyond the highest energy
level is the ionization energy for that electron in that atom. Therefore, in
the case of hydrogen, the convergence frequency for the Lyman series
must correspond to the ionization energy of hydrogen.
The convergence frequency of the Lyman series is of the order of
3.3 x 1015 Hz. Taking Planck’s constant as 4 x 10713 kJ s mol-1, and
using the equation E = hr, the ionization energy is given by
EAC 1OT Kes Sen”
= 13x LU) Ky OL
Ionization energies, derived from spectroscopic data, are very important
as their interpretation in terms of electron energy levels enable chemists to
rationalize a lot of the chemical behaviour of elements.

Distribution of Energy Levels


Atomic emission spectra indicate that there are only certain energy levels
available in an atom. Ionization energies provide an indication of how
these energy levels are distributed.
Consider first, successive ionization energies, i.e. the energy required to
remove the first electron, then the second electron, then the third electron,
and so on, from atoms of a particular element. A graph of the logarithm of
successive ionization energies of potassium against the number of electrons
removed (Figure 5.3) shows that, as expected, each electron is more
difficult to remove than the previous one. However, in addition to this
general trend, there are some distinct breaks in the graph, indicating that
the energy levels are arranged in groups. These start with 2 electrons, with
fairly similar ionization energies, which are both near the nucleus and
difficult to remove, followed by 8 electrons with fairly similar ionization
energies (but distinctly less than the previous 2), followed by another group
of 8 and finally by a single electron which, compared to all the other
Electrons in the Atom 49

>
w
a

N
mol’)
log
ionization’
(kJ
energy

1 i

OR 2S 4 567 Se NO Film i20 138 145015416 17 18) 19


number of electrons removed
Figure 5.3 Log of successive ionization energies of potassium

electrons, is very easily removed. These groups of energy levels are referred
to as electron energy shells. Each shell is denoted by a particular value of
n (principal quantum number) or by a letter. For example, the nineteen
potassium electrons are arranged as follows:
K L M N
Shell n= h=2 n=3 n
Number
of 2 8 8 1
electrons

The electrons of the first twenty elements arranged in energy shells are
_ given on page 50.
A point which will be taken up later (Chapter 6) is that elements which
are chemically similar are found to have the same number of electrons in
their outer shells.
A more detailed insight into the distribution of energy levels is obtained
by plotting a graph of the first ionization energies of the elements against
their atomic numbers (see Figure 5.4, on page 51).
For the first twenty elements the general shape of the graph follows the
2, 8, 8, 2 pattern of the principal electron energy shells. In addition the
8 electrons of the nm = 2 shell and the 8 electrons of the n = 3 shell appear
to be subdivided into 2, 3, 3 arrangements. A more advanced analysis of
spectra shows that there are only two energy levels, one holding 2 electrons
50 Physical Chemistry

Elements of the first twenty elements arranged in energy shells


1 2 | 3 |4

1H 1
2 He 7)

3 Li 2 1
4 Be #) 2
5B 2 3
6G 2 4
7N a 5
80 2 6
or 2 7
10 Ne 2 8

11 Na 2 8 1
12 Mg 2 8 2
13 Al 2 8 3
14 Si 2 8 4
15P Z 8 5
16S 2 8 6
17 Cl 2 8 7
18 Ar f. 8 8 :

19K 2 8 8 1
20 Ca 2 8 8 2

and the other 6, and the extra break in the graph occurs at the point where
the 6 level is half full, which is a particularly stable arrangement. In order
to account for the graph beyond the first twenty elements it is necessary to
include subshells, one type holding 10 electrons (e.g. atomic numbers
21-30) and one holding 14 electrons (e.g. atomic numbers 58~71). These
subsidiary energy shells are denoted by the azimuthal (or subsidiary)
quantum number, /, which can have values of 0, 1, 2 etc. up to a maximum
ofn — 1. They are also known by the letters s, p, d, and f which are the initial
letters of the names given to certain spectral lines.

Subshell : a 4 .

Maximum
number of 2 6 10 14
electrons
2500 He

')
mol”
(kJ
ionization
energy

0 10 20 30 40 50 60
atomic number

Figure 5.4. Variation of first ionization energies with atomic number

energy

As

3d

0 20 40 60 80
atomic number
Figure 5.5. Variation in energy of the 4s and 3d levels with atomic number
52 Physical Chemistry
It is therefore possible to characterize an electron by its principal quantum
energy shell and its subsidiary quantum energy shell. Thus an electron in
the s subshell of the third shell is known as a 3s electron.
The first 2 electrons of all elements are in the 1s energy level, the next 2
in the 2s, the next 6 in the 2p and so on. The 2s electrons of, say, sodium
will be in a different environment than the 2s electrons of, say, uranium;
it might therefore be expected that the absolute energy levels of these
electrons will be different. It is of particular interest to consider the
variation in energy of the 4s and 3d electrons with atomic number (Figure
5)
It can be seen from Figure 5.5 that the 4s is at a lower energy level than
the 3d in the atomic number region 20. The electron shells are filled in
order of increasing energy and therefore the two outer electrons of calcium
are in the 4s shell and not the 3d shell. The next ten elements after calcium
have their extra electrons added to the 3d shell, thus giving rise to the
d-block or transition elements.
The outer electron configurations of chromium, 3d® 4s, and copper,
3d?° 4s!, are anomalous and indicate that there is a special stability
associated with the half filled and completely filled d subshell.
The relative energies of the subshells for the first thirty-six elements are
given in Figure 5.6.

4p
3d Te :

3p
3s

: 2p
2s

1s

Figure 5.6.

The subshells are filled in order of increasing energy according to the


Aufbau Principle.
Figure 5.6 can be used to determine the electron arrangement in an
atom, a superscript being used to denote the number of electrons in a
particular subshell. For example, chlorine has an atomic number of 17 and
therefore 17 electrons around its nucleus. These electrons will be arranged
as follows:
1s? 25% 2p%6s%3p5
Electrons in the Atom 53
Iron has an atomic number of 26 and its electron arrangement is
1s? 2s? 2p® 3s? 3p® 3d® 45?
An alternative way of remembering the order in which the subshells are
filled is provided by Figure 5.7.

(2) (6) (10) (14)

ies
1 Es"

at
eee
ae
ihwag
Cs as
bok
Figure 5.7.

Electron spin
On closer examination, some apparently single spectral lines are found to
be double lines. These doublets are accounted for by electrons having spin
and there being two directions of spin. Two electrons with opposite spin
can form a stable pair of electrons. Thus the 2 electrons in a 1s subshell
can berepresented by , the 6 electronsina2psubshell willexistas three

pairs and can be represented by , the 10 electrons in a 3d sub-

shell will exist as five pairs and can be represented by

and the 14 electrons in a 4f subshell will exist as seven pairs and can be

represented by | NIN INT AINA The electronic configur-

ation of, for example, fluorine may be written as

rN) [M] [NINT


54 Physical Chemistry

Within a particular subshell electronic configurations are always written


with the maximum number of unpaired electrons (Hund’s Rule). For
example, oxygen (atomic number 8) is written as

1s 2s 2p

and not as

1s 2s 2p
ENT |
Quantum Numbers
In describing the energy state of a particular electron two quantum num-
bers have so far been used:

1. The principal quantum number, n, where 7 = 1, 2, 3 etc. and denotes the


principal quantum energy shell in which the electron is found.

2. The azimuthal or subsidiary quantum number, /, which can have values


of /=0,/=1,]=2 up to] =n — 1. This means that when x = 3,
I can have values of 0, 1, and 2. That.is, there are three’ types of sub-
shell (s, p, and d) in the third principal shell.

In order to describe completely the energy state of an electron two


additional quantum numbers are required.

3. The magnetic quantum number, m, which can have values from + / to


— |. Thus, when / = 2 (d subshell), m can have five values, + 2, + 1,
0, — 1, — 2. The use of this quantum number stems from work on the
effect of magnetic fields on spectral lines which gave an indication of
the number of energy levels within a particular subshell. Thus when
1 = 2, m can have five values; therefore there are five energy levels in
this subshell.

4, The spin quantum number, s, which can have values of + 4 or — } and


describes the fact that each energy level in a subshell can hold two
electrons of opposite spin. This means that for the / = 2 subshell there
are five energy levels (five values of m) each of which can hold 2
electrons giving a total of 10 electrons for a d subshell.

The quantum numbers of the electrons in the first three principal energy
shells are as follows.
Electrons in the Atom 55

=] = “

First shell

Second shell
:
)
2p subshell

3s subshell

Kan)
3p subshell
y

Third shell

3d subshell

WD
S
WW
WW KBP
eK
KH
°O
COCO
HSH
aH
SH
NNNNNNNNNN
PH
P
Koo
aa
Sa
PPINNNNNNDN
WWW ae
|
a
eee
Leer
Pee?
t+
1b NJ=—
N=
N[—
NJ—
NI=—
Nl=—
NJ=
N/—
N|—
v=
y=
]=
|=
=
=
|—
|—
NV]

It can be seen from this that no two electrons in an atom can have all
four quantum numbers with identical values (Pauli Exclusion Principle).
For example, the two electrons in the 3s subshell have quantum numbers

= 3)7=0,m
= 0,5 = 4-4
n=3,1=0.m=0,s=—}

Atomic Orbitals
In the early days (Bohr 1914), the definite energy levels indicated by
atomic spectra were visualized as definite electron orbits. At a later stage
it was realized that, due to its extremely small mass, it would not be
56 Physical Chemistry
possible to determine experimentally the exact path and velocity of an
electron in an atom (Heisenberg 1927). However, by considering the wave
properties of an electron, it has proved possible to calculate the probability
of finding an electron in a particular position.
If the probability of finding the 1s electron of hydrogen at a particular
distance from the nucleus is plotted as a radial charge density against the

charge
radial
density

distance from nucleus

Figure 5.8.

distance from the nucleus (Figure 5.8), it can be seen that the most likely
place to find the electron will be within a spherical shape fairly close to the
nucleus. This shape can be thought of as the volume within which there is,
say, a 99 per cent chance of finding the electron. The volume is called an
atomic orbital.

Py P, Pp,

Figure 5.9. Two-dimensional representations of the shapes of s, Pz» Pys and p.. orbitals
Electrons in the Atom 57
The s atomic orbitals are spherical and hold a maximum of 2 electrons.
The 6 electrons of a p subshell are found to be in three orbitals, each
holding 2 electrons. The p orbitals, which resemble dumb-bells in shape,
are on axes at 90° to each other and are referred to as the p,, p,, and p,
orbitals (Figure 5.9).
6. The Periodic Classification of the
Elements

This very important scheme of classification was first put forward by the
Russian chemist, Mendeléeff, in 1869. It had, however, certain partial
forerunners in the form of Ddébereiner’s Triads and Newlands’ Law of
Octaves.

Doébereiner’s Triads
In 1829, Doébereiner drew attention to the existence of sets of three
chemical elements (hence triads), the members of which resemble each
other closely in chemical properties and have one of two relations among
their relative atomic masses — either the relative atomic masses of all three
are similar, or the middle one is close to the arithmetic eae of the other
two. Actual cases are:

Cl 35.5 Ca 40.1 Fe 55.85 Relative


Br 80 mean—81.25 Sr 87.6 mean-— 88.75 Co 58.94 atomic masses
I 127 Ba 137.4 Ni 58.69 similar

These triads are actually selections from much broader similarities, but
are of historical interest because they represent the earliest known
attempt to trace an interdependence between the properties of elements
and their atomic masses.

Newlands’ Law of Octaves


This idea was put forward in 1867. Newlands developed the idea that, if
‘the elements are written in order of increasing relative atomic masses (as
'below), a Law of Octaves can be discerned. By the operation of this law,
the first, eighth, and fifteenth elements show chemical similarity, as do the
second, ninth, and sixteenth, and the third, tenth, and seventeenth, and so
on. Written as in the table below (which includes the known elements of
Newlands’ time), the law should place chemically similar elements in the
The Periodic Classification of the Elements 59
same vertical columns. Notice that no noble gases appear; they remained
unknown till the 1890s.

Hise Be Been NO
FSi Ne PMpecsaAl) Si PB -=S
Chey ‘KeerCa Cre Ti? *Mn “Fe

Similarities such as those of F and Cl; Li, Na, and K; C and Si; N and PR
are obvious, but the scheme soon broke down, e.g. Fe is not suitable for
inclusion with O and S.
Here Newlands had the first glimmerings of the idea that has produced
the Periodic Classification, but he did not develop it adequately. In
particular, he failed to allow for the imperfect chemical knowledge of the
mid-nineteenth century. It is enough to point out that the third horizontal
period should have read:

Cl K Ca (Sc) Ti (V) (Cr) Mn Fe

The bracketed elements were unknown or misplaced in the Newlands


table.

The Periodic Classification of the Elements (Mendeléeff, 1869)


This classification was based on the Periodic Law which, as given by
Mendeléeff, stated that the properties of the elements are periodic functions
of their relative atomic masses. That is, if the elements are arranged in
order of increasing atomic mass, elements with similar properties occur at
regular intervals. Instead of arranging the elements in the form of a
continuous list, Mendeléeff placed the elements in horizontal rows
(periods) so that elements with similar properties appeared in the same
vertical column (group). The noble gases and a few other elements, such as
gallium and germanium, were not discovered until later and were therefore
absent from the early Periodic Table.
_ On the basis of relative atomic mass, two anomalies appeared in the
early table. Cobalt (58.94) and nickel (58.69), and tellurium (127.6) and
iodine (126.9), had to have their relative atomic mass order reversed to
bring them into correct placing on chemical grounds. The strict order of
relative atomic mass would have separated iodine from the other halogens
into Group VI, while tellurium would have been separated from the
closely-related elements sulphur and selenium, and placed in Group VII
with the halogens. Another anomaly arose in the 1890s when the newly
discovered argon (39.9) had to be put before potassium (39.1) to keep
argon with the other noble gases and potassium with the alkali metals.
These cases showed clearly that relative atomic mass is not really the true
basis of the table.
The Periodic Table
(Mendeléeff’s arrangement)

Se
34 35

ie Ru Rh Pd
: 43 44 45 46
.

|
53
127.6 126.9

W Re Oss lraene
74 75 MO 79h TAS)
i Po At
84 85
Th U
90 92

The Ce | Pr Sm Gd| Tb} Dy| Ho} Er |Tm| Yb] Lu


Rare Earths|58 | 59 62 64 |65] 66 | 67 |68 | 691 70 | 71

Transuranium| Np} Pu | Am| Cm) Bk | Cf Fm No


Elements |93 | 94] 95} 96] 97 | 98 100 102

Atomic masses are quoted where their order reverses the order of
atomic number.
The Periodic Classification of the Elements 61
The anomalies have disappeared with the recent recognition of atomic
number instead of relative atomic mass as the true basis of the Periodic
Table. The atomic number of an element, being the number of protons on
the atomic nucleus, determines the number of electrons in the atom,
hence their arrangement and hence the properties of the element. The
Periodic Law is now restated to read:
The properties of the elements are periodic functions of their atomic
numbers.
The three anomalies mentioned on page 59 on the basis of relative
atomic mass arose out of isotopy. The case of argon and potassium will
illustrate the position. The principal isotopes of these elements are:

Argon Potassium

Nucl Protons 18 18 Niged Protons 19 19


ucieUus) Neutrons 18 22: ucleUS) Neutrons 20 22
Ar = 36 Ar = 40 K = 39 K = 41

Atomic number 18 for both Atomic number 19 for both


Electrons 2, 8, 8 for both Electrons 2, 8, 8, 1 for both

In the case of argon, the heavier isotope predominates, giving an average


atomic mass of 39.9; with potassium, the lighter isotope predominates,
giving an average atomic mass of 39.1. But the order of atomic numbers
clearly puts argon before potassium and there is no anomaly on this basis.
The explanation is similar in the case of cobalt (no isotopy, mass number
59) and nickel (five isotopes, 58-64) ;and tellurium (eight isotopes, 120-130)
and iodine (no isotopy, mass number 127).
Isotopes of a given element do not, in general, exceed a range of eight
units of atomic mass. This range keeps the atomic mass order the same as
the atomic number order, except for the three cases mentioned. This
enabled Mendeléeff to recognize the Periodic Law and Table on the
relative atomic mass basis though its real basis is atomic number.

Lothar Meyer’s Atomic Volume Curve


The work of Mendeléeff was mainly concerned with periodicity of chemical
properties, whereas, during the same year, Lothar Meyer noted a
periodicity in a physical property, namely atomic volume. The atomic
volume of an element is given by the relationship
mass of | mole of atoms
atomic volume = ;
density
62 Physical Chemistry

60

SiS

Ww(S)

(cm*)
volume
atomic

10

(¢) 10 20 30 40 50
atomic number

Figure 6.1. Variation of atomic volume with atomic number

The graph of atomic volume against atomic number (Figure 6.1) or, as in
the case of Lothar Meyer, relative atomic mass, consists of a series of
maxima and minima showing that the atomic volumes of elements are
periodic functions of their atomic numbers. There is a close resemblance
between this periodicity and that noted by Mendeléeff. For example, the
elements which appear at the maxima of the curve all occur in the same
group of Mendeléeff’s Periodic Table.

Periodic Table and Electron Arrangements


With the development of theories of atomic structure and, in particular,
the energy levels occupied by electrons, it became apparent that periodicity
in physical and chemical properties is related to the electronic configur-
ations of the elements. Considering the first twenty elements and their
principal quantum energy shells, it can be seen that elements in the same
group have similar outer electron arrangements:
The Periodic Classification of the Elements 63
H He
1 2
Li Be B (c N O F Ne
2,1 2,2 2,3 2,4 2,5 2,6 2,7 2,8

Na Mg Al Sie aad Ss Cl Ar
2,8,1 2,8,2 2,8,3 2,8,4 2,8,5 2,8,6 2,8,7 2,8,8

K Ca
2,8,8,1 2,8,8,2

When the elements beyond the first twenty are considered and the
distribution of electrons into subshells is taken into account, it becomes
desirable to modify the layout of the Periodic Table. The elements are
better arranged in what is known as the Jong form of the Periodic Table
(Figure 6.2).

LH|He|
[Alsi]PS [cllAr]
Ti |V|Cr)Mr{ Fe|Co|Ni|CulZn|GalGe/As|Sel
Br]Kr
Zr |NbMaTc|Ru|Rh|Pd|AgiCd| In|Sn|SbiTe| |[Xel
|Eu|Gd{Tb|Dy|Ho! Er Hf|Ta]W|Re|Os| Ir |Pt /Au|Hg| TI[Pb]Bi]Po|At
|Rn|
reall eo vereat

s, p, d, and f blocks of elements

Figure 6.2. s, p, d, and f blocks of elements in the long form of the Periodic Table

On examination of the electronic structures of the elements, it can be


seen that the elements in Period 4 (atomic numbers 19-36) consist of two
elements, K and Ca, with their last electrons added to an s subshell, ten
elements, Sc to Zn, with their last electrons added to a d subshell, and six
elements, Ga to Kr, with their last electrons added to a p subshell. Hence
this part of the Periodic Table can be divided into s, d and p blocks of
elements. Starting at element 58 it is also necessary to include an f block
of elements.
A more detailed discussion of the variation in properties of elements and
compounds in relation to the Periodic Table is given in Inorganic Chemistry,
the companion volume to this book.
7. Bonding I: Ionic and Covalent

All pure solid substances consist of definite arrangements of particles.


These particles may be atoms, charged atoms, or groups of atoms (called
ions), or collections of atoms (called molecules). The physical properties of
a substance give an indication of the type of particle it contains and the
nature of the bonding forces involved.

1. Conduction If a compound, when it is converted to its liquid form,


conducts electricity then the compound contains ions. Whereas, if the
liquid form of a compound does not conduct it probably exists as
molecules.

2. Melting point A high melting point indicates that the forces being
overcome during melting are strong. This usually indicates that the solid
exists in the form of a giant lattice. This is a large continuous arrangement
of ions or atoms which can only be broken down (i.e. melted) by over-
coming the main bonding forces between the atoms or ions.
A low melting point usually indicates that the substance exists as a
molecular lattice. This consists of molecules held in a particular arrange-
ment by weak attractive forces. When the substance melts, the weak
attractive forces between the molecules are overcome but the molecules
themselves remain intact.

3. X-ray diffraction When light is passed through a series of very fine


slits (diffraction grating) each slit acts as a secondary wave source. This
diffraction of the light only occurs when the width of the slits is of the
same order as the wavelength of the light. The wave fronts from these
secondary sources will meet and interference will occur. Where the waves
are in phase they will reinforce each other and produce a brighter light, but
where they are out of phase they will tend to cancel each other out and
leave a dark area. Thus, on viewing the light through the grating, a series
of maxima and minima, called a diffraction pattern, is observed. In a
similar manner, when X-rays, which have a much smaller wavelength than
visible light, are reflected from the electron clouds of successive layers of a
crystal, reinforcement will occur when the reflected rays are in phase. The
angles at which the reinforcements occur will depend on the dimensions of
the crystal. The relationship between this angle, the wavelength of the
Bonding I: Ionic and Covalent 65
X-rays, and the distance between successive layers in the crystal was
originally determined by Bragg in 1912.

ray1

Figure 7.1.

In Figure 7.1. AB and CG represent successive layers in a crystal and d


is the distance between the layers. A parallel beam of X-rays of wavelength
A is incident at an angle 0. Ray 2 travels a distance ED + DF further than
ray 1. If this distance is equal to a whole number of wavelengths, the
reflected rays will be in phase and reinforcement will occur.
ED = d sin 0
DF = dsin 0
ED+ DF = 2d sin 0
For reinforcement nda = 2d sin 8

where v is a whole number.


It is therefore possible to use the results of X-ray diffraction to calculate
a value for d and, if the crystal is investigated from various angles, the
dimensions of the crystal lattice can be determined. This technique is
called X-ray crystallography and a more detailed discussion of structures
which have been elucidated by this method is given on page 86.
The wavelength of electrons (page 22) is such that electron diffraction
patterns may be obtained in much the same way as X-ray diffractiv:
patterns. Electron diffraction is of particular value in organic chemistry as
it reveals the positions of hydrogen atoms more readily than the X-ray
method.
In addition, X-ray and electron diffraction provide useful information
concerning the way in which the atoms are bonded together. The X-rays
are reflected by the electrons and the intensities of the reflections are
related to the electron densities around the nuclei. The distribution of
electrons can be represented by an electron density map which is drawn so
that it is consistent with the results of the X-ray diffraction. Such a map
consists of a series of contours where each contour joins all the points with
66 Physical Chemistry
identical electron densities. The electron density map of a substance
containing ions is very different from a molecular substance (see Figure
7.2).

(b)
Figure 7.2. Electron density maps of (a) an ionic compound, sodium chloride, and
(b) a molecule, hydrogen

The electron density map of the ionic compound indicates that the
substance consists of an arrangement of almost spherical ions with very
low electron density between them, suggesting that the bonding forces
which hold the ions together are due entirely to the electrostatic attraction
between the oppositely charged ions.
The map of a molecular substance shows that a molecule consists of an
arrangement of atoms with a relatively high electron density between the
Bonding I: Ionic and Covalent 67
atoms. It therefore appears that the bonding forces in this case are due to
the sharing of electrons. Any electrons shared between two nuclei will be
attracting both nuclei, which will result in a net attractive force holding the
nuclei together (see Figure 7.3).

+
e
$
Figure 7.3. A shared electron holding two nuclei together

The bonding which involves ions is called electrovalent or ionic bonding


and that which involves sharing of electrons is called covalent bonding. In
order to see how the two types of bonding arise it is necessary to consider
the electronic configurations of the elements involved.

lonic Bonding
We shall first consider a specific case of electrovalent combination, that
between sodium and chlorine. A sodium atom contains 11 nuclear protons
and has the electron structure 2, 8, 1. It differs from the nearest noble gas
electron structure, that of neon, 2, 8, by the presence of one extra electron
in the third principal shell. A free chlorine atom contains 17 nuclear
protons and has the electron structure 2, 8, 7. It differs from the argon
electron structure, 2, 8, 8, by a deficiency of one electron in the third
principal shell.
During chemical combination of the sodium and chlorine atoms, the
single electron from the outermost orbit of the sodium atom passes over
to the outermost orbit of the chlorine atom. In this way, two ions are
produced. The sodium ion is positively charged, as Nat, by the nuclear
proton left in excess after the departure of the electron, and the electron
structure is now 2, 8. The chlorine ion is negatively charged, as Cl~, by the
acquired electron, and its electron structure is now 2, 8, 8. In both cases,
the ions have now the electron structure of a noble gas (neon and argon
respectively) with the external electron octet. They differ, however, from
the neutral atoms of these noble gases by carrying their respective ionic
charges. No molecules are formed in electrovalent combination. Any
experimental quantity of sodium chloride contains an enormous number
of chloride and sodium ions in equal proportions. The electrical attraction
resulting from their opposite charges constitutes such chemical “bond” as
exists at all. The ions arrange themselves into a crystal lattice (see Figure 7.4)
—in this case, a face-centred cube—but there is no specific pairing of ions
68 Physical Chemistry

Figure 7.4. Crystal lattice of sodium chloride

which could be considered as a molecule of sodium chloride. The change


during combination is expressed diagrammatically below.

x (ou e)
x XxX Ono ,
x ie)
x x fe) 0 Oo
: llep . @) fe l7ep ae
x (e)
x x (oe)

oO
Na atom Cl atom
Electrons 2,8,1 Electrons 2,8,7

(omme)
x xX (oe)
x (@)
x x Oo oO Oo Oo
. llep + sh 17®@p aa
x ©)
x x (ome)

(oe)
Na?* ion ion
Electrons 2,8 Electrons 2,8,8

Before combination Before combination


Neutral Na atom, electrons, 2,8,1 Neutral Cl atom, electrons, 2,8,7
After combination After combination
Na* ion, electrons, 2,8 Cl ion, electrons, 2,8,8
Bonding I: Ionic and Covalent 69
In terms of subshells, each sodium atom loses one electron from a 3s
subshell and each chlorine gains an electron into a 3p subshell.

Na + Cl —> Nat + Cla


1s? 2s? 2p® 3s 1s? 2s? 2p® 3s? 3p> ~—s1s? 2s? 2p® _ 1s? 2s? 2p® 3s? 3p®

An ionic compound will be formed if the total energy change involved


is favourable. In the case of sodium chloride, the energy required to remove
an electron from sodium is more than made up for by the energy released
when the electron is gained by chlorine and the oppositely charged ions
come together to form the ionic lattice. A more detailed discussion of the
energy changes involved in the formation of ionic compounds is given on
page 189.
Other examples of ionic compounds are shown in the following table.

Potassium sulphide
Before combination: Two K atoms One S atom
Nuclear protons 19 each 6
Electrons 2,8,8,1 2,8,6
2,8,8,1

After combination:
Nuclear protons 19 each 16
Electrons 2,8,8 2,8,8
eee
Two K* ions One S?- ion
Potassium sulphide, 2K*+-S?-

Magnesium oxide
Before combination: One Mg atom One O atom
Nuclear protons 12 8
Electrons 2,8,2 2,6

After combination:
Nuclear protons 12
Electrons 2,8 2,8
One Mg?* ion One O?- ion
Magnesium oxide, Mg?*+-O?-

The three examples discussed so far exhibit certain features in their


bonding which are common to a large number of ionic compounds, i.e.
1. A metal loses one or more electrons to form a positive ion (cation).
2. A non-metal gains one or more electrons to form a negative ion (anion).
3. By obtaining an external octet of electrons (complete s and p subshells),
both elements attain an electron structure which is identical to that of
the nearest noble gas.
4. The ratio of the numbers of ions combining to form the lattice is such
that the compound is neutral over all.
70 Physical Chemistry
The octet rule does not apply to d-block elements. For example, zinc has
an electron structure of 2, 8, 18, 2 and zinc ions have a relative charge of
2 — and therefore an electron structure of 2, 8, 18. Also, more complex
ions, such as NH} and SO?2-, which will be discussed in detail later,
achieve octets by electron sharing within the ions.

Characteristic properties of ionic compounds


1. Ionic compounds contain no molecules. They are aggregates of ions.
If the ions are made mobile by dissolving the compound in water or by
melting it, the resulting solution or melt will conduct an electric current
with decomposition, i.e. ionic compounds are electrolytes.

2. The electrical forces between the oppositely charged ions are powerful.
Consequently, ionic compounds form comparatively rigid crystal
lattices and are solids. They melt at high temperatures (compared with
simple covalent compounds) and are non-volatile.

3. Ionic compounds are rarely soluble in organic liquids, but many of


them dissolve in water.

Covalency
®

We shall consider, as typical of covalency, the formation of a chlorine


molecule from two chlorine atoms. These two atoms each have an electron
structure of 2, 8, 7. The seven electrons in the third principal quantum
shell are distributed as

3s? 3p®

Ny
Each chlorine atom has one unpaired electron and is one electron short of
the nearest noble gas structure. No other element is available from which
electrons may be obtained; therefore the stable pairing of electrons can
only be achieved by the unpaired electron from each chlorine atom being
shared between the two atoms. This means that neither atom actually
achieves a noble gas structure but, by forming the shared pair, they each
have a share of eight electrons. The theory is expressed diagrammatically
below.
x Xo ok
x x x
Cie Clee
MR OX
Bonding I: Ionic and Covalent 71
A shared pair of electrons is said to constitute one covalent bond and is
conventionally represented by a stroke, —:

Cci—Cl
It is important that you are not misled, by either of these diagrams, into
thinking that the sharing of a pair of electrons is a static affair. Electrons
are in a continuous state of motion and, as discussed on page 55, it is not
possible to define both the path and velocity of an electron. It is therefore
best to regard the shared pair as a pair of electrons which spend most of
their time between the two atoms. (See page 80 for a discussion of covalency
in terms of atomic and molecular orbitals.) Other simple examples of
covalency are shown below.

Ammonia Steam Hydrogen chloride

H H
@x x@ ee
®
06 Nix
@ @
ereaitcer:
x
Salat
® ®

@x 6@ @ ©
H

i
tH
i
O—H Cil—H
4 x = hydrogen electron
® — electrons of other atoms

The following are rather more complicated examples of covalent molecules.


Note that some of these molecules include atoms which share more than
one pair of electrons and are thus being held together by multiple bonds.
In all these examples, no attempt is made to indicate the three-dimensional
shape of the molecules. This will be dealt with later, page 78.

Properties of covalent compounds


1. Covalency produces genuine molecules, never ions. Consequently,
covalent compounds are non-electrolytes, except for a few examples
which form ions when they dissolve in water (e.g. hydrogen chloride).
W2 Physical Chemistry

H
i
Ethane, CoHe

Ho C°H H—C—H
|
Ho oH H—C—H
|
H
Tetrachloromethane, CCl,

+
oo Ba eo
Cl

Ethanol, C2H;0H
HH
foye) © 00

Ethana! (acetaldehyde),
CH,CHO H;
Le ots
H—C—C
ere AS
H
BC)
‘ae
oe
oeee
Ethene, CoHy
i H
|
C=C
hed
0000

H H

|
Methyl cyanide,
CH3CN

@ H—C—C=
00
00 |
H
Carbon dioxide,
CO,
O=C=
Bonding I: Ionic and Covalent 73
we Unlike a solid ionic compound, which consists of a continuous giant
lattice of oppositely charged ions, the covalent compound consists of an
arrangement of molecules called a molecular lattice. Whereas the forces
holding the atoms together within the molecules may be of comparable
strength to the forces holding ions together, the forces which hold the
molecules together are weak. The molecules.which..make up the
molecular lattice are easily separated and simple,covalent compounds
are always volatile. Many of them are gases (e.g. chlorine, Cl,; oxygen,
O,; methane, CHs; ammonia, NHg; and carbon dioxide, CO,) or
volatile liquids (e.g. trichloromethane, CHCl,; ethanol, CH;CH,OH;
and carbon disulphide, CS.). More complex covalent compounds may
be solids, e.g. the higher alkanes, C,H2,+2, where 7 is 18 or more.
The weak forces which hold the molecules together in a molecular
lattice are known as van der Waals’ forces (page 103). Mh

3. Covalent compounds are often insoluble in’ water, unless they contain
such groups as the hydroxyl (—OH), but they usually dissolve in other
covalent, organic solvents such as benzene or ethoxyethane (ether).

It is important to realize that the two types of bonding described so far,


ionic and covalent, represent two extreme cases. In fact, covalent bonds
between atoms of different elements can rarely be said to be entirely
covalent as they usually have some ionic character (page 97) and only
certain metal salts can be thought of as 100 per cent ionic (page 191).

Dative or Co-ordinate Covalency


This is a type of covalency. A shared pair of electrons is formed between
the combining atoms in both these bonding types. The difference is that in
covalency each atom contributes one electron to the shared pair, while in
a dative bond one atom provides both electrons. The other atom only
accepts the shared pair.
An example of this type of covalency is the formation of the stable
molecular compound between ammonia, NHs;, and boron trichloride,
BCl. The latter compound is purely covalent but boron, having only three
electrons in its outer energy shell, makes up only a sextet of electrons by
covalent sharing. Ammonia is a covalent compound with three shared pairs
of electrons and an unshared pair (lone-pair). The lone-pair is donated from
the nitrogen to be shared between the nitrogen atom and the boron atom,
thus making up the boron octet of electrons.
74 Physical Chemistry
xx
rca 7
xx OX Ox
x ° o

xX Ox xO

tel
xx
H

x a x o = electron of B or N
x cl x H x = electron of Cl or H
xx *0O OX
S ChaceBe oeNaet
x ce) ce) fe)

xX Ox Ox

A co-ordinate or dative bond may be represented in several ways: either as


a stroke, —, as in normal covalent bonding; e.g.

Cl H
Cok at
ua
or as an arrow, — ,which indicates the direction in which the electrons are
donated, e.g.

piney
lis oa
xk ig &

CloH

or by a stroke and, because one atom has in effect lost an electron and the
other has gained one, a positive and negative charge.

Cl H
|-.1
Cl—B—N—H
Wal
Cl H
Bonding I: Ionic and Covalent 75
The arrow is the Most common and preferred way of representing the
bond. Two more important examples of co-ordinate or dative covalency
are the ammonium, NH}, and the hydroxonium, H,O*, ions.

The ammonium ion, NH}, and ammonium chloride


The molecule of ammonia, NHs, is covalent and contains a lone-pair of
electrons. (See page 73). It is electrically neutral. If ammonia comes into
contact with an acid, e.g. HCl in ionized condition, H*+Cl-, the hydrogen
ion (or proton) accepts a share in the lone-pair of electrons, combines in
this way with the NH; molecule, and produces the ion, NH;, positively
charged by the accepted proton. This particle is associated by electro-
valency with the Cl-, which has accepted an electron from a hydrogen
atom to supply the ion, H+. Ammonium chloride thus shows all three of
the common valency types. The electrovalency is predominant in deter-
mining the properties of the salt, which is, consequently, a crystalline solid,
soluble in water and an electrolyte when in solution.

H ate

i 33

Formula of ammonium chloride

The arrow, as a symbol for a co-ordinate bond, is a useful means of


indicating the source of the electrons which constitute the bond, but in
cases such as the ammonium ion it can be misleading as it suggests that one
bond is different from the other four. In fact this is not the case as the
nitrogen is bonded to four atoms of the same element and all four bonds,
once they have been formed, are identical.

The hydroxonium ion, H;0*


The characteristic behaviour of an acid is the production of ‘hydrogen
ions’ in water. In their simplest form, H*, these are protons. The observed
electrical conductivities of acids are, however, too low to admit of the
76 Physical Chemistry
presence of such a light, highly charged, mobile particle and it is certain
that the proton is at once hydrated to form the hydroxonium ion.

H +
ox
H re) O .e)
O H
OOo
© = electron from O
x = electron from H

i ae

H—O—H

Complex lons
The so-called complex ions involve co-ordinate or dative covalency, and
are formed by ions or molecules which possess lone-pairs of electrons,
donating their lone-pairs to be shared between themselves and a metal ion.
The group which donates the lone-pair is called a Jigand and the metal ion
is usually an ion of a transition metal.

CN).]*~ . ,
The hexacyanoferrate(II) ion, [Fe(
This complex ion is formed by the addition of an excess of a solution of
cyanide ions to a solution containing iron(II) ions. Each CN~ has a
lone-pair of electrons which it is capable of donating. A neutral atom of
iron has the following electron arrangement:
Fe? 22,8514, 2 or ist 2s* 2p e357, op ode 4s"
An iron(II) ion has the structure
ethene ici or 182. 2S eae. soa ods
Each iron(II) ion is capable of co-ordinating with six cyanide ions, thus
giving it a share in another twelve electrons. This means that the iron atom
becomes associated with the same number of electrons as the nearest noble
gas, krypton, 2, 8, 18, 8.
The charge on the complex ion produced is 4—, which is the algebraic
sum of the charges on the Fe?* and the six CN-.
Fe?+ + 6(CN-) > [Fe(CN)¢]?-
Each iron combines with six ligands and is said to have a co-ordination
number of six.
Bonding I: Ionic and Covalent 77
This particular ion is a complex anion, hence the use of the -ate ending
in the name.

Tetraamminecopper(II) ion, [Cu(NH;).]?+


This ion is formed when an excess of ammonia solution is added to a
solution containing copper(II) ions. The ammonia molecule is electrically
neutral and it has a lone-pair of electrons. A neutral atom of copper has
the electron structure

Cu PeSial Sach or 133(2s?,:2p°;3s7,:3p*,, 3d?°; 4s}


A copper(I]) ion has the structure

ree 2 S17 or is2s7, 2p, 35°, 3p, od?

Each copper(II) ion can co-ordinate with four ammonia molecules (the
resulting number of electrons associated with the copper atom being one
less than the krypton electron structure). The complex ion formed has a
net charge of 2*.

Cu?* + 4NH; — [Cu(NHs).]?*

Strictly, the formula of the ion is [Cu(NH3), (H.O),]?* and should be


called tetraamminediaquacopper(II). The two water ligands are more
distant from the copper ion and are held less strongly than the ammonia
ligands. The structures of this ion and other complex ions are discussed in
Inorganic Chemistry, the companion volume to this book.

The Importance of Electron Pairing


Some elements in the third period and above form compounds in which
they do not adhere to the octet rule, thus indicating that the pairing off of
unpaired electrons is of more fundamental importance than achieving a
noble gas structure.
Phosphorus forms two chlorides: the lower one, PCl3, conforms to the
octet rule, but the higher one, PCl,, does not. Phosphorus has five electrons
in the outer shell and they would normally be distributed as

Chlorine, with its electron structure of 2, 8, 7, has one unpaired electron.


When PCI, is formed all the unpaired electrons are paired off and both
atoms obtain a share in a complete octet of electrons,
78 Physical Chemistry

(oye)

6cls Cl
xO (eye) |

Psa Pci
xO (eye)

2c13 cl
(eye)

The formation of PCI, can be accounted for by the electrons of phosphorus


rearranging to a state of slightly higher energy in which there are five
unpaired electrons.

filha ale
Each of these unpaired electrons forms a shared pair with an electron from
a chlorine atom.

fe Ob ie
25 2% Cl

cS. P +,20 0 Cr '


reysoil ie) oClo >»
OCle
[ofe) oO Cl
cl

This results in phosphorus having a share in ten electrons in its outer shell.
Another example of expansion of an octet is in the formation of sulphur
hexafluoride in which the sulphur gains a share in twelve electrons. The
expansion above the octet does not occur in the second period (Li—Ar)
as there is not a corresponding d energy level available.

The Shape of Covalent Molecules


The covalent bonding force arises from the electrostatic attraction that a
shared pair of electrons exerts on two nuclei and, as such, it is specific to
the two atoms between which the two electrons are shared. On the other
hand, the bonding forces within an ionic lattice are due to the electrostatic
attraction between oppositely charged ions. Such forces will be non-
directional as a positive ion will be able to attract negative ions from all
directions.
Bonding I: Ionic and Covalent 79
Thus, one very important distinction between covalent and ionic
bonding is that the covalent bond is directional and the ionic bond is not.
The directional nature of the covalent bond and the repulsive forces
between the pairs of electrons involved account for the observed shapes of
covalent molecules. (The shapes of covalent molecules can be deduced
from X-ray and electron diffraction studies (page 64) and they are usually
defined by stating the angles between the bonds.)
The four covalent bonds of carbon are arranged with an angle of 109.5°
between any pair of bonds. This is explained in terms of the repulsions
between the bonding pairs of electrons, resulting in them being as widely
spaced as possible and forming a tetrahedral shape.

|
Aes
The angles between the N—H bonds in ammonia are found to be
107.3°. This angle is slightly less than the tetrahedral angle and is explained
by the repulsions between three bonding pairs and the slightly greater
repulsion from a lone-pair (non-bonding pair). The reason for the greater
repulsion of the non-bonding pair is thought to be that its electron cloud
is not stretched out by the presence of another atom, which results in
there being a greater concentration of negative charge nearer the nucleus
and hence nearer to the other bonds.

H 107.3°

Water has two bonding pairs and two non-bonding pairs. Therefore, as
would be expected, the angle between the OH bonds, 104.5°, is slightly less
than that between the NH bonds in ammonia.

xo 104.5°
80 Physical Chemistry
The shapes of some other molecules and ions are given below.

H
oO

ae Lye
j H— | H
ty H
BF, is planar as boron NH, is tetrahedral as there
does not have a lone-pair. are four bonding pairs.

itor Pastor
“ipl
a pied

A single molecule of PF, exists _ and PFs are both octahedral.


in the form of a trigonal bipyramid.

The Orbital Approach to Covalent Bonding


As previously stated (page 55), it is not possible to determine both the path
and velocity of an electron but it is possible to determine the rough outline
of a shape in which the electron is most likely to be found. This shape is
called an atomic orbital,
The formation of a covalent bond can be considered as the overlapping
of two atomic orbitals, each containing one electron, to form a new shape,
called a molecular orbital, in which there is a high probability of finding the
pair of electrons forming the bond.
When a hydrogen molecule is formed, the 1s orbital of one hydrogen
atom overlaps the 1s orbital from another to form a molecular orbital as
shown below
Bonding I: Ionic and Covalent 81
The bond formed by the overlapping of two s orbitals is called a
(sigma) bond. When a covalent bond is formed by two p orbitals (as
shown below), each containing one electron of opposite spin, it is called
a 7m(pi) bond.

ee
GX
The z bond, unlike the o bond, has a very low electron density along the
axis which joins the two nuclei.
For carbon to form four covalent bonds it must have available four
unpaired electrons. These are obtained by the electrons of carbon re-
arranging to form a state of slightly higher energy than the ground state of
carbon.

2 2p
C (ground state) itu vehicle |

2s 2p
C (rearranged)

In order to form four identical bonds, as in methane, it is considered that


the one electron from the s orbital and the three electrons from the p
orbitals are amalgamated to form four bonds of equal energy which point
towards the corners of a tetrahedron.

SG
The changing of a number of orbitals of more than one type to form the
same number of orbitals of equal energy is called hybridization. The
82 Physical Chemistry
hybridization involved in the tetrahedral covalency of carbon is called sp®
hybridization which indicates the number and origin of the electrons
involved. In methane each of the sp* hybrid orbitals overlaps with an s
orbital of a hydrogen atom and forms a o bond.
In ethene (ethylene) sp? hybridization occurs and each hybrid orbital
forms a o bond.

a ae eee
oe “AG
The fourth, unpaired, electron of each carbon atom is in a p orbital.
The two p orbitals overlap to form a z bond.

NOs Gforce Naeem


Or (ON: VEEN
It is the z bond of ethene which is particularly susceptible to attack by
an electron seeking group (electrophilic) during the characteristic addition
reactions. The z bond also accounts for the geometric isomerism (see
below) exhibited by alkenes. This type of isomerism is due to the restricted
rotation of the carbon-carbon double bond. The rotation of such a bond
is difficult as it would destroy the overlap between the two p orbitals
forming the 7 bond.

H CO.H HO.C H
ees
C
I |
C C
ra on y ON
CO,H H CO.H
cis-butenedioic acid trans-butenedioic acid
(maleic acid) (fumaric acid)

Delocalization
Benzene

Using the theory that a covalent bond results from a shared pair of
electrons, benzene requires an arrangement of alternate double and single
bonds.
Bonding I: Ionic and Covalent 83

There is a lot of evidence to suggest that this does not represent the true
structure.

1. X-ray and electron diffraction experiments show that the carbon-carbon


internuclear distance is identical for all six bonds. The bond length is
0.139 nm which is intermediate between that of a single carbon-carbon
bond, 0.154 nm, and a double carbon-carbon bond, 0.134 nm.

2. The experimentally determined standard enthalpy of hydrogenation of


benzene to form cyclohexane, i.e.

C.H,(1) + 3H.(g) oe C,H; .2()) AH®? = —208 kJ mol7?

is less than three times the standard enthalpy of hydrogenation of a


normal carbon-carbon double bond as found in alkanes.

C.H.(g) + H.(g) > C2H,(g) AH® = —120kJ mol!


For three double bonds the predicted enthalpy change is —360 kJ. The
difference between these values, 152 kJ mol-+, represents the margin by
which the actual structure of benzene is more stable than the theoretical
alternate double and single bond structure.

From such evidence it is clear that all six bonds in benzene are identical
and involve more electron sharing than a normal carbon-carbon single
bond but less sharing than a normal double bond. It is considered that such
a structure is achieved by the six electrons, which might be expected to
form the second bond of each double, being shared over all six carbon
atoms. The sharing of electrons over more than two atoms is referred to as
delocalization and energy associated with the resulting extra stability is
known as the delocalization energy.
Using the molecular orbital approach, each carbon atom forms three
normal covalent bonds by sp? hybridization. The fourth electron from each
carbon atom is in a p orbital at right angles to the plane of the benzene
84 Physical Chemistry
ring. Each of the six p orbitals overlaps with both adjacent orbitals
producing delocalized z orbitals as follows:

This view of the bonding in benzene leads to the coventional symbol for
benzene being

Carboxylic acid anions


Using normal covalent bonds the ethanoate (acetate) ion could be written
in two ways:

Criz—C or CHe—C

It is not suggested that these two structures are isomers but that, if one
particular ion is being considered, there are two possible ways in which
the bonds could be arranged.
The carbon—oxygen bond lengths are found to be identical, which
indicates that neither of the above formulae are correct. The true structure
is intermediate between those given above and is formed by electrons being
delocalized over all three atoms. The true structure could be represented
as follows:
Bonding I: Ionic and Covalent 85
Nitrate ion
The bonds in a particular nitrate ion could be arranged in several ways;
three are

However, delocalization occurs and all three nitrogen-oxygen bonds are


identical. The bond length is intermediate between the length of a N—O
bond and a N=0O bond.

Sulphate ion
Using the available electrons, there are several different ways of repre-
senting the bonds in a sulphate ion, but it is found that the ion contains
four identical bonds arranged tetrahedrally around the central sulphur
atom. Thus, once again, delocalization must have occurred.

ieee
S
00
8. Crystal Structure

It is important to make the general distinction between giant lattices and


molecular lattices. The term giant lattice is used to describe the large
continuous arrangements of particles (atoms or ions), which make up
certain solids. The bonding forces in a giant lattice may be directional
(covalent), as in diamond, or non-directional, as in metals and ionic
compounds. A molecular lattice, unlike the above, consists of an arrange-
ment of molecules in which the forces holding the molecules together are
very much weaker than the forces which hold the atoms together within
the molecules.
As previously stated, the geometrical arrangement of the particles (atoms,
ions, or molecules) which make up a lattice is determined by X-ray and
electron diffraction techniques (page 64).

Giant Lattices

1. Metal lattices
The bonding forces within a metal are considered to arise from the
delocalization of the ‘valency’ electrons of the metal atoms. The picture is
that of an arrangement of metal ions with the valency electrons free to
move within the metal. It is the mobility of these electrons which accounts
for the electrical conductivity of the metal.
In a metal, unlike the ionic lattice of a salt, there is only one type of ion.
Therefore, from the geometrical point of view, the packing of ions in a
metal can be compared to the packing of spheres of identical size.
When a single layer of identical atoms or spheres is packed as closely
as possible, it is found that each atom is surrounded by six near neighbours,
as shown in Figure 8.1. If a second layer of spheres is added so that they
rest in the hollows of the first, the plan view would look like Figure 8.2.
A careful examination of Figure 8.2 reveals that if yet another layer of
atoms is added, they can be placed in one of two possible positions.

(a) The third layer can be placed vertically above the first layer (position
1). If the atoms in the first layer are denoted by A and those in the
second layer by B, then, because the third layer is a kind of repetition
Crystal Structure 87

Figure 8.1.

Figure 8.2.

of the first, and the fourth of the second, and so on, the structure is
called an ABA type.
(b) Alternatively the third layer can be placed in position 2 where they
would not be vertically above either of the two previous layers. This
structure is called an ABC type.
It is worthwhile looking more closely at each of these alternative
structures.

ABA (called hexagonal close-packing)


Consider atom X (Figure 8.2) which is in the second layer; it has twelve
near neighbours, six in its own layer, three in the first layer, and three in
88 Physical Chemistry
the third layer (the positions of which are indicated by the three 1s around
atom X). The atom is said to have a co-ordination number of 12.
If the structure is rotated about an axis at right angles to the layers of
atoms, six identical positions are possible. This axis is called a six-fold axis
of symmetry and the structure is known as hexagonal close-packing.

ABC (called cubic close-packing or face-centred cubic)


As in the hexagonal close-packing, each atom has a co-ordination number
of 12, having six near neighbours in its own layer, three in the layer below,
and three in the layer above.
O A

®
e¢$e

® @ @6 ® @c

O A ;
Figure 8.3.

Figure 8.3 represents the side elevation of parts of four layers (i.e.
ABCA) looked at from slightly above the plane of the layers. One atom in
the uppermost layer (A) is shown, six atoms of the next (B), six atoms of
the next (C), and one atom of the lowest (A).

Figure 8.4,
Crystal Structure 89
Figure 8.4 is the same as Figure 8.3 except that construction lines have
been added to show that the atoms form a cube, standing on one corner
with an atom at the centre of each face. The structure is called cubic
close-packing or face-centered cubic. In this case, a vertical line through
the planes of atoms is a three-fold axis of symmetry.

Body-centred cubic
This third type of structure which is adopted by some metals cannot be
described as close-packed. Each atom has a co-ordination number of 8.
The structure is described as body-centred as each atom is at the centre of
a cube formed by eight others (Figure 8.5).

Figure 8.5.

Examples of metals which adopt each type of structure

Hexagonal close-packed magnesium and zinc,


Cubic close-packed copper and aluminium,
Body-centred cubic alkali metals.

A considerable number of metals can adopt two structures. For example,


iron is known to occur in both cubic close-packed and hexagonal close-
packed forms.

2. lonic lattices
When a sphere is placed in a hollow in a previous layer, it comes into
contact with three other spheres and the overall shape of this unit is
tetrahedral. In between the four spheres there is a small space which is
called a tetrahedral site (see Figure 8.6).
90 Physical Chemistry

Figure 8.6.

When a triangle of spheres is placed on the top of another triangle of


spheres, in such a manner that the centre of one triangle is immediately
above the centre of the other, the overall shape of the spheres is octahedral.
The space which is left at the centre of these spheres is called an octahedral
site (see Figure 8.7).

Figure 8.7. :

Figure 8.8 shows all the tetrahedral (T) and octahedral (O) interstitial
sites for two layers of close-packed atoms. It is very useful to relate the
structures of ionic compounds to the various metal structures and the
interstitial sites.

Figure 8.8.
Crystal Structure 91
The main factors which determine the type of lattice adopted by an
ionic compound are:
ih The relative charges on the ions which dictates the relative numbers of
10ns present.
2. The relative sizes of the ions.
The size of an ion is denoted by its ionic radius which is determined by
X-ray diffraction methods. The X-ray analysis gives only the inter-nuclear
distance which is the sum of the two ionic radii. It is necessary to decide on
a value for one particular ion and then by investigating a whole range of
salts it is possible to build up a table of ionic radii.

Atomic radius lonic radius


(nm) (nm)
Na 0.19 Nat 0.095
Kone 0.235 Ko 0'133

Cl 0.099 Cl- 0.181


O 0.074 O7- 0.140

It can be seen from the table that for a positive ion the ionic radius is
much smaller than the atomic radius (e.g. K* and K), and in the case of a
negative ion the reverse is true (e.g. Cl” and Cl). This difference is con-
sistent with the protons being in excess of the electrons in a positive ion
and vice versa for a negative ion.
Some of the more common types of ionic lattice are described below.

Sodium chloride (ionic radii, Na+ = 0.095 nm, Cl~ = 0.181 nm)

Figure 8.9.

The geometry of this structure may be described in the following


manner.
92 Physical Chemistry
1. Each chloride ion is in close contact with six sodium ions and each
sodium ion with six chloride ions. The crystal is said to have 6:6
co-ordination.
2. Considering the chlorides only, it can be seen that although the ions
are not actually in contact, they can be regarded as being in a face-
centred cubic arrangement (i.e. ABC type which is cubic close-packing).
The sodium ions are also in a face-centred cubic arrangement. The
whole lattice consists of two interlocking face-centred cubes with the
sodium ions appearing in the octahedral sites of the chloride arrange-
ment and vice versa for the chlorides (Figure 8.10).

Figure 8.10.

Caesium chloride (ionic radii, Cs* = 0.167 nm, Cl- = 0.181 nm)

Cs* surrounded
by eight Cl-

Cl surrounded
by eight Cs*
Figure 8.11.
Crystal Structure 93
This structure resembles sodium chloride in that it is a cubic structure, but
it differs in the following respects:

1. Each caesium ion has eight chloride ions as near neighbours and each
chloride has eight ceasiums as near neighbours. Hence the co-ordination
is 8:8.

2. Each type of ion is arranged in a simple cubic form. The whole structure
may be regarded as two interlocking simple cubic forms with each
caesium ion appearing at the centre of a cube of chlorides and vice versa
for chloride ions.

Note that, although in some ways Figure 8.11 resembles a body-centred


cubic arrangement, it is not correct to label it as such, because the ion at
the centre is of a different type from the ions at the corners of the cube.

Zinc blende (ionic radii, Zn?* = 0.074 nm, S?- = 0.190 nm)

Figure 8.12.

1. There is 4:4 co-ordination.

2. The sulphide ions are arranged in a face-centred cubic pattern.

The zinc ions can be regarded as being in a simple cubic pattern with
each zinc ion appearing in a tetrahedral site of the sulphide cubic close-
packed system.
Note that for every atom in a close-packed system there are two
tetrahedral sites. In the zinc blende type of structure half of the sites are
occupied.
94 Physical Chemistry
Fluorite (ionic radii, Ca?* = 0.094 nm, F~ = 0.133 nm)

e-F-
O - Ca?*

Figure 8.13.

1. Each calcium ion has 8 co-ordination around it and each fluoride has
4 co-ordination.
2. The calcium ions can be considered as being in a face-centred or cubic
close-packed arrangement in which, unlike the zinc blende structure,
all the tetrahedral sites are occupied by fluoride ions. The fluoride ions
themselves are in a simple cubic pattern.

3. Giant covalent lattices :


In the two types of giant lattices which have been discussed, the bonding
forces have not been directional in nature. The third type of giant lattice
involves covalent bonds which are directional.
Even though the nature of the bonding forces is different, the fact that
they are giant lattices and not molecular means that, like metals and ionic
compounds, it is necessary to overcome the main bonding forces when
melting these substances.

Diamond

Figure 8.14,
Crystal Structure 95
Each carbon atom is covalently bonded to four other carbon atoms giving
the structure a co-ordination of 4. The system is cubic and the atom at the
centre of a tetrahedron can be compared to a zinc ion in a tetrahedral site
in the zinc blende structure. Remember, in the case of diamond it is the
number of unpaired electrons possessed by each carbon atom which
determines the 4 co-ordination.

Graphite

Figure 8.15.

Graphite consists of a layer structure. Within each layer, each carbon atom
is covalently bonded in a planar arrangement to three others. The C—C
bond length in diamond is 0.154 nm which is the normal length for a single
covalent bond between carbon atoms. In graphite the bond length is
0.142 nm which suggests that there is some delocalization of the fourth
electron which results in extra bonding. However, the extra bonding is not
as powerful as that in benzene in which the bond length between the
carbon atoms is 0.139 nm. The fact that graphite conducts electricity
suggests that the delocalization occurs throughout the layer.
The distance between the layers in graphite is quite large at 0.34 nm and
indicates weaker bonding forces. These are van der Waals’ forces (page 103)
and it is their weakness which allows the layers to move over each other
and so result in graphite being soft and suitable for use as a lubricant.

Molecular Lattices
In particular, these lattices are characterized by their low melting and
boiling points. These properties arise because the bonding forces between
the molecules are weak. The important point to remember is that when
these substances melt the covalent bonds within the molecules are un-
affected and the molecules remain intact.
96 Physical Chemistry
lodine

Figure 8.16.

This lattice consists of an arrangement of iodine molecules each of which


contains two iodine atoms bonded together by a single covalent bond. To
melt iodine it is only necessary to overcome tk weak attractive forces (van
der Waals’ forces) between the molecules and, as expected, its melting
point and boiling point are comparatively low at 387 K and 456 K respec-
tively.
;

Naphthalene
The molecular formula of naphthalene is C,)>H, and its structural formula
is

Solid naphthalene consists of an arrangment of these molecules held


together by weak van der Waals’ forces which result in a relatively low
melting point 353 K, and boiling point, 491 K.
9. Bonding II: Intermolecular

The wide range of energies associated with intermolecular bonding


suggests that not all of the bonds are formed in the same way.

Bond Polarity
The covalent bond which holds two atoms together within a molecule is
envisaged as a pair of electrons shared between the atoms. Unless the
atoms are identical, the electrons will not be equally shared. When
unequal sharing does occur it will result in a slight separation of charge.
The atom with the greater share of the electrons will be slightly negative
and the other atom will be slightly positive. This phenomenon is referred
to as bond polarity and the bond is said to be of intermediate type, i.e.
partly covalent and partly ionic.
The relative tendencies of elements to attract electrons are expressed
numerically on an arbitrary scale, as their electronegativities. The larger the
difference between the electronegativities of the atoms forming the bond,
the more unequal the sharing will be. The comparative electronegativities
of some elements are as follows:

F O N Ce © H
(hi leo Mee 8A Beene 1 lle 2 aa |

The values quoted above show that there will be unequal sharing in the
C—Cl bond and in the C—O bond, but that the electrons forming the
C—H bond will be more equally shared.

Polar Compounds
Bond polarity within a diatomic molecule results in a separation of
charge over the whole molecule. The molecule is said to be polar and
to possess a dipole moment which can be expressed numerically by taking
into account the magnitude of the charge and the distance separating the
charges.
eee made up of more than two atoms and exhibiting some bond
polarity does not necessarily have a dipole moment. If the molecule is
symmetrical, the effects of the individual dipoles of the bonds, as far as the
98 Physical Chemistry
molecule as a whole is concerned, cancel each other out. Thus, trichloro-
methane has a dipole whereas tetrachloromethane does not.
H ‘<

C
the|NY ae
Cl Cl Cl Cl
Cl
Trichloromethane Tetrachloromethane

The lack of a dipole where one might be expected is an indication of


symmetry within a molecule. For example, carbon dioxide is made up of
two elements with differing electronegativities but because the compound
is linear, there is no resultant dipole.
O—C—O
Carbon dioxide

Dipole-Dipole Interactions
Electrostatic attractions between the oppositely charged ends of dipoles
account for the intermolecular bonding in such compounds as propanone
and the energy associated with this type of bonding iis about il 100 of that
associated with a covalent bond. s

CH, CH;
arose taienes arees

Ve
CH3 CH3

Hydrogen Bonds
When electrons are withdrawn from a hydrogen atom, either by it being
bonded to an electronegative atom (as in water) or by the presence of
electronegative atoms in the rest of the molecule (as in CHCIs), it acquires
a slight positive charge. Due to the smallness of the hydrogen atom, the
positive charge is more concentrated and is therefore more effective in
forming links with other molecules. Hydrogen atoms in this type of
situation are found to make particularly strong intermolecular bonds with
electronegative atoms which possess lone-pairs (non-bonding pairs) of
electrons. Fluorine, oxygen, and nitrogen are all good examples of such
elements.
An intermolecular bond of this type is called a hydrogen bond and the
energies associated with such bonds are usually of the order of 20 to
40 kJ mol~? as compared with energies of the order of 300 to 400 kJ mol-!
for normal covalent bonds.
Bonding IT: Intermolecular 99
Certain compounds, HF, H,O, and NHs, possess both of the essential
requirements for hydrogen bond formation, i.e. one or more hydrogen
atoms in a suitable environment and an electronegative element with a
lone-pair of electrons.
The effect of the unusually powerful intermolecular forces which exist
between molecules of these compounds is made obvious if their melting
points and boiling points are compared to those of hydrides of the other
elements in their respective groups (see Figure 9.1).

g
8

(K)
temperature
DSoIOo

150
period number

Figure 9.1. Boiling points of the hydrides of


Groups 5, 6, and 7

The melting points and boiling points of HF, H,O, and NH; do not fit
in with the general trends within their groups showing that they possess
unusually powerful intermolecular forces.
Methane, CH,, which is in the corresponding position in Group IV,
fits in perfectly with the general trend for the hydrides of its group as it
possesses neither of the requirements for hydrogen bond formation (see
Figure 9.2, on page 100).
The hydrogen bond is a very specific case of intermolecular bonding and
a few important examples of its occurrence will be discussed in more
detail.
100 Physical Chemistry

250
SnH,

8 GeH,

(K)
temperature
—solOo

100
period number
Figure 9.2.

1. Ice

In the crystal lattice of ice each HO molecule forms:


(a) two hydrogen bonds by virtue of its non-bonding pairs being attracted
to the slight positive charges on hydrogens of other H,0 molecules,
(b) two hydrogen bonds by virtue of the slight positive charges on its
hydrogen atoms being attracted to non-bonding pairs of electrons
associated with oxygens of other H,O molecules.
The four bonds form an approximately tetrahedral arrangement around
each oxygen atom. The resulting structure is cubic and wurtzite in type.
Compared to the more random arrangement of H,O molecules in water,
the ice structure is more open, which results in ice having a lower density
than water. Thus when water is cooled its density increases to a maximum
at 4 °C, after which the more open structure begins to form and the density
decreases.

Hydrogen bonds to es H
lone-pairs of other
H20 molecules
O

NS
A Hydrogen bonds to hydrogen
¢ atoms of other H2O0
molecules

Hydrogen bonding in ice


Bonding II: Intermolecular 101
2. Gypsum
Calcium sulphate occurs naturally in two forms, CaSO1.2H2O which is
gypsum, and the anhydrous form which is called anhydrite. The physical
characteristics of these two forms are quite different. Anhydrite exhibits
no obvious cleavage and is fairly hard (Mohs’ scale 3.5), whereas gypsum
is much softer (Mohs’ scale 2) and has a very obvious cleavage. These
observations, together with evidence from X-ray crystallography, are
interpreted as:

(a) Anhydrite consists of a giant lattice of Ca?*+ and SO?- ions. This
lattice is not easily cleaved along any plane.

(b) Gypsum is a layer structure, with each layer consisting of a lattice of


Ca?* and SO2~ ions. The layers are held together by hydrogen bonds
which act between the sulphate ions in the layers and the water
molecules which are situated between the layers. Due to the comparative
weakness of the hydrogen bonds, very obvious cleavage planes exist
between the layers.

3. Carboxylic acids
In non-polar solvents, such as benzene, a number of organic acids are
found to have relative molecular masses which are approximately twice the
expected values. The double molecules, or dimers, are formed by two
molecules being held together by hydrogen bonds.

A dimer of ethanoic acid

Ethanoic (acetic) acid, CH;CO.H, and benzenecarboxylic acid (benzoic


acid), C,H,CO.H, both form dimers in this way.

4. Biochemistry
Proteins are essential constituents of all living material and the structure
adopted by a protein depends to a large extent on hydrogen bond forma-
tion. Amino acids combine in various sequences to form compounds of
large relative molecular mass which are known as proteins. In general
terms;
102 Physical Chemistry
H R O ne t ye

eA +4 Vienkidan

7 | > whe H O—H

Ha. O R’ O
| lect Aggie 4a
= N—C—C C—C
Ves PE OAL OS
etc H N H etc

th
The amino acid chain has the necessary requirements for hydrogen bond
formation, namely, hydrogen atoms (H) bonded to an electronegative
atom (N) and electronegative atoms with lone-pairs of electrons (O). It is
hydrogen bonds between these sites which keep the protein chain in a
coiled form, as shown in Figure 9.3.

Figure 9.3.

Another striking illustration of the occurrence of hydrogen bonding is in


the structure of deoxyribonucleic acid (DNA), which is the substance
responsible for the process of heredity. It consists of two nucleic acid
helices joined together by hydrogen bonds between the hydrogen atoms of
N—H groups and other nitrogen atoms or the oxygen atoms of C=O
Bonding I: Intermolecular 103
groups. When the hydrogen bonds break, each helix can then, by reforming
the specific hydrogen bonds, produce a replica of the original DNA.

van der Waals’ Forces


Certain compounds and elements, for example alkanes, iodine, and the
noble gases, possess none of the characteristics required for hydrogen bond
formation or dipole-dipole interaction. However, these substances can be
obtained in liquid and solid form, and so intermolecular forces of some
description must exist. The forces are called van der Waals’ forces and are
thought to be electrostatic in nature and due to slight molecular dipoles
which are temporarily induced when molecules come into close contact.
In support of this explanation, it is found that ‘straight’ chain alkanes are
more easily liquefied than ones with branched chains. The more linear
shape of the straight chain, as opposed to the roughly overall spherical
shape of the branched chain, provides more possibility of interactions
between the electron clouds of neighbouring molecules.
The energy associated with van der Waals’ forces is usually about ten
times less than that associated with hydrogen bonds.

Covalent and van der Waals’ radii


As mentioned on page 96, solid iodine is a molecular lattice in which each
molecule consists of two atoms covalently bonded together and the
intermolecular forces are of the van der Waals’ type. This arrangement
results in two different internuclear distances for iodine atoms. Iodine
atoms within a molecule are relatively close together and half the inter-
nuclear distance is called the covalent radius of iodine. Adjacent atoms
from neighbouring molecules are further apart and half the internuclear
distance is known as the van der Waals’ radius for iodine.
The types of structure and bonding forces which have been discussed in
this text are shown summarized overleaf.
104 Physical Chemistry
Summary of types of bonding and crystal lattice
Evidence Type of lattice Type of bonding

es metallic
atomic —
covalent
Dae
gu itt
ionic ———————— ionic
m.p., b.p.,
solubility,
conductivity,
X-ray, and
electron:
diffraction.
atomic
(noble gases)
molecular ae nse van der Waals
et molecular
(covalent dipole-dipole
bonds within
the molecules)
hydrogen bonds

In the case of a diatomic hydride (e.g. HF) there are contributions from
both dipole-dipole interactions and the more directional hydrogen bonds.
10. Colloids and the Colloidal State

The founder of the science of colloids is generally taken to be Graham


because of work done by him in this field between 1851 and 1861.
Colloids were once regarded as a separate class of materials, such as
starch, gum, gelatine, and glue, which had the characteristics of not
diffusing through a colloidal membrane, such as parchment paper, and
not being generally known in crystalline form. The word colloid is derived
from the Greek kolla, meaning glue. The contrast was with crystalloids,
such as common salt, copper sulphate, or oxalic acid, which diffused
readily through the membrane and were readily obtained as crystals. It is
now known, however, that most elements and compounds can be obtained
in colloidal solution in a medium in which they are not soluble in the
ordinary sense of that term, so that colloids are regarded collectively as
being in a particular state of matter rather than a separate class of
materials.
In ‘true’ solution, the dissolved substance is (with very few exceptions)
in the state of single molecules or ions which are distributed in the solvent.
The whole is regarded as a homogeneous mixture. The solute is in its
ultimate state of division for chemical purposes. At the other end of the
scale, there are insoluble particles so large that they settle out of the liquid.
Such a mixture is definitely heterogeneous and is called a suspension. A
colloidal solution occupies the intermediate region between true solution
and suspension. At both ends, the line of division is indefinite and the size
of colloidal particles passes gradually into that of suspended particles at
one end of the scale and that of the molecules and ions of true solution at
the other. In general, colloidal particles have a diameter of the order of
1-10? nm. Above this range, suspended particles begin to appear; below it,
the particles cannot be distinguished from those of true solution.
Colloidal solutions are generally referred to as sols; if the dispersion
medium is water, the term hydrosol (or aquasol) is employed. In some cases,
fairly concentrated colloidal solutions are obtainable, and they may set to
a jelly-like material. In that case, they are referred to as gels (or hydrogels
if aqueous). The colloidal particles are called the disperse phase and the
liquid in which they are carried (usually water) is called the dispersion
medium.
106 Physical Chemistry

Preparation of Sols
Since the disperse phase is, on the average, intermediate in particle size
between suspension and true solution, sols can be made either by aggre-
gating the particles of true solution or further dispersing those of suspen-
sion. The following are some typical examples of both types. In general,
each sol will need purification by dialysis. This process is described
after the preparations of sols.

1. Dispersion methods
(a) Mechanical grinding If sulphur is finely ground using an agate pestle
and mortar, with urea (acting as an abrasive) and the mixture is stirred
with water, some of the sulphur will be fine enough to form a sol with
the water. The rest can be filtered out and the urea removed by dialysis.

glass tube

Figure 10.1. Bredig’s arc

(b) Bredig’s Arc method for metal sols In this process an arc is struck,
usually under conductivity water, between two wires of the metal
required in the sol (Figure 10.1.) The containing vessel may be cooled
in ice. For the preparation of gold sols, a trace of alkali in the water is
advantageous. The repeated striking of the arc detaches minute particles
of the metal and they pass into colloidal solution. Larger particles may
be filtered out. In aqueous sols, contamination by traces of metallic
oxide or hydroxide is common.
(c) By peptization Peptization is the dispersal of a precipitated material
into colloidal solution by the action of an electrolyte in solution. The
Colloids and the Colloidal State 107
following is an example of it. The gradual mixing of solutions of
potassium hexacyanoferrate(II) and iron(II) chloride (both 3 per cent)
precipitates Prussian blue. After a few minutes, the precipitate is
filtered off and washed well. Ethanedioic acid solution:(per 5 cent),
acting as the peptizing electrolyte, is poured through the filter-paper
and the Prussian blue passes into a sol. It should be dialysed to remove
most of the ethanedioic acid. DF ic N
~~

2. Aggregation methods
In these cases, aggregation of molecular or ionic materials occurs to the -
size of colloidal particles. Among the chemical methods used for this:
purpose are the following. E ~ F-
(a) Reduction This method is commonly used to produce metal sols. For
example, the preparation of a silver sol can be carried out as follows.
To 5cm® of a solution of silver nitrate (1 per cent), add very dilute
ammonia, drop by drop, with shaking, till the precipitated silver oxide
is just redissolved. Dilute with 95 cm® of water. Then add 0.5 cm® of
tannin solution (0.5 per cent) as reducing agent. A silver sol will be
obtained, brown in colour.
(b) Hydrolysis This process is usually applied to the production of oxide
or hydroxide sols. An example of this is the production of a sol of
hydrated iron(III) hydroxide by pouring a few cm® of a concentrated
solution of iron(II]) chloride into a large volume (say 1.5 dm®*) of boiling
distilled water. The sol is ruby-red and can be dialysed to remove most
of the electrolyte, HCl, though over-dialysis will cause precipitation of
the colloidal material
FeCl,(aq) + 3H,O(1) = Fe(OH),(sol) + 3HCl(aq)
(c) Double decomposition An arsenic(III) sulphide sol can be made by
utilizing the reaction
As,O,(aq) + 3H.S(g) — As,S,(sol) + 3H,O()

To 100 cm? of hot distilled water add 1 g of arsenic(III) oxide and


let it dissolve as fully as possible. Cool and filter the solution. Saturate
200 cm? of distilled water with hydrogen sulphide. While still passing
this gas, gradually add the arsenical solution till a bright yellow sol is
formed. Remove the electrolyte, H.S, by passing a stream of hydrogen.

Dialysis
This is the process of removing crystalloids (usually electrolytes) from sols
by the use of a colloidal membrane. The crystalloid can pass through the
membrane; the colloid cannot. The membrane most used has been
108 Physical Chemistry
parchment paper, but cellophane is now available and collodion mem-
branes are also suitable. Though it is not absolutely necessary, the mem-
brane is usually used with a dialyser; a U-shaped hollow tube of parchment
paper is also quite efficient (Figure 10.2).
A slow, continuous flow of fresh water is maintained in the outer vessel.
As a result, the concentration of crystalloid in the outer vessel continually
tends to zero and any crystalloid in the sol passes gradually through the
membrane and is washed away. Eventually the sol is left pure, though in
fact most sols require a trace of electrolyte present to stabilize them and
over-dialysis may precipitate the colloid.

Lyophobic and Lyophilic Sols


Sols are classified into lyophobic and lyophilic sols, these terms meaning
solvent-hating and solvent-loving respectively. (When water is the dispersion
medium, the terms hydrophobic and hydrophilic are used.) These sols are
distinguished by the following features:

Lyophobic sols Lyophilic sols

1. On evaporation or cooling give 1. With similar treatment give gels,


solids which do not readily form which easily form sols again by
sols again by the action of water. the action of water.
2. Are precipitated by relatively 2. Are not precipitated by relatively
small amounts ofelectrolyte. small amounts of electrolyte.
(May be salted out by large
amounts of electrolyte.)
3. Have surface tension and vis- 3. Generally have higher viscosity
cosity similar to those of the and lower surface tension than
dispersion medium. the dispersion medium.
4, The colloidal particles all migrate 4. In similar conditions, the parti-
in one direction under the in- cles may not migrate at all, or
fluence of an electrical P.D. may move in either direction.
5. The colloidal particles are easily 5. The particles are not easily
detected in the ultra-microscope. detected in the ultra-microscope.

Examples colloidal metals, Examples gums, starch, proteins.


metallic sulphides and hydroxides;
other inorganic colloids.

Properties of Sols
1. Optical properties
The smallest particles visible in the ordinary optical microscope are about
0.2 um in diameter. This is greater than the diameter of the largest
Colloids and the Colloidal State 109
-U- shaped membrane
water —=—

Ca ne
water —— dialyser

membrane
Figure 10.2. Dialysis

Figure 10.3. Ultra-microscope

colloidal particles, consequently colloidal particles cannot be directly


viewed. They can, however, be detected in the ultra-microscope. In this
instrument (Figure 10.3), the sol is brightly illuminated laterally and is
observed by a microscope in a direction at right angles to the direction of
illumination. The colloidal particles scatter the light and can be recognized
as points of light against a dark background. The particles in hydrophobic
sols are easily detected, those of hydrophilic sols much less easily and those
of true solutions not at all.

2. The Brownian movement


The Brownian movement is so called from its discovery by the botanist
Brown (1827) as a property of pollen particles in water. The colloidal
particles of a sol can be observed by the ultra-microscope, as described in
the above paragraph, and are found to be generally in a state of random
110 Physical Chemistry
motion, which is most rapid when the particles are very small and the
liquid least viscous. This motion is called the Brownian movement. In the
case of large colloidal particles, the movement may be only an oscillation.
At first, it was ascribed to convection currents in the liquid, but is now
known to be caused by collisions of the molecules of the dispersion medium
with the colloidal particles. The Brownian movement continues indefinitely
(at constant temperature) with particles of colloidal size.

3. Electrical properties
If inert electrodes are put into a lyophobic sol and a considerable potential
difference is applied across them, it is found that the colloidal particles will
migrate towards one of the electrodes, showing that the particles are
electrically charged. This phenomenon is called electrophoresis. For
particles of colloidal silver, a typical rate of flow is 3 x 10~* cm s~+ for
a potential gradient of 1V cm~+. A summary of the situation in water is:

Colloids positively charged Colloids negatively charged

Metallic hydroxides Metallic sulphides


Metals
Organic colloids

The source of the charge is believed to-be ionic. The colloidal particles
adsorb (at the surface) ions derived from electrolyte dissolved in the

Figure 10.4. Electrical double layer

dispersion medium and acquire the corresponding charge, which, in the


majority of cases, is negative by adsorption of anions. Cations are dispersed
in the neighbouring liquid to form an electrical double layer (Figure 10.4)
and this system normally remains comparatively undisturbed, moving as a
unit. Under the influence of a potential gradient, the charged colloidal
particle moves in one direction towards the oppositely charged pole and
the ionic layer moves in the other direction. For a positively charged
colloidal particle, the position is the reverse of that shown in Figure 10.4.
Colloids and the Colloidal State 111
Inorganic lyophobic sols undergo precipitation if the particles lose their
electric charge. This can happen if dialysis is carried so far that the
electrolyte concentration in the sol becomes virtually nil and none is
available for adsorption. An over-dialysed iron(III) hydroxide sol, for
example, precipitates gelatinous iron(III) hydroxide. Precipitation can
also be caused if colloidal particles are made to adsorb an ion opposite in
charge to their usual condition. For example, the particles of a silver sol,
normally negative, will adsorb Al®* ions from an added aluminium salt
until they reach neutrality (the isoelectric point) and then precipitate out. If
the isoelectric point is passed rapidly so that the colloid has no time to
coagulate, it may be stabilized with a charge opposite to its usual charge.
It is obvious, also, that coJloidal particles of opposite electrical charge
will bring about each other’s precipitation when mixed, by neutralizing
each other’s electrical charges. For example, iron(III) hydroxide sol
(positive) and arsenic(III) sulphide sol (negative) precipitate each other
when mixed and, in electrical equivalency, precipitation is complete.

4. Coagulation of colloids
Lyophobic sols can be forced to precipitate the colloid by addition of
electrolyte. It is found (as would be expected) that the colloid is pre-
cipitated by an ion of opposite charge to its own. The ion of opposite
charge is known to be adsorbed by the colloid because it can be detected
in the precipitated material.
For precipitation of a given colloid, a certain minimum concentration
of a given electrolyte is needed. The values of these minimum concentra-
tions have been closely studied. Figures for an arsenic(III) sulphide sol
(1.85 g dm-3) are given below. It is a negatively charged colloid, pre-
cipitated by cations. The cations were present as metallic chlorides in all
cases except that of Ce+, which was added as nitrate.

Minimum conc. for coagulation


Cation in 10-8 mol dm-® (as chlorides)

Lit 58
Nat 51
K+ 50

Mg2+ 0.72
Ca2t 0.65
Sr2t 0.63
Zn2+ 0.68

Als+ 0.093
Ces+ 0.080
112 Physical Chemistry
It will be observed that all the univalent cations require about the same
minimum concentrations to precipitate the colloid; so do the divalent
cations, but the concentration is much smaller; so also do the trivalent
cations, but the concentration is smaller still. That is, the effectiveness of
an ion in precipitating an oppositely charged colloid increases rapidly with
increase of the charge of the ion. A rough average shows that the minimum
concentrations required for tri-, di-, and univalent cations are in the
proportion of 1:100:700 respectively. (The situation is similar for anions
in the precipitation of positively charged colloids.) This is the reason why
alum is so effective as a styptic, i.e. in checking minor bleeding by co-
agulation of colloids in blood. The effective agent is the trivalent ion, Al**.
This property of the aluminium ion is also made use of during the treat-
ment of sewage.
Another example of the coagulation of sols by electrolytes is the
formation of deltas when material carried in colloidal form by rivers is
precipitated on contact with sea water.
Lyophilic sols are much more stable towards added electrolyte. They can
be obtained in an electrically neutral (isoelectric) state in some cases. They
are usually little affected by small concentrations of electrolytes. For this
reason, a hydrophilic sol, such as gelatine or gum-arabic, will often prevent
or delay precipitation of a hydrophobic sol by added electrolyte. This effect
is called protection.
The phenomenon called thixotropy is interesting. It occurs when the
addition of electrolyte to a sol (e.g. a gelatine sol) causes the production of
a pasty gel, which will liquefy when shaken but recover its pasty character
on standing. This phenomenon has been exploited in the recent introduction
of thixotropic paints of a pasty consistency, which spread readily when
brushed but do not spill in use.

Adsorption Indicators
Adsorption indicators are used for titrations with silver salts in which
precipitation of silver halide occurs. The use of eosin in the titration of
potassium bromide by silver nitrate solution is typical.
Silver bromide is precipitated and can, in the appropriate conditions,
adsorb Br-ion, Ag* ion or the indicator dye, eosin. So long as potassium
bromide remains in excess (i.e. the titration is incomplete), bromide ions
are adsorbed on to the precipitate, which remains pinkish-yellow. As soon
as the titration is completed and Ag* ion has gone into excess, Ag* ion and
eosin are adsorbed on to the precipitate, which turns red in colour. This
adsorption indicator is very effective with dilute solutions, say about
0.01 M. With adsorption indicators, the colour change takes place on the
precipitate, not in the solution.
11. Physical Behaviour of Gases:
The Kinetic Theory of Gases

The following laws were formulated from the results of the experimental
observation of the behaviour of gases.

Boyle’s Law
Boyle’s Law states the relation between gaseous pressure and volume in the
following way:
The volume of a given mass of gas is inversely proportional to its
pressure at constant temperature.
This is expressed mathematically in the form:
pV =k, (when T is constant)

Charles’ Law
Charles’ Law states the relation between the volume of a gas and its
temperature in the following way:
The volume of a given mass of gas is directly proportional to Its
absolute temperature at constant pressure.
This is expressed mathematically in the form:

n =k, (when p is constant)

In the above expressions, p, V and JT denote the pressure, volume, and


absolute temperature respectively of a given mass of gas and k, and k, are
constants.
The two expressions given above as the mathematical expressions of
Boyle’s Law and Charles’ Law can be combined into a single expression,
which is:
pV
va
114 Physical Chemistry
This equation can also be stated in the form:
pV =kT

If the equation refers to one mole of a gas, the constant k has a value
denoted by R and the equation is written
pV
= RT

Expressed in terms of joules, R has the value of 8.314 J K~* mol~*.


If a given mass of a particular gas has pressure, volume, and absolute
temperature, p,, V;, and T; in one set of conditions, and po, V2, and T, in
another set,
f.J PiVi _ P2V2
STP eatecks vig orl
both fractions being equal to k, which is always constant for the same mass
of the same gas. This is a very valuable equation, which enables a gaseous
volume (relating to a fixed mass of a particular gas) to be ‘corrected’ from
one set of temperature and pressure conditions to another. For reference
purposes, the internationally agreed set of standard temperature and
pressure conditions (s.t.p.) is 0°C or 273 K and 760 mmHg (101 325
N m-“).* These figures should be memorized and are in continual use in
physical science. ‘
Students are reminded that the Kelvin scale of temperature uses the
Celsius degree, but starts from a zero which is, for approximate purposes,
273 below the Celsius zero, which is the equilibrium temperature between
ice and water at 760 mmHg (101 300 N m~). Consequently, the tempera-
ture on the absolute scale corresponding to ¢t °C is (t + 273). For further
details of the absolute scale, the reader should consult a text-book of physics.

Dalton’s Law of Partial Pressures


For a mixture of gases which do not chemically react, the total pressure is
equal to the sum of their partial pressures, where the partial pressure of a
gas is the pressure the gas would exert if it occupied the volume alone.
Thus the partial pressure of a gas depends on the number of moles of that
component present and is related to the total pressure by the equation
partial pressure of a component = total pressure x mole fraction,
where
mole trattick ie number of moles of that component present
total number of moles of gas present
*The unit for pressure, N m2, is sometimes called a Pascal (abbreviation Pa), and
1000 N m-2 a kiloPascal (kPa).
Physical Behaviour of Gases: The Kinetic Theory of Gases 115
~Graham’s Law
It is a common observation that gases, unlike solids and liquids, tend to
fill any container into which they are placed. Also, a low density gas will
pass more rapidly through a porous material than a high density gas. Thus,
in the apparatus represented by Figure 11.1 the levels of liquid in the
U-tubes will move in the directions indicated. The movement of gases is
called diffusion and it is obvious from experiments such as those illustrated
in Figure 11.1 that the rate of diffusion of a gas is inversely related to its

co,
Sar ee porous pot

porous pot

coloured liquid e.g.


KMn0O,, solution
Figure 11.1.

density. In fact, Graham found from experimental results that the rate of
diffusion of a gas is inversely proportional to the square root of its density.

Kinetic Theory of Gases


The above experimental laws can be accounted for qualitatively by
considering a gas as being made up of widely spaced molecules which are
free to move in any direction. If this picture of a gas is treated quantitative-
ly, it is possible to derive theoretically the same gas laws. This agreement
between theory and experimental results is good evidence for the validity
of the theory.
In the mathematical treatment of the kinetic theory it is necessary to
make several assumptions, in particular:

1. the individual molecules of a gas have negligible volume,


2. the attractive forces between the molecules are negligible,
3. the average kinetic energy of the molecules of a gas is proportional to
the absolute temperature on the Kelvin scale.
116 Physical Chemistry
Consider n molecules, each of mass m, in a cubical box of side /. The
velocity, c,, of any one molecule can be resolved into three components,
X1, Yi, and Z,, along the directions of the edges of the cube.
afitste 2 2
ef SHxP FYE tT 2
Considering the x component only, a molecule travelling with this velocity
will experience a change of momentum of 2mx, on colliding with one wall
of the box.
The molecule would collide x,// times per unit time. Therefore the
change of momentum per unit time, or the rate of change of momentum,
for this component is

2m, x 2 me2

Similarly the rates of change of momentum for the other components


are
2my?/1 and 2mz?//
Thus the total change of momentum per unit time for one molecule is
2mx? e 2my? ei2mz? _ 2mc?
l l l l
For n molecules the total change in momentum is e
2 2
= (B+ cf + ch... +2)

This can be simplified to 27né?/I, where ¢ is the root mean square velocity.
The rate of change of momentum is equal to the force exerted by the
collisions. This force is acting on a total area of 6/? and, as pressure is equal
to force per unit area,

But /® is equal to the volume of the cube, and therefore


pV = 4 mnée??
This fundamental gas equation can be used to derive the gas laws.

Boyle’s Law
As stated on page 115, the kinetic energy of a gas is proportional to the
temperature on the Kelvin scale and so, for a fixed mass of gas at a constant
Physical Behaviour of Gases: The Kinetic Theory of Gases 117
temperature, the kinetic energy, }mné?, is a constant. Rearranging the
fundamental gas equation,
pV = & X 3 mné?
therefore pV = constant
which is Boyle’s Law.

Charles’ Law
Re-arranging the fundamental gas equation,
1
V=— xX 3x 4 mn?

At constant pressure,

V=k x 3mnéc??
and the kinetic energy is proportional to the absolute temperature, 7,
therefore
V=kT
which is Charles’ Law.

Dalton’s Law
Consider a mixture of two gases in the same volume, V. Then for each gas,
PAV =%32 X dm,n,C2
DoV = %
2
X kmon2¢7

In the mixture of the gases the total kinetic energy is equal to the sum of
the individual kinetic energies. Then, where p is the total pressure,
pV = 2(4m,nyc? i b/MgM23)

pV= 3(3piV+ BpeV)


P=Pi + Po

which is a statement of Dalton’s Law.

Graham’s Law
Rearranging the fundamental gas equation,

C= 3pV
mn
118 Physical Chemistry
but mn/V is equal to the density of the gas. Therefore
pe tedinat)
d
ee Bis
Son ad
At a constant pressure,
rclge ¢ m7

and as the rate of diffusion is directly proportional to the velocity of the


molecules,
Ee, 1
rate of diffusion « 7d

which is Graham’s Law.

Molecular Velocities
By substituting values for the density and pressure, using appropriate units,
into the equation
= f3p
d
it is possible to calculate the root mean square velocity of the molecules of
a gas. : ‘

Distribution of molecular velocities


Although the total kinetic energy of a gas is constant for a constant
temperature, not all the molecules of the gas travel with the same velocity.

of
number
molecules

velocity of molecules
Figure 11.2. Distribution of molecular velocities
Physical Behaviour of Gases: The Kinetic Theory of Gases 119
The distribution of molecular velocities was determined theoretically by
Maxwell and Boltzman and it is represented graphically in Figure 11.2.
At a higher temperature, T>, the shape of the distribution curve will be
similar, i.e. the majority of the molecules will have velocities near to the
mean, but the mean velocity will be higher. The importance of the variation
of the distribution with temperature is discussed in the chapter on reaction
kinetics, page 261.

Variation of the Behaviour of Gases from Boyle’s Law


It has been shown on page 117 that an ideal gas, which obeys Boyle’s Law
exactly, would fulfil the requirement
pV =k (when T is constant)
In fact, no actual gas does so. If a graph of the product pV at 273 K
is drawn against p (as in Figure 11.3), it is found that, for most gases pV is

at
pV
K273

(é) 500
increasing pressure (atmospheres)
Figure 11.3. Variation of value of p¥ with pressure

lower in value than Boyle’s Law requires until very high pressures are
reached (of the order of 250 atmospheres or more). Hydrogen and the
noble gases, however, show values of pV which are too high for the
requirements of Boyle’s Law throughout the whole range of pressures from
one atmosphere upwards. The variations from Boyle’s Law are greatest in
gases which, like carbon dioxide, are most easily liquefied.
120 Physical Chemistry
These variations arise from two assumptions which are inherent in the
formulation of Boyle’s Law. These assumptions are:
1. That the gaseous molecules occupy a volume which is completely
negligible compared with the space occupied by the gas. This assump-
tion is correct only at very low pressures.
2. That there are no attractive forces between the gaseous molecules.
The best-known attempt to deal with these variations from ideal
gaseous behaviour, and to express them mathematically, is that of van der
Waals.
To allow for the reality of the volume occupied by the gaseous molecules,
van der Waals introduced a correction to the volume, V, which became
(V — b), where 6 is a constant appropriate to each gas and related to the
volume occupied by the molecules.
The attractive forces normally exercised between the molecules of gases
cancel each other when exerted on a molecule in the interior of the gas,
because they are exercised at random. In the case of a molecule near the
containing wall of the vessel, however, the resultant attractive force must
be inwards, since there are no gaseous molecules on the outside. Con-
sequently, the pressure exerted on the containing wall by such a molecule is
lessened by this attractive force of the other molecules from inside. The
attractive force is proportional to the number of molecules in the gas, i.e.
to its density. Further, the number of molecules of gas striking the con-
taining wall at a given moment is also proportional to the density of the
gas, so that the total inward attracting force, representing loss of pressure,
is proportional to the square of the density of the gas. This is inversely
proportional to the square of the volume of the gas so that the Joss of
pressure resulting from molecular attraction can be represented by a
quantity a/V*, where a is a constant. This quantity must be added to the
observed pressure of the gas to represent its ‘true’ pressure.
Allowing for these two corrections, van der Waals’ equation takes the form

(> + ju) — = Rr
and, in this form, relates to one mole of the gas.
Each gas has a particular value of its own for constant, a, and another
value for constant, b. The constant, a, was at first thought to be indepen-
dent of temperature but, in fact, is not so.
Since gases do not obey Boyle’s Law exactly, they also show small
variations from other laws and generalizations relating to gases.

The Work of Andrews on Carbon Dioxide


Some experiments of Andrews (1861) on carbon dioxide have come to be
regarded as classical. They are closely related to the theoretical work of
van der Waals.
Physical Behaviour of Gases: The Kinetic Theory of Gases 121
Andrews investigated the effect of changes of temperature and pressure
on the volume of carbon dioxide.
The experimental results are expressed by isothermals as in the approxi-
mate graph (Figure 11.4). Each isothermal refers to the same mass of
carbon dioxide and shows the variation of the volume of carbon dioxide
with pressure for a particular temperature. The isotherm of 321 K, MN,
approximates to a rectangular hyperbola, which is the correct plot of the

~a

(atmospheres)
pressure

volume

Figure 11.4. Isothermals of carbon dioxide

equation, pV = k, which is the expression of Boyle’s Law. In this isotherm,


carbon dioxide is showing the ‘normal’ behaviour of a gas.
The isotherm of 286K, ABCD, is markedly different. Its course is
interpreted in the following way. The portion AB represents the usual
contraction of the volume of a gas as pressure increases, and approximates
to a hyperbola. The portion BC represents a large contraction in volume
for almost no pressure change. The portion CD represents almost no
volume change for a very large pressure change. Thus at B the pressure has
become large enough to start the Jiquefaction of carbon dioxide at 286 K,
and from B to C this liquefaction continues and completes itself with only
slight increase of pressure. At C liquefaction is completed and, liquids
being almost incompressible, further increase of pressure produces no
appreciable volume change. The isotherm of 295 K shows similar effects,
122 Physical Chemistry
with EF corresponding to AB, FG (a shorter stretch at the higher tempera-
ture) representing the liquefaction stage, and GH the behaviour of liquid
carbon dioxide as almost incompressible.
The isotherm of 304.1 K is very significant. It shows only a slight
inflection (at L). This corresponds to the point at which the liquefaction
stage of BC and FG in the lower isothermals has decreased so that, in
effect, the points C,B or G,F have come to coincide. This means that the
liquefaction stage just barely occurs, and above this temperature (as in
MN) does not occur at all; that is, 304.1 K represents a temperature above
which carbon dioxide cannot be liquefied no matter what pressure may be
applied to it. This temperature is known as the critical temperature of
carbon dioxide.
This situation is quite general and every gas or vapour has its own
critical temperature, which is defined as
that temperature above which no amount of pressure will serve to
liquefy the material.
The least pressure which suffices to liquefy the material at its critical
temperature is called its critical pressure. The volume occupied by one mole
of the material at the critical temperature and pressure is called the
critical volume. The values of critical temperature (to the nearest degree)
and critical pressure are shown below for some common materials:
——. $< $< sw
Critical Critical
temperature (K) pressure (atmospheres)

Hydrogen 38 20.0
Nitrogen 127 33.0
Oxygen 155 50.0
Carbon dioxide 304 73.0
Chlorine 417 76.1
Sulphur dioxide 430 77.6

It will be noticed that the last three gases tabulated have critical
temperatures above room temperature and so can be liquefied at room
temperature by the exercise of pressure alone. These gases were among the
earliest to be liquefied. The first three gases, however, have very low
critical temperatures. These could not be attained in the early days of
experimentation on liquefaction; consequently these gases (and others such
as carbon monoxide and methane) were regarded as permanent gases.
There is, in fact, no truly permanent gas. All known gases have now been
liquefied by cooling each below its critical temperature and applying the
necessary pressure.
Physical Behaviour of Gases: The Kinetic Theory of Gases 123
Liquefaction of Gases
The work of Andrews on carbon dioxide showed clearly that to liquefy any
gas it is necessary to cool it below its critical temperature and then to
apply enough pressure to induce liquefaction. The amount of pressure
required is at its maximum at the critical temperature and falls as the
temperature of the gas is reduced progressively below the critical tempera-
ture. This gives the following general methods of liquefaction of gases.

1. By cooling only
If room temperature is well below the critical temperature of a gas, a
moderate amount more of cooling may reduce the temperature to the
point at which the gas will liquefy under one atmosphere pressure, i.e. the
gas can be liquefied merely by cooling. For example, nitrogen dioxide and
sulphur dioxide can be liquefied by passage through a tube immersed in a
freezing-mixture of ice-salt, giving about 261 K. The gas must be dry to
prevent formation of ice.

2. By the joint application of cooling and pressure


Though others also worked on this method of liquefaction, the classical
work in this field is associated with Faraday. His work on the liquefaction
of chlorine is typical. He used an inverted, V-shaped glass tube, which was
sealed, with one end containing chlorine hydrate, Cl,..8H,O. The empty
end was cooled in a freezing-mixture and the chlorine hydrate was
warmed. It liberated a relatively large volume of chlorine, which generated
high pressure. Under the influence of this high pressure and the cooling
effect of the freezing-mixture, drops of liquid chlorine were obtained.
Chlorine is now liquefied by pressure alone and stored dry in steel
containers.

3. By using the Joule-Thomson effect. Liquefaction of air


If a highly compressed gas is allowed to expand into a region of low
pressure, a cooling effect is observed. This is called the Joule-Thomson
effect. It arises because as the molecules separate during expansion,
internal work is done in overcoming the attractive forces between them.
A perfect gas, with no attractive forces between the molecules, would show
no Joule-Thomson effect.
The Joule-Thomson effect is more marked at lower temperatures and
was used in the Linde process for the liquefaction of air. In this process, air
is purified by passage over soda-lime to remove carbon dioxide and is then
dried. It is pumped to about 200 atmospheres pressure, cooled back to
ordinary temperature, and passed through closely spiralled thin copper
tubing surrounded by thicker copper tubing (Figure 11.5). At the valve, A,
124 Physical Chemistry

cold air_—=[ |
at 200 atmospheres
pressure

insulation

Figure 11.5. Liquefaction of air

the air expands to about 20 atmospheres pressure and is cooled by the


Joule-Thomson effect. The cooled air passes out to the pumps in
the wider copper tube, cooling the incoming, highly compressed air in the
inner tube. This, in turn, is further cooled by the Joule-Thomson effect and
still further cools incoming air as it passes out. Eventually, this cascade
effect so reduces the temperature of the air that, at 20 atmospheres pressure,
it liquefies. All the tubing is protected from heat loss by insulating material.
The process was later improved by Claude by making the expanding gas
do work, and so lose heat energy, in an expansion engine. The engine
drives a dynamo, which recovers some of the energy used in compressing
the gas and also makes the liquefaction process more rapid.
Hydrogen is anomalous with regard to the Joule-Thomson effect,
showing a rise of temperature on expansion at room temperature. This
continues down to 193 K; then, at this point (the inversion temperature),
hydrogen becomes normal and cools under expansion. Hydrogen was
liquefied (1898) by Dewar, who cooled the gas below the inversion
temperature by liquid air and then used the principle of the Linde pro-
cess as described above. Helium is anomalous like hydrogen, with a
very low inversion temperature (33 K).
Physical Behaviour of Gases: The Kinetic Theory of Gases Hs
Densities of Gases and Relative Molecular Masses
Cannizzaro’s use of the densities of gases relative to the density of hydrogen
for the determination of relative molecular masses and ultimately relative
atomic masses was discussed on page 10. It is possible to use the ideal gas
equation for the calculation of the relative molecular mass of a gas from its
density.
For example, it was found that 100 cm® ofbutane at 298 K anda pressure
of 102700 N m~? had a mass of 0.233 g.
The ideal gas equation is
pV = RT
where p is the pressure in N m~?,
V is the volume in m® of 1 mole of gas,
T is the temperature on the Kelvin scale,
and_ R the gas constant, is equal to 8.31 J K~? mol7?.
Re-arranging the equation, the volume occupied by one mole is given by

ye
Dp

Under the stated conditions of temperature and pressure,


_ 8.31 x 298
~ 102700
= 2.41 x 10-7 m?
But 100 cm® of butane has a mass of 0.233 g. Therefore 1 m*® has a mass
of
U.2a9 le
= 2.33 x 10°g
10?
Therefore 2.41 x 10-2 m® will have a mass of
2,30 % 10° < 2.41 x 107? = 56.15g
Thus this density measurement gives a value of 56 for the relative molecular
mass of butane.

Experimental Methods for Determining the Densities of Gases


1. Direct weighing
The density of a substance which is a gas at room temperature can be
found with sufficient accuracy by direct weighing. The mass of a container
of known volume, such as a glass syringe or an air-tight plastic bottle, is
determined. Then, using a predetermined value for the density of air, the
126 Physical Chemistry
mass of air in the container is calculated and subtracted from the first
weighing to give the mass of the container. The container is then filled with
the gas under investigation and reweighed. It is thus possible to find the
mass of a known volume of gas.
The densities of the vapours of volatile liquids can also be used for
calculating their relative molecular masses. The densities of such com-
pounds can be determined by the traditional nineteenth century methods
or their modern equivalents.

2. Victor Meyer’s method for finding the vapour density of a volatile liquid
This method is used chiefly in connection with liquids such as diethyl
ether (ethoxyethane), trichloromethane, ethanol, and carbon disulphide,

loose-fitting cork ———

glass inner tube ‘water

copper or glass outer tube

cotton wool

Figure 11.6. Victor Meyer’s apparatus

all of which have relatively low boiling points. The apparatus is shown in
Figure 11.6. The experiment will be described for the special case of ether.
A small stoppered tube is weighed empty, and then weighed again (check-
ing immediately before use in the apparatus) full of ether. The mass of
ether contained is usually about 0.1 g.
The water below C in the outer tube is heated, with the tube ED not yet
in position. Air bubbles will be seen escaping from the narrow tube at D
Physical Behaviour of Gases: The Kinetic Theory of Gases 127
as the inner tube warms up, but eventually steam will begin to escape from
the loose cork below B. The flame heating the water is adjusted to maintain
a slight but steady escape of steam near B. Shortly after this, escape of
bubbles at D will cease, showing that the inner tube from B to C is ina
steady state with most of its length at about 100°C. The graduated tube
ED is placed in position as shown, but full of water.
The small weighed tube of ether is then inserted at A and held at B
by the glass rod, G, until stopper A is replaced. The rod G is pulled gently
aside so that the tube of ether falls to C where cotton-wool prevents
breakage. The ether vaporizes, forcing out the stopper of the small tube,
and air is expelled rapidly at D and collected (as shown) in the graduated
tube. When no more air is expelled, the tube ED is moved clear of the
narrow tube at D and lowered into the water till E and F are at the same
level, i.e. the air in the tube is at atmospheric pressure. The volume of the
air is read off; the temperature of the water in the collecting vessel is noted
and the barometer is read. The vapour pressure of water at the relevant
temperature is also required (from tables).
The following points are important:
(a) The temperature in the tube BC need not be known, but should remain
steady during the determination and should be high enough to vaporize
the experimental liquid easily, i.e. should be at least 20 K above the
boiling point of that liquid.
(b) The vapour of the experimental liquid should remain near C and, in
particular, must not reach B where it could liquefy. The wide bulb near
C, and the length of narrower tubing below B, through which diffusion
is slow, are designed to secure this. The air expelled from the tube by
the experimental liquid then represents the volume which that liquid
would occupy if it could exist as vapour under room conditions.
(c) The advantages of Meyer’s method are its rapidity and the small
amount of liquid required. It is not very accurate, but this is a minor
matter. For a consideration of this point, see after the calculation
below.

Calculation
Mass of stoppered tube = 0.672 g
Mass of stoppered tube and ether = 0.783 g
Mass of ether = 0.111 g
Volume of air displaced = 36.0 cm
Temperature = 285K
Pressure = 100800 N m-? (756 mmHg)
Vapour pressure of water at 285K = 1467 N m~? (11 mmHg)
True air pressure = 100 800 — 1467 = 99333 N m~*
128 Physical Chemistry
: 99333 _ 273
Volume of air corrected to s.t.p. = 36.0 X {01300 x 585

= 33.8 cm®
: 0.09
Mass of this volume of hydrogen = 33.8 X Totes 0.003 04 g

(since 1000 cm® of hydrogen at s.t.p. has a mass of 0.09 g).


Mass of ether
Relative vapour density of ether
~ Mass
ofhydrogen
Or aes
~ 0.00304
Relative molecular mass of ether = 36.5 x 2 = 73.0
Alternatively the calculation may be based on the ideal gas equation in
a similar manner to the example on page 125.
The true molecular mass of ether is 74. The method of Meyer is not very
accurate but this hardly matters, for the following reason. If ether were a
new material, analysis would show its molecular formula to be C,H,9O or
some multiple of this, i.e. its molecular mass must be 74 or 2 X 74 0r3 X
74, etc. The method of Meyer is needed merely to determine which of
these numbers, 74, 148, 222, etc., is the correct multiple. Moderate
inaccuracy hardly matters in such a case. :

3. Dumas’s method for finding the relative molecular mass of a volatile liquid
This method will be described for the special case of ether. A Dumas bulb
(Figure 11.7) is weighed full of air at room temperature and pressure,
which must be known. The bulb is then warmed and allowed to cool with
its jet, A, under ether. As the bulb cools, ether will be drawn into it and
should fill about one-quarter of the bulb. The bulb is then placed in position
as shown in Figure 11.7, and held so that the greatest possible fraction of
the bulb is immersed in the water. The water is then heated. In a short time,
the ether begins to vaporize and escape from A, where it can be burnt. As
it does so, air is swept from the bulb. Eventually the water boils, by which
time the flame at A will be very small. The water is kept boiling for five
minutes or so to secure constant temperature and the jet A is sealed off by
heating with a burner while the water is still boiling. In a successful
experiment this gives a bulb full of ether vapour at the boiling point of
water, which is read from the thermometer, and barometric pressure,
which is also read. The bulb is dried and weighed when cool.
A file scratch is gently made near the sealed end, A, of the bulb and the
end is broken off under the surface of cold, boiled-out water. This water
should rush in and fill the bulb. Boiled-out water is used because it contains
no air, bubbles of which could prevent the bulb from filling properly with
Physical Behaviour of Gases: The Kinetic Theory of Gases 129

Figure 11.7. Dumas’s apparatus

water. The bulb is dried and weighed full of water, including the broken
end of the jet.

Calculation
Mass of bulb full of air = 78.385 g
Temperature of room == 287K
Atmospheric pressure = 102 900 N m~? (772 mmHg)
Temperature at which bulb was
sealed = 373 K
Mass of bulb and ether == 78.16)
Mass of bulb full of water #2/355.7 2
Capacity of bulb = (355.7 — 78.4) = 277.3 cm®
This capacity referred to s.t.p. is
102900 _ 273 = 3

Mass of air in bulb = 1,293 x 268


1000
= 0.347 g
(since 1000 cm? of air at s.t.p. has a mass of 1.293 g).
130 Physical Chemistry

Mass of evacuated bulb = (78.385 — 0.347)


= 78.038 g
Mass of ether vapour sealed in = (78.765 — 78.038)
= 0,727.2
273 _. 102 900
Volume of ether vapour at s.t.p. 2713. x 373 x 101 300

= 206m;
: 206
Mass of this volume of hydrogen = 0.09 x T000

= 0.0185 g
(since 1000 cm® of hydrogen at s.t.p. has a mass of 0.09 g).

Relative vapour density of ether = Sas

==)39)3
Relative molecular mass of ether = 78.6
A more rigorous method of calculation would regard the first weighing
as the mass of the bulb alone and would consider that the mass of the bulb
full of vapour was too small by an amount equal to the mass of the air
displaced by the sealed bulb. The final answer will be the same whichever
method is employed.
A slight error is involved in taking the capacity of the bulb to be the same
at 373 K as at room temperature. The expansion of the glass is ignored.
Dumas’s method is capable of greater accuracy than Meyer’s, but, in a
general way, the value obtained need only be accurate enough to distinguish
between multiples of an empirical molecular mass as was indicated at the
end of the account of Meyer’s method. The principal disadvantages of
Dumas’s method are:
(a) It needs a large quantity of material.
(b) If the material contains any impurity of higher boiling point than itself,
this impurity accumulates in the final vapour and makes the vapour
density inaccurate.
(c) It is difficult to expel the whole of the air. A small bubble is usually left
after filling with water.
In the modern equivalent of the last two methods a known mass of the
volatile liquid is injected through a serum cap into a glass syringe enclosed
in a transparent steam bath. The volume of vapour produced can be read
directly from the syringe. Thus the volume of a known mass of vapour at
the temperature of the steam bath and at atmospheric pressure is known.
S ~ *

12. Phase Equilibria

A phase is a physically distinct part of a system. Thus examples of two-


phase systems are those containing: :
a gas and a solid,
a gas and a liquid,
a solid and a liquid,
a solid and a solid,
two immiscible liquids.
Examples of three-phase systems are:
a solid, a liquid and a gas,
two immiscible liquids and a gas.
Phase equilibria involves a study of the conditions (temperature,
pressure, and concentration) under which different phases are in equi-
librium.
The number of components in a system is the minimum number of
chemical entities required to make up the system. A liquid and its vapour
constitute a two-phase, one-component system, whereas a solution of a
solute and the solvent vapour is a two-phase, two-component system.

One-component Systems
Concentration is not a variable when there is only one component in a
system and the phase equilibria are only affected by changes in temperature
and pressure.
A liquid in a closed container will reach a state of equilibrium with its
vapour. This state is realized when the rate of movement of molecules from
liquid to vapour is exactly balanced by the rate of movement from vapour
to liquid. At equilibrium the concentration of molecules in the vapour
phase will be constant which will result in the liquid exerting a constant
vapour pressure for that temperature (saturated vapour pressure). The
saturated vapour pressure and the temperature represent conditions for
equilibrium between the liquid phase and the vapour phase.
An increase in temperature causes an increase in the rate of movement of
molecules from liquid to vapour phase and a new equilibrium is set up at a
higher vapour pressure.
132 Physical Chemistry
A graph representing the equilibrium conditions, a phase diagram, over
a range of temperatures and pressures is given in Figure 12.1.

liquid

vapour

pressure
vapour

temperature
Figure 12.1.

Also for a one-component system it is possible to determine the con-


ditions for equilibrium between solid and liquid and at low pressures
between solid and vapour. The complete phase diagram for water is given
in Figure 12.2. '

pressure
vapour

temperature
Figure 12.2.

Two-component Systems
In the case of two-component systems all three variables, temperature,
pressure, and concentration, may affect the equilibria. It is customary to
Phase Equilibria 1o3
keep one of these conditions constant and to investigate the effects of
variations in the other two.
Six general types of two-component systems are considered in this
chapter.

1. Dilute solutions of non-volatile solutes


(a) Lowering of vapour pressure

The presence of a non-volatile solute reduces the rate of movement of


molecules from the liquid phase to the vapour phase and so lowers the
saturated vapour pressure. This phenomenon was investigated by Raoult,
who, keeping the temperature constant, found that the lowering of vapour
pressure was related to the concentration of the solution by the equation

Paci Pate Beal


Pi N+a

where p; = vapour pressure of pure solvent,


P2 = vapour pressure of solution,
n = number of moles of solute,
N = number of moles of solvent.

The left-hand side of the equation is the relative lowering of the vapour
pressure and the right-hand side is the mole fraction of solute. This
relationship, which is one form of Raoult’s law, only holds for dilute
solutions of non-volatile non-electrolytes.
If the solution is very dilute, 1 is small relative to N and the law can then
be used in the approximate form

Py Ppa
P1 als

The following calculation illustrates the application of the law.

Calculation The vapour pressure of ether (molecular mass 74) is


58 930 Nm=~2 (442 mmHg) at 293 K. If 3g of a compound A are dis-
solved in 50 g of ether at this temperature, the vapour pressure falls to
56800 N m~2 (426 mmHg). Calculate the molecular mass of A.
Using the same symbols as above,

le
134 Physical Chemistry
where M is the molecular mass of A.
50
N = =; p, = 58 930;pe = 56 800
74
Using the complete formula:
58 930 — 56 800 _ 3/M
58 930 50/74 + 3/M
From this, by the usual algebraic methods, M = 118.
The approximate form of the equation gives:
58930 — 56800 _ 3/M
58930 50/74
M = 123
Direct measurement of vapour pressure change is rather difficult. The
following gravimetric method can be applied to water and its solutions.
A current of dried air is passed slowly through several bulbs containing a
solution of known concentration of the compound of which the molecular
mass is required. The air becomes saturated to the vapour pressure, Po.
The air then passes on through a set of weighed bulbs containing pure
water. Here the air is saturated to the vapour pressure, p,. The bulbs are
weighed after the experiment and the Joss of mass is proportional to
(pi — Pz). The air then passes through weighed calcium chloride tubes,
which are weighed again after the experiment. The gain in mass of these
tubes is proportional to p,.
If the /oss of mass in the bulbs of water is m, and the gain in the calcium
chloride tubes is mg,

and the calculation is on the same lines as before.


When one particular solution is considered the concentration will
remain constant and the variation of the vapour pressure of the solution
with temperature can be investigated. The vapour pressure curve of a
solution will be similar in shape to but lower than that of the pure solvent.
Figure 12.3 shows vapour pressure curves of the pure solvent and two
solutions. The concentration of solution 2 is greater than that of solution 1.

(b) Elevation of boiling point


The third variable, vapour pressure, is most easily kept constant by
considering the temperatures required to raise the vapour pressures of
solutions to the pressure of the atmosphere at which points the solutions
will boil. It is clear from Figure 12.3 that the presence of a non-volatile
Phase Equilibria 135

atmospheric pressure

pressure
vapour solvent
solution 1

solution 2 |
| |
| |
os
2 Se ae
temperature tt t

Figure 12.3.

solute raises the boiling point of a solvent. The boiling points of the pure
solvent, solution 1, and solution 2 are t¢, t,, and f., respectively. The
magnitude of the elevation is related to the lowering of the vapour pressure
which in turn, according to Raoult’s Law, is related to the mole fraction of
solute. The elevation of boiling point must therefore be dependent on the
mole fraction of solute. In fact, it is found that a constant mole fraction of
various solutes in one particular solvent always produces the same
elevation. If the mole fraction used is 1 mole of solute in 1000 g of solvent
the elevation produced is known as the boiling-point constant, K. The
elevation of boiling point provides a method for the determination of
relative masses of solutes, the limitations being that the solutes must be
non-volatile and non-electrolytes and that only dilute solutions are used in
the experimental work.

Determination of relative mass by boiling-point elevation


Several methods are available for this purpose. The one described below
was introduced by Landsberger and modified by Walker and Lumsden.
The apparatus is shown in Figure 12.4. A quantity of the pure solvent is
placed in the inner tube and the apparatus is completed as shown. Some
more of the pure solvent is boiled in a flask and the vapour is passed in as
shown until the thermometer registers a constant temperature, which is read
(with a lens) as the boiling point of the pure solvent. (Notice that the inlet
tube for vapour ends in a small perforated bulb to break up the vapour
into small bubbles.)
136 Physical Chemistry
thermometer reading to 0.01 K
—— vapour of solvent

hole for escape of vapour

graduated inner tube

vapour jacket to prevent


irregular cooling

—- to condenser

Figure 12.4. Elevation of boiling point

A known mass of the solute (of which the relative molecular mass is
required) is then dissolved in the solvent and a second (higher) boiling
point of the solution is observed as before. At once, to avoid further
dilution, the supply of vapour is cut off. The fittings are removed from the
inner tube and the volume of solution is read off from the graduations. If
desired, more solvent vapour can then be passed, which will dilute the
solution, and other similar observations of boiling point and volume of
liquid can be made.
The following is a typical set of results:

Boiling point of trichloromethane = 332.80 K


Mass of camphor added = 0.400¢
Boiling point of the solution = 333.10K
Volume of solution = 23.0 cm
(Density of trichloromethane is 1.50 g cm~*. K is
3.9 K per 1000 g of trichloromethane.)

The density of the solution can be taken as the same as that of trichloro-
methane, so that the mass of trichloromethane used is (1.5 x 23.0 g) =
34.5 g. The boiling-point elevation is (333.10 — 332.80) K = 0.30 K. From
the results:
Phase Equilibria 137
0.30K is the elevation in 34.5 g of trichloromethane by 0.40g of
camphor.
3.9 K is the elevation in 1000 g of trichloromethane by
3.9 _ 1000
0.40; x —_
030 x 534.5 g == 151 g of camphor

That is, the relative molecular mass of camphor is 151.


If desired, this kind of calculation can be generalized into the following
formula. Let x K be the elevation of boiling point produced by dissolving
a g of a solute A in b g of a solvent B. Then, if y K is the boiling-point
constant per 1000 g of B,

M=ax%
x ——
aR
where & is the relative molecular mass of the solute, A.

(c) Depression of freezing point


The vapour pressure curves of solutions 1 and 2 cross the line which
represents the transition to solid (Figure 12.5) at lower and lower tempera-
tures, indicating that a non-volatile solute depresses the freezing point of a
solvent.
solvent
solution 1
solution 2

vapour
pressure

t, tt temperature
Figure 12.5.

In Figure 12.5, t, t; and ft, are the freezing points of the pure solvent,
solution 1, and solution 2 respectively. It is clear from the graph that the
magnitude of the depression is related to the lowering of the vapour
pressure, which, according to Raoult’s law, is related to the mole fraction
138 Physical Chemistry
of solute. The laws which govern depression of freezing point are identical
to those for elevation of boiling point, that is, a constant mole fraction of
any non-volatile non-electrolyte, when dissolved in the same solvent, will
produce identical depressions. When the mole fraction is 1 mole of solute
dissolved in 1000 g of solvent, the depression is known as the freezing-
point constant, for that solvent. Freezing-point depressions can also be
used for relative molecular mass determinations.

Determination of molecular mass by freezing-point depression


Apparatus suitable for this determination is shown in Figure 12.6. The
freezing mixture is not necessarily ice or ice-salt. It is a material which can

Beckmann thermometer
reading to 0.001 K
— stirrer A

stirrer B

air-gap to reduce
the rate of cooling

freezing mixture

Figure 12.6. Depression of freezing point

maintain a temperature about 7 K below the freezing point of the solvent


used in the inner tube. If this solvent is water, ice-salt is suitable, but, in
some cases, the freezing mixture may be warm water. The choice is
relative to the freezing point to be observed.
In making the determination, a known mass of the solvent is placed in
the inner tube. Stirrer B is used occasionally to keep the freezing mixture
reasonably uniform. The solvent is stirred vigorously by stirrer A and
temperature observations are made every quarter-minute. From them a
graph can be drawn which will usually take the form shown in Figure 12.7.
Phase Equilibria 139

super-cooling

temperature


freezing point
of pure solvent
freezing- point
depression

freezing point
of solution

time

Figure 12.7. Cooling curves

There is usually some degree of super-cooling. If this occurs, the tempera-


ture will rise as the solvent starts to freeze and will become steady at the
freezing point, as shown in Figure 12.7. The solvent is then allowed to
warm up and melt, and a known mass of the solute (preferably in pellet
form) is introduced by the side-tube, C. (If it is more convenient, a solution
of known concentration can be substituted for the pure solvent.) When the
solute has dissolved, a cooling curve is plotted from observations as before.
Super-cooling will occur again (see Figure 12.7) and should not exceed
0.5 K. If it does, so much solid solvent will separate rapidly when freezing
begins that the concentration of the solution will be appreciably altered.
Super-cooling can usually be kept at a tolerable level by vigorous stirring,
but in some cases ‘seeding’ by a small crystal of the solid solvent may be
required to induce freezing. The solution freezes at a lower temperature
_ than the pure solvent and the freezing-point depression is read off as
shown on the graph. Other observations can be made by addition of
further amounts of the solute.
The following are typical results:

Volume of benzene used = 60.0 cm


Freezing point of benzene = 278.495 K
Mass of naphthalene dissolved = 1.2508
Freezing point of the solution = 277.515 K
(Density of benzene = 0.880 gcm~°
K = 5.1K per 1000 g of benzene.)
Mass of benzene used = 60 x 0.880 g
= 52.82
140 Physical Chemistry
From the results:
0.980 K is the depression in 52.8 g of benzene by 1.250 g of naphthalene.
That is, 5.1 K is the depression in 1000 g of benzene by

1.250x oe x= dea = 123 g of naphthalene


0.980 ~~52.8
From this, the relative molecular mass of naphthalene is 123.
This type of calculation can be generalized in the following way. Let
x K be the depression of freezing point produced by dissolving a g of a
solute A in b g of a solvent B. Then, if y is the freezing-point constant per
1000 g of B,
Mi=saxX y ,, 1000b
where M is the relative molecular mass of the solute, A.
The freezing-point method for relative molecular mass determination is
probably more accurate than the boiling-point method, but in practice
neither really needs to show a high degree of accuracy for the following
reason. Consider the above result for naphthalene. Analysis shows that
the composition of this hydrocarbon corresponds to the simplest formula,
C;H,. The molecular formula of naphthalene must then be (C;H,), and
the relative molecular mass 64 x 1 or 64 x 2 or 64 x 3, etc. The method
for determination of relative molecular mass need only be accurate
enough to distinguish among these multiples, and the occurrence of a few
per cent of error is unimportant. The figure 123 in the calculation above
clearly indicates the multiple 128 and the formula C,)Hg.
Freezing-point and boiling-point constants for some common solvents
are given below:

K per 1000 g of solvent


Solvent Boiling point Freezing point
(K) (K)
Water 0.52 1.86
Benzene 2.60 5.10
Camphor _— 40.0
Tribromomethane — 14.3

Method of Rast for determination of relative molecular mass by freezing-


point depression
This method utilizes the exceptionally high freezing-point constant of
camphor, which results in a large depression for a small mole ratio of
Phase Equilibria 141
solute to solvent and hence there is no need to use a very accurate ther-
mometer.
A known mass (probably a few mg) of the solute is melted with a known
mass (10-12 times its own mass) of camphor. The product is cooled and
ground to powder. Its melting point is determined in a small capillary tube
attached to a thermometer bulb and heated in propane-1,2,3,-triol (glycerol)
as in the usual method of finding melting points of organic compounds.
The melting point of the pure camphor having been similarly determined,
the depression is known and the calculation can be made as before.

The Beckmann thermometer

subsidiary bulb

scale reads about 8 K


by 0.01 K graduations

main bulb

Figure 12.8. Beckmann thermometer

The Beckmann thermometer is a very accurate and sensitive form of


centigrade thermometer. Its important features are:
(i) A large bulb of mercury and a very thin capillary tube in which the
mercury moves. This combination makes the instrument very sensitive.
It is usually graduated in hundredths of a degree.
142 Physical Chemistry
(ii) The instrument has only a small range of temperature on its scale,
usually about 8 K. It is used for measuring temperature differences.
(iii) It has a subsidiary bulb at the top of its capillary tube which can be
used to vary the temperatures registered by the scale. Thesystem works
in the following way. Suppose the Beckmann thermometer (Figure
12.8) has been used for observations on the freezing points of aqueous
solutions. It is then adjusted so that the scale reads about — 6 K to
42 K. If it is to be used for boiling-point observations on aqueous
solutions, the scale should read about 371-379 K. This requires some
mercury to be forced out of the main bulb into the upper, subsidiary
bulb. To allow for the thread of mercury above the scale, the
thermometer is gradually heated in a liquid boiling at about 383 K (say
toluene) and this, at the boiling point, forces mercury up into the
subsidiary bulb. When the thermometer is registering the temperature
of 383 K it is tapped sharply so that the mercury expanded into the
upper bulb is detached from the thread of mercury in the capillary
tube. The top limit of the thread then registers 383 K at the entrance to
the upper bulb. On cooling 3-4 K, the end of the thread will fall back
on to the top of the scale which then represents about 379 K. The
8 K range of the scale will take it down to about 371 K as required.
To reverse this process, the detached mercury is tapped back to the top
of the upper bulb and the main bulb is then gradually heated in a bath of
suitable liquid till the mercury thread rejoins. t

(d) Osmosis and osmotic pressure


It is found that certain membranes exist which allow water to pass through
them, but do not allow dissolved solutes to pass. Such membranes are said
to be semi-permeable. The best known and most efficient semi-permeable
membrane is an artificially prepared one of copper hexacyanoferrate(II).
The preparation is described below. Animal membranes, such as pigs’
bladders, are often approximately semi-permeable but rarely perfectly so.
The different phases to be considered in this case are two aqueous
solutions of the same solute, but of different concentrations, which are
separated by a semi-permeable membrane. The rate of movement of
solvent molecules through the membrane from the less concentrated
solution to the more concentrated will be greater than in the reverse
direction and so there is a net movement in that direction and the solutions
tend to become equal in concentration. When a solution is separated from
pure solvent by a semi-permeable membrane, the pressure which must be
applied to the solution in order to balance and hence prevent the movement
of solvent into the solution, is called the osmotic pressure of the solution.
The passage of the solvent through the membrane is called osmosis. This
can be illustrated by the apparatus of Figure 12.9.
Phase Equilibria 143
A sugar solution inside the pot is separated from distilled water (acting
here as an infinitely dilute sugar solution) in the outer vessel by the semi-
permeable membrane. At the beginning, the inner and outer liquid levels
are equal. With the passage of time, distilled water moves through the
membrane into the pot and the level of liquid rises. As it does so, the
hydrostatic pressure of the rising column of sugar solution from the pot
opposes the effects of osmotic pressure. Eventually the column of sugar
solution becomes so long that the hydrostatic pressure of the column of
sugar solution just equals the osmotic pressure of the solution.

final level of liquid

original levels equal

sugar solution

distilled water
porous pot with
semi-permeable membrane
of copper hexacyanoferrate(II)
Figure 12.9. Illustration of osmotic pressure

The solution inside the pot at this stage is more dilute than the original
and it is the osmotic pressure of this more dilute solution which is being
balanced and not that of the original solution.

Preparation of a semi-permeable membrane


Natural semi-permeable membranes, such as animal bladders, are not much
use for scientific purposes. They are variable in quality, not fully semi-
permeable, and weak. A copper hexacyanoferrate(II) membrane is, for all
practical purposes, completely semi-permeable, i.e. it allows water to pass
quite freely but prevents the passage of dissolved solutes. It is, however,
very weak. Pfeffer (1877) strengthened the membrane by depositing it in
the walls of a porous pot so that the membrane acquired the mechanical
strength of the walls of the pot while losing little in semi-permeability.
144 Physical Chemistry
In Pfeffer’s method, the porous pot was cleaned with alkali solution and
then with potassium nitrate solution. It was thoroughly washed to remove
these materials, after which it was filled with water and subjected to
pressure so that water was forced into the pores. The pot was filled with a
3 per cent solution of potassium hexacyanoferrate(II) and placed in a
similar solution of copper(II) sulphate. The two solutions diffused into the
walls of the pot, depositing the membrane of copper hexacyanoferrate(II).
2Cu?* (aq) + Fe(CN)§~ (aq) > Cu.Fe(CN)¢(s)
The pot was then thoroughly washed with distilled water. A later method
of Morse and Frazer improved on Pfeffer’s by assisting the production of
the membrane by electrolysis as well as diffusion.

Measurement of osmotic pressure

Pfeffer’s method
The apparatus used is illustrated diagrammatically in Figure 12.10. It
must be kept at constant temperature. With the passage of time, water

graduated manometer
tube containing nitrogen

distilled water

porous pot with


copper hexacyanoferrate(I!)
membrane
Figure 12.10. Pfeffer’s method for the measurement of osmotic pressure

enters the porous pot by the operation of osmotic pressure, forcing the
mercury up the manometer tube and compressing the nitrogen. Eventually
the reverse pressure of the nitrogen equals the osmotic pressure of the
Phase Equilibria 145
solution and equilibrium is reached. The pressure can then be read from
the manometer scale. The weaknesses of this method are:
(a) Water enters the porous pot diluting the solution. The extent of the
dilution is known by the movement of mercury in the manometer tube
but it may not be equally distributed in the solution. Hence there is
some slight doubt about the effective concentration of the solution.
(b) The pressure is exerted against the walls of the porous pot. The strength
of the pot sets a limit to the osmotic pressures which can be measured.
In the modern equivalent of Pfeffer’s method the membranes are usually
made of cellulose or cellulose nitrate supported by perforated metal plates.
The osmotic pressure is determined either by the rise in a capillary tube or
by measuring the flow rates for certain applied pressures and interpolating
to find the pressure for zero flow rate.

Laws of Osmotic Pressure


These laws apply to unionized solutes only. The effects of ionization on
Osmotic pressure are considered later.

1. Concentration effect
For dilute solutions of a given solute at a constant temperature, osmotic
pressure, 77, is directly proportional to concentration, C.
The mathematical expression of this law is
7 . °
—=k (if T is constant)
c
Since concentration is inversely proportional to the volume of the solution,
this law can also be written
a xX V=k (T constant)

This is analogous to Boyle’s Law for gases, expressed in the form


px V=k (T constant)

2. Temperature effect
The osmotic pressure of a given solution is directly proportional to its
absolute temperature.
The mathematical expression of this law is:

- =k (if concentration is constant)


146 Physical Chemistry
This is analogous to the law for gases, expressed in the form

A =k (if Vis constant)

3. Molecular effect
Equimolar dilute solutions of different solutes at the same temperature
have the same osmotic pressure.
This law shows that osmotic pressure is determined by the number of
molecules of solute present in unit volume of solution and not by their size
or chemical nature. Thus, if 46 g of alcohol, 60 g of urea, 180 g of glucose,
and 342g of cane sugar are each dissolved separately to make, say,
10 dm® of solution, all the solutions will have the same osmotic pressure at
a given temperature. The figures are the relative molecular masses of the
solutes. Notice that, because of its lighter molecule, urea is, mass for mass,
three times as effective as glucose and nearly six times as effective as cane
sugar in setting up osmotic pressure.
By combining the mathematical statements of laws (1) and (2) above,
we obtain the composite equation

ap Se IY
= a constant

This is analogous to the expression of the gas laws, which takes the form

pxVv
= a constant
T

This similarity is not merely qualitative. It is quantitative too, the two


constants being identical in numerical value, as the following calculations
show.

For gases
1 mole of any gas at 273 K in 22.4 dm? exerts a pressure of 1 atmosphere.
Applying these figures,

pXx Lae xX 22.4


constant =
T Zs

= 0.082 dm® atmosphere K-! mol-#


Phase Equilibria 147

For osmotic pressure


Experiment shows that 10g of cane sugar (molecular mass 342) in
1 dm® of solution produce an osmotic pressure of 0.66 atmospheres at
273 K. From this,

constant = 7x _ 9.66
x 34.2
273
= 0.083 dm? atmospheres K-! mol-1
(V is the number of dm® containing 342 g of cane sugar.)
The same calculations using SI units are given below.

For gases
1 mole of any gas at 273 K in 2.24 x 10-2 m® exerts a pressure of
1£.0130x'.105 Nm-?.
The gas equation for 1 mole is pV = RT. The gas constant
Mert Ol xpl0" Xi2.24 x 1074
R
273
= 8.31 J mol~1 K7+

For osmotic pressure


Experiment shows that 10g of cane sugar (molecular mass 342) in
1 x 10-* m® of solution produces an osmotic pressure of 6.68 x 10* N m=?
Be zi3K.
1 x 10-3 x 342
Volume of solution containing 1 mole =
10
m= 3.42 %.1052.m?
Using 7 V = RT,
_ 6.68 x 10* x 3.42 x 107?
R
¢ 273
= 8.36 J mol~1 K7}

The slight difference in the figures is caused by experimental error in the


osmotic pressure data. This quantitative identity means that:
A solute produces in dilute solution the same osmotic pressure as it
would produce gas pressure, if it existed as a gas in the same volume
as the solution at the same temperature.
148 Physical Chemistry
That is, suppose a dilute sugar solution to be contained in an exactly full,
covered beaker. Then suppose the water to be suddenly whisked away,
leaving the sugar molecules suspended in otherwise ,empty space. The
sugar then functions as a gas, exerting a pressure equal to its former osmotic
pressure in solution. From this it follows that, just as 1 mole of any gas in
22.4 dm3 at 273 K exerts 1 atmosphere pressure, so 1 mole of any solute
exerts 1 atmosphere osmotic pressure. This offers a means of finding the
molecular mass of a soluble compound, for the relative molecular mass is
that mass in grammes which, at 273 K in 22.4 dm of solution, gives |
atmosphere osmotic pressure. Any observation of the osmotic pressure of
a suitable solution at a known temperature enables this calculation to be
made. An example is given below.

Calculations on osmotic pressure


These calculations can be performed by formula or from first principles.
Both methods are given below. The most convenient form of the osmotic
pressure formula is

7 Ty

ONT © CeeT,

where C is concentration and T is absolute temperaturé. This formula


combines laws (1) and (2) given above.

Example 1 20g of cane sugar in 2 dm? of solution at 283 K produce an


osmotic pressure of 0.68 atmosphere. Calculate the relative molecular mass
of cane sugar.
In using the above formula, one side of it (say the left-hand side) expresses
the experimental data given; the other side expresses the fundamental
principle that the relative molecular mass of cane sugar in grammes in
22.4 dm® of solution at 273 K gives 1 atmosphere osmotic pressure.
Let the relative molecular mass of cane sugar be M.

alll, Saeed Ce
GXe Ee CRxXAT.

0.68 rf 1
20/2 x 283 M/22.4 x 273

, 20 1 1
From this, MKB 5 ree OK rere
2 0.68 * 373 * 22.4
i.e. relative molecular mass of cane sugar = 341
Phase Equilibria 149
Working from first principles, the calculation takes the form:

0.68 atmospheres is the osmotic pressure developed in 2 dm® of solu-


tion at 283 K by 20 g cane sugar

1 atmosphere is the osmotic pressure developed in 22.4 dm® of solution


at 273 K by

Die
068
A eo
5 x 773 7 341 g cane sugar.

i.e. relative molecular mass of cane sugar = 341.


In the above fraction, the part 1/0.68 expresses the fact that the change
from 0.68 atmosphere to 1 atmosphere of osmotic pressure requires an
increase in the mass of cane sugar; the fraction 22.4/2 expresses the fact
that an increase in the volume of solvent from 2 to 22.4 dm® requires a
corresponding increase in the mass of cane sugar to maintain the same
osmotic effect; the fraction 283/273 expresses the fact that the /fa/l in
temperature from 283 K to 273 K requires an increase in the mass of cane
sugar to maintain the same osmotic effect.
The same calculation may be done by direct application of the gas
equation. SI units will be used in this case.
20 g of cane sugar in 2 x 10~° m® of solution at 283 K produce an
osmotic pressure of 6.89 x 10* N m~?. The gas constant is 8.31 J mol~* K7?.
The volume of solution containing | mole is given by

Vath
7

8.31 X 283
~ 6.89
x 104
= 3.414 x 10°? m®

But, 2 x 10-% m® contains 20g of sugar; therefore 3.414 x 10-* m®


contains
20 x 3.414 x 10-?
ewes = 341.4¢

The relative molecular mass of the sugar is 341.


Osmotic pressure effects for comparable mole fractions are much
higher than the effects on freezing and boiling points. Thus when the
relative molecular mass of a polymer, from which it is only possible to
prepare a solution with a small mole fraction, is determined the osmotic
pressure method is preferred.
150 Physical Chemistry

Abnormal relative molecular masses


When the relative molecular masses of certain compounds are determined
by the foregoing methods, abnormalities appear in the results. Some
compounds show higher relative molecular masses than would be expected
from their accepted chemical formulae; others show lower values.

Abnormally high relative molecular masses


These are comparatively rare and arise from polymerization of the com-
pound, i.e. from the formation of complex molecules by the combination
of two or more normal molecules. The normal relative molecular mass
is multiplied by the factor n.
nA— A,
Examples are benzoic acid and ethanoic acid (acetic acid) in benzene
solution, where dimerization occurs to (CsH;CO.H). and (CH3CO2H)z.
This phenomenon probably involves hydrogen-bonding (page 101).
Ethanoic acid is also dimerized in the vapour just above its boiling point.
Methanoic acid (formic acid) dimerizes in the vapour state, but only
partially.

Abnormally low relative molecular masses


g

These arise when the observed effect (osmotic pressure, elevation of


boiling point, etc.) is greater than expected. These high observed values
are due to the solute dissociating into ions.
Consider an electrolyte such as ethanoic acid (acetic acid) in aqueous
solution. Let | mole of the acid be dissolved to make a definite volume of
solution. Suppose that a fraction « of this mole is ionized when equi-
librium is reached at constant temperature (say 298 K). The position is
then:
ie CH,CO,H = CH,CO,- + Ht
Originally 1 mole _ -
At equilibrium (1 — «) mole a « mole
« is called the degree of ionization or dissociation of the acid.
The ‘normal’ osmotic pressure developed in the solution would be
related to 1 mole of unionized acid. But each of the ions exercises the same
influence in solution as the molecule from which it was derived; therefore
the total number of effective particles produced per molecule of the original
acid is (1 — a) + 2a, and the abnormally high osmotic pressure (and
eae point and boiling point changes) is related to this figure. From
this
(1 — a) + 2« _ __ osmotic pressure observed in practice
1 osmotic pressure calculated for no ionization
Phase Equilibria 151
The numerator of the fraction on the left-hand side of the equation is
always greater than unity. This explains the abnormally high osmotic
pressure produced by ethanoic acid in solution and, if the actual osmotic
pressure is experimentally determined, provides a means of finding «, the
degree of ionization of ethanoic acid. Parallel conclusions follow for
freezing-point depression and boiling-point elevation.
In general, if an electrolyte produces n ions per molecule and has, under
the relevant conditions of concentration and temperature, a degree of
ionization «,
=a) + osmotic pressure observed in practice
1 osmotic pressure calculated for no ionization

and correspondingly for freezing-point depression and boiling-point


elevation.
A dilute solution of a strong binary electrolyte such as sodium chloride
always produces an effect which is twice that calculated from the assump-
tion that it exists as an ion pair of relative formula mass 58.5. This indicates
that sodium chloride is completely dissociated in dilute aqueous solutions.

2. Solubility of a solid in a liquid


In this system there are two components, the solute and the solvent, and
two phases, the solution and the undissolved solid.

Definition of solubility
If a fixed amount of water is shaken with, say, common salt (preferably in
a state of fine division) at a fixed temperature, say 288 K, it will be found
that the first small quantities of common salt added will disappear into the
water, leaving a colourless mixture. This mixture is called a solution (more
strictly an aqueous solution) of common salt; it consists of two parts:
(a) the liquid which is said to dissolve the common salt and is called the
solvent, (b) the solid, which undergoes dissolution, and is called the solute.
If more and more common salt is added (with no temperature change),
it will finally be found that some of it remains undissolved and settles to
the bottom of the container. The solution then contains as much common
salt in solution as it can for that quantity of water and at that temperature;
it is said to be a saturated solution, and can be defined as follows:

A saturated solution is one which, at the temperature concerned, is


in equilibrium with undissolved solute.
The solubility of a substance must be defined with reference to a
saturated solution of it at a stated temperature in a standard amount of
solvent, and the following definition has been adopted:
152 Physical Chemistry
The solubility of a substance in water at a given temperature is the
mass of the substance in grams which will saturate 100 grams of water
at that temperature.

Supersaturation
If a boiling-tube is filled with water to a depth of about 2 cm, then filled up
with sodium thiosulphate crystals and heated, the sodium thiosulphate will
dissolve to give a colourless solution. If the liquid is stirred (to make it
uniform) and then cooled under the tap while being held quite still, it will
be found that no crystals separate. If then the liquid is seeded (orinoculated)
with a very small sodium thiosulphate crystal, crystals will at once begin
to separate, starting from the seeding crystal as centre and growing
steadily downwards into the solution. An almost solid mass of crystals will
be formed (and the temperature will rise considerably as they separate,
see page 191). After the crystals have separated, the solution left must still
be saturated because it is in equilibrium with crystals of the solute. Before
the crystals separated it must have been more-than-saturated or super-
saturated; this may be defined as follows:
A supersaturated solution is one that contains more of the solute in
solution than it can hold, at that temperature, in the presence of crystals
of the solute.
It is important to notice that supersaturated solutions are unstable.
They usually require the following conditions:
1. If aqueous they are given (to any marked extent) by relatively few
compounds, e.g. sodium sulphate Na,SO,.10H.O, and sodium thio-
sulphate Na,S.O3.5H.O. Supersaturation is more common with
organic solvents.
2. They generally require exclusion of dust particles, which might act as
centres of crystallization.
3. They usually require s/ow cooling. Sodium thiosulphate is exceptional
in giving supersaturation with rapid cooling.
4, They require avoidance of stirring or shaking.

Determination of the solubility of a solid, e.g. sodium chloride, in water at


room temperature
A suitable amount of distilled water (say 150 cm’) is placed in a beaker and
warmed to a temperature appreciably (but not greatly) above room
temperature, say to 300 K. Finely divided sodium chloride (pure) is added,
with stirring, until a little is left undissolved, i.e. the solution is saturated
at about 300 K. The liquid, protected from dust by a cover-glass, is then
cooled to room temperature. As it cools, more crystals of common salt
Phase Equilibria 153
will separate. The mixture is allowed to stand, with occasional stirring,
till it has had time to reach room temperature throughout and attain
equilibrium between solution and crystals. It is then filtered with completely
dry apparatus. The filtrate should be an exactly saturated solution of
common salt at room temperature, which should be noted.
In principle, the easiest analysis of the solution consists in adding some
of it to a weighed evaporating dish and reweighing, then evaporating the
liquid to expel all the water and weighing the (cooled) dish and solid. The
calculation is as follows:
Mass of evaporating dish =ag
Mass of evaporating dish + solution = bg
Mass of evaporating dish + solid =cg
Mass of water =(b—c)g
Mass of solid =(c—a)g

Solubility of common salt eli on) x 100g


in 100 g of water at t K. (6 — ¢)

In practice, this method offers many difficulties. Evaporation by steam is


essential, as evaporation by burner causes splashing and great inaccuracy.
When water has been expelled to the fullest possible extent by steam, the
solid will have to be dried further, say at 390 K, but even then it is not easy
to be sure of complete expulsion of water. This is especially so if the solute
forms a hydrate with water. There is also the possibility of hydrolysis of the
solute during evaporation; this is very common with chlorides.
In view of these difficulties, a chemical analysis may be preferable. In the
case of common salt, the following is suitable. A small beaker (with
watch-glass cover) is weighed. About 5cm® of the saturated sodium
chloride solution is added and the beaker is weighed again. The liquid is
then diluted to, say, 100-120 cm® with distilled water, transferred, with the
usual analytical precautions and washings, to a 250 cm® measuring flask,
made up to the mark with distilled water and shaken. It is then titrated, in
batches of 25.0cm® by 0.1M silver nitrate solution, with the usual
potassium chromate indicator. The following illustrates the calculation:
Mass of beaker and cover-glass = 75.010 g
Mass of beaker, cover-glass, and solution = 80.363 g
Mass of solution = 5.353 g

Solution diluted to 250.0 cm?


25.0 cm? of this dilute solution required 24.13 cm® (average)
of 0.1 M AgNO; solution
AgNO, = NaCl
10 dm® of 0.1 M 58.5 g
154 Physical Chemistry

24.13. 2000
Mass of sodium chloride = 58.5 x 10 000 x 750 g

=1411¢
Mass of water = (5.353 — 1.411) g
= 3.942 g
is 1.411
Solubility of common salt = 3943 * 100 g

= 35.8 g (in 100 g of water


at t K.)
It must be admitted that the actual procedure in such cases depends (to a
certain degree) on fore-knowledge of the result. For example, the choice of
‘about 5 cm® of the saturated solution for dilution depends on knowledge
that about one-quarter of the mass of the solution is common salt and the
density of the solution is about 1.2 g cm~ °. This choice then provides about
1.5 g of sodium chloride in 250 cm® of diluted solution, which is about
right for titration for 0.1 M AgNO solution. In the absence of such
knowledge, a trial run would probably be needed to acquire it.

Determination of solubilities at temperatures other than room temperature


The principles of the determination are the same as those of the method
already given, but the saturated solution must be prepared at the required
temperature and is then somewhat more difficult to manipulate and
analyse.
A suitable volume of distilled water is heated to the temperature of the
determination, placed in a thin-walled glass bottle with the finely divided
solid and immersed in a thermostat at the appropriate temperature, t K.
The mixture is stirred, preferably mechanically, with excess of the solid
continually present. Enough time must be allowed for saturation to be
reached (not less than 2 hours in general). The mixture is then allowed to
settle and some of the liquid is extracted by a dry pipette which has been
heated to a temperature somewhat above ¢t K and fitted with a filter to
prevent the entrance of crystals of solid. The liquid is weighed (as before) in
a pre-weighed vessel and analysed as before.

Solubility curves
If determinations of the solubility of a substance are made at suitable
temperatures, a graph can be drawn from the results with temperatures
along one axis and the solubility values along the other. Except in special
cases, such solubility curves are continuous. Generally, solubility of solids
increases with rise of temperature; in a few cases (e.g. calcium hydroxide),
Phase Equilibria 155
solubility decreases with rise of temperature; occasionally (e.g. calcium
sulphate), solubility increases over a certain temperature range and then
decreases. This Is Summarized in Figure 12.11. In some cases the solubility
curves are discontinuous. This is usually caused by the fact that the
compound in question can exist in two or more states of hydration.

100 KNO,
90
80
70
KCIO,
60

Na 2804

40 NaCl

solubility
water)
(g
of
100
per
g

CaSO,
0-283. _(303--323.. 343 363
temperature (K)
Figure 12.11. Solubility curves

It should be noted that most solids dissolve in water with absorption of


heat; consequently, if a solution is in equilibrium with its solid solute at a
certain temperature and the temperature is then raised, Le Chatelier’s
Principle (page 204) requires the equilibrium to shift so that the temperature
tends to fall again. That is, heat must be absorbed by solution of more
solute and the solubility therefore rises.

Solubilities of sparingly soluble compounds


The methods given above are not suitable for determining the solubilities
of the so-called ‘insoluble’ salts, such as barium sulphate or silver chloride.
For these, the following methods are available.

(i) Conductance measurements


A saturated solution of the compound (say silver chloride) is prepared in
very pure water at the temperature in question (usually 298 K). The
156 Physical Chemistry

resistivity (page 277) of the solution is determined and, from its reciprocal,
the conductivity is available. From this, the conductivity of water is
deducted to obtain the conductivity of silver chloride. In such very dilute
conditions, the silver chloride is taken as fully ionized so that the
following relation holds:

molar conductivity for = 1000 x conductivity of AgCl


AgCl at infinite dilution no. of mole of AgCl per dm?

The molar conductivity is the sum of the ionic conductivities of Ag* and
Cl- and is 138 ohm-! cm2. The conductivity of saturated silver chloride
solution, and water, are 3.41 x 10-6 and 1.60 x 10-® ohm! cm-!
respectively at 298 K. By subtraction, the silver chloride figure is 1.81 x
10-6. From this,

respectively ~ no. of mole of AgCl per dm®


(1.81 x 107°)
solubility of AgCl in moldm~* = 1000 x
138
= 1.31 x 10-5 (298 K)
In grams per dm® this is (1.31 x 1075) x 143.5 =10.001 88g dm>=
(AgCl = 143.5.)

(ii) Colorimetric method


This can be applied when a metal forms a suitably coloured salt. The case
of lead sulphate is typical. A saturated solution of this salt is prepared at
the relevant temperature. Hydrogen sulphide is passed through it, giving a
brown liquid containing colloidal lead sulphide.
Lead ethanoate (acetate) solution of known concentration is then
progressively diluted until, on passage of hydrogen sulphide, it matches the
colour from the lead sulphate solution. The concentration of Pb?* in the
sulphate solution is then equal to the known concentration in the ethanoate
solution. The solubility of the sulphate is then obtained by multiplying by
the ratio PbSO,: Pb.

(iii) Radioactive method


This method can be applied (say to lead sulphate) in the following way.
Lead nitrate is dissolved in water and a known (very small) proportion of
the same salt of a radioactive isotope of lead is added. All the lead is then
precipitated as sulphate and the precipitate is thoroughly washed. The
sulphate is then agitated with water till a saturated solution is obtained at
the relevant temperature. A known mass of the filtered solution is evapor-
ated to dryness and the amount of radioactive material is estimated in the
Phase Equilibria 157
residue by a Geiger counter. The proportion of this being already known,
the total mass of solid can be calculated and hence its solubility.

Some further features of solubility curves


The behaviour of mixtures of potassium iodide and water
If pure water is cooled at a pressure of 101 300 Nm-? (760 mmHg) and
super-cooling is avoided, ice will separate as soon as the temperature of the
water reaches 273 K. If a very dilute potassium iodide solution is similarly
cooled, the freezing point will be a little depressed and ice will separate at a
temperature a little below 273 K. If the concentration of potassium iodide
is gradually increased, the freezing-point depression is accentuated and ice
separates at progressively lower temperatures. This behaviour is expressed
in the curve AB of Figure 12.12.

|
|
|

9iE 373 r |
!
oe atid
| |
8 © | |
ms
5 8B solutions |
| solid KI - |
46 + saturated |
oD solution
3 E2734
rae
oO
|
:
-
ES oO
M |
£2 ice + KI solution
250}-~~—--—-—-—-—-—=f-------- =
: |
ice + eutectic Kl + eutectic |
!
0% KI 52% KI 70% KI
100% HO
Figure 12.12. Cryohydrate of water-Kl

If a hot, concentrated solution of potassium iodide is cooled, it will


finally reach a point at which crystallization of potassium iodide will begin.
This point of crystallization will occur at lower temperatures for more
dilute solutions and at higher temperatures for more concentrated solutions.
The expression of this in graphic form is the curve BC, which is, in fact, the
solubility curve of potassium iodide in water.
158 Physical Chemistry
The situation can be illustrated by considering what happens when
particular mixtures of water and potassium iodide are cooled.
If a solution of composition L (about 12 per cent KI) is cooled, ice
begins to separate at the point M (about 268 K). As cooling is continued,
more ice separates, and the composition of the solution follows the curve
MB until the temperature falls to 250 K. By this time the solution has the
composition B. It then remains constant in temperature while a mixture of
composition B separates as solid crystals.
If a solution of composition N (or B), i.e. 52 per cent KI, is cooled, no
separation of solid occurs at all until the temperature reaches 250 K. Then
the temperature remains constant while a mixture crystallizes out which
has the same composition as that of the solution, i.e. 52 per cent KI and
48 per cent HO. This mixture is called the eutectic between potassium
iodide and water. If, as in this case, one of the constituents of the eutectic is
water, the eutectic may also be called a cryohydrate. The eutectic is clearly
that mixture of its constituents which has the lowest freezing point.
If a solution of composition O is cooled, potassium iodide crystals begin
to separate at P. As they form, the composition of the solution follows the
curve PB; then, at 250 K, the eutectic again separates as before.
At all points above the curves AB and BC, only solutions can exist. For
this reason, the line ABC may be called the Jiquidus. In a similar way, at
all points below the horizontal line of 250 K, only solids can exist. Con-
sequently, this line may be called the solidus. In other areas, equilibrium
mixtures can exist as marked.
As the name cryohydrate implies, it was at one time thought that a
cryohydrate (or eutectic) was a compound because of its constant com-
position and constant melting point. This is not so, because:
1. The eutectic can be seen to be heterogeneous under the microscope,
whereas a compound would be homogeneous.
2. The composition of a eutectic rarely corresponds to that of a compound
and then only by chance.
3. The properties of the eutectic, e.g. its heat of solution, are simply the
sum of the properties of its constituents. This is very unlikely for a true
compound.

Other simple cases of eutectic formation


The typical feature of the water-KI system treated above is that the
constituents form no compound. Many other similar cases occur, especially
between metals. The case of zinc-cadmium is well known and is described
in its main features by Figure 12.13.
If a mixture of composition A is cooled, it will remain liquid to the point
B. Then solid zinc will begin to separate. The liquid will follow the curve
BC in composition and will finally reach the composition of the eutectic,
which will separate as solid at 543 K.
Phase Equilibria 159
A liquid of the eutectic composition C will remain entirely liquid to
543 K if cooled, and will then deposit the solid eutectic mixture at that
temperature.
A mixture of composition E will remain liquid when cooled until point
D is reached. Then cadmium will separate as solid and the composition of
the liquid will follow the curve DC. Then, at 543 K solid eutectic will
separate.

692
E
|
|
|
|
|
|
|
z |
2 |
= Zn + liquid |
= I
om
=
£
543
Zn + eutectic Cd + eutectic

100% Zn 100% Cd
Figure 12.13. Eutectic of Zn-Cd

Compound formation
The formation of a compound produces a maximum in the freezing-point
curve as in Figure 12.14. The freezing point and composition of the
compound of X and Y (of the type X,,Y,) are represented at the point F.
The addition of X or Y to this compound depresses the melting point of
the compound, since both X and Y act as impurities with respect to the
compound; consequently the freezing point is lowered on each side of F,
producing the typical ‘hump’ effect (DFH) associated with compound
formation.
Two eutectics occur in this case, of compositions represented by E, and
E,. E, is a eutectic between X and the compound; E, is a eutectic between
the compound and Y. There is no eutectic between the materials X and Y.
The following effects accompany the cooling of the various mixtures
shown:
A. At temperature B, solid X begins to separate. The composition of
liquid follows that of the curve B to E,. When the point E, is reached,
eutectic E, begins to separate and continues (at constant temperature)
till all is solid.
160 Physical Chemistry

E Y + compound
1 |
X + ‘compound |

100% X | 100% Y
composition of compound of X and Y

Figure 12.14. Eutectics with formation of compound

C. At temperature D, the compound begins to separate out solid. As it


does so, the composition of the liquid follows the curve D to E,. When the
point E, is reached, eutectic E, begins to separate out solid and so on as
in A.
E. This is the composition of the compound of X and Y. At F this
compound begins to separate as pure crystals and continues to do so (at
constant temperature) till all is solid.
G. At temperature H, the compound begins to separate as solid. The
liquid follows the compositions represented by the curve H to E,; when
E, is reached, the eutectic E, begins to separate as solid and does so (at
constant temperature) till all is solid.
J. At temperature K, solid Y begins to separate. As it does so, the
liquid follows the compositions represented by the curve K to Es. When
the conditions of E, are reached, this eutectic begins to separate as solid and
so on as in G.
Mixtures of the eutectic compositions E, and E, behave in practice like
the compound, except that the appropriate eutectic separates out and the
temperatures of solidification are lower.
The curve B E,D F H EK is the liquidus above which (in the figure) only
liquid exists; the horizontal dotted lines through E, and E, represent a
solidus below which only solid exists. The other areas of the figure represent
the conditions marked in them.
The metals magnesium and tin behave in accordance with the situation
represented in Figure 12.14. The compound is Mg,Sn (Mg, 28.8 per cent).
Phase Equilibria 161
Much more complex cases of compound formation occur, with hydrates
in the solubility curves of compounds, usually salts. The case of iron(II)
chloride is given in illustration in Figure 12.15. The curve shows iron(IID)
chloride forming four hydrates, Fe,Cl,.xH.O, where x has the numerical
values shown at the maxima of the curve.

330

330

temperature
(K)
273

210
5 10 15 20 25 30
solubility
Figure 12.15. Hydrates of iron(III) chloride

3. Vapour pressure of salt hydrates


The behaviour of the three hydrates of copper(II) sulphate (which contain
5, 3, and 1 mole of water combined with one mole of copper sulphate) is
typical of that of salt hydrates generally.
If the pentahydrate CuSO,.5H,O is placed in an otherwise evacuated
space at 323 K, it is found that it dehydrates gradually to the stage
of CuSO,.3H,O. However, the vapour pressure remains steady at
6266 N m~2(47 mmHg) so long as any of the pentahydrate remains.
As soon as all is converted to the trihydrate, a rapid fall of vapour
pressure occurs to 4000 N m~? (30 mmHg). It then remains steady at this
value as dehydration to CuSO,.H,O occurs, and there is no change of
vapour pressure so long as any CuSO,.3H,0 is left.
As soon as CuSO,.H,0 is the only constituent of the solid, another rapid
fall to 600 N m~? (4.5 mmHg) occurs in the vapour pressure and this value
is maintained so long as any monohydrate is present. These changes are
summarized in Figure 12.16.
162 Physical Chemistry

oe CuSO,.3H,0 CuSO,4.5H,0

(mmHg)

323K
at
pressure
vapour

x in CuSO,.xH,0

Figure 12.16. Hydrates of copper(II) sulphate

The phenomena of deliquescence and efflorescence are closely connected


with vapour pressure of hydrates.
Deliquescence is the phenomenon by which a solid absorbs water-
vapour from the air and passes into solution.
The explanation of this phenomenon is as follows. It is characteristic of a
deliquescent solid that its saturated solution has a /ow vapour pressure,
e.g. a saturated solution of the deliquescent compound, calcium chloride,
at room temperature has a vapour pressure of only about 999 N m=?
(7.5 mmHg). Consequently, when an aqueous film from the air condenses
on the solid salt, the vapour pressure of the solution formed is so low that
the vapour pressure of water in the surrounding air is usually greater.
Consequently more water condenses into the solution until finally the solid
is entirely dissolved. Still more water may then be absorbed till the vapour
pressure of the diluted solution is equal to that of water in the air. It will be
obvious that if the aqueous vapour pressure in the air falls below
999 N m~’, calcium chloride will not deliquesce. The vapour pressures of
saturated solutions of most salts are greater than the usual aqueous
vapour pressure in the atmosphere, which is usually about 1600 N m=?
(12 mmHg). Consequently, most salts are non-deliquescent.
Efflorescence is the loss of water of crystallization from a salt hydrate
on exposure of it to air at room temperature.
The loss may be only partial, e.g. sodium carbonate decahydrate effloresces
according to the equation:
Na,CO3.10H,O(s) — Na,CO3.H,O(s) + 9H,O(g)
The loss of the water is usually accompanied by a collapse of the crystal
lattice, so that the crystals of the hydrate fall to powder.
Phase Equilibria 163
Efflorescence occurs when the dissociation vapour pressure of the
hydrate is greater than the partial pressure of aqueous vapour in the
atmosphere. For example, the vapour pressure of sodium carbonate
decahydrate is about 3200 N m~? (24 mmHg) at room temperature. This
exceeds the usual 1600 Nm~? pressure of aqueous vapour in the
atmosphere, so that washing soda will usually effloresce. Sodium and
iron(II) sulphates, Na,SO,.10H,O and FeSO,.7H.O, are also usually
efflorescent. It is obvious that in very humid surroundings efflorescence
may be suppressed because aqueous vapour pressure in the atmosphere
balances that of the hydrate.

4. Solubility of gases in liquids


As far as is known at present, no gas has a high solubility in water unless it
reacts chemically with the water. Gases such as nitrogen, hydrogen, and
oxygen are typical of those which dissolve in water without chemical
action; their solubilities are in the region of 2-4 volumes of gas in 100
volumes of water at room temperature. Gases of much higher solubility, of
which chlorine, carbon dioxide, hydrogen chloride, and ammonia are
representative, have solubilities in the region of 100-80000 volumes in
100 volumes of water at room temperature. They all react with water, e.g.

CO.(g) + H,0() = H,CO;(aq) = 2H* (aq) + CO2- (aq)


Cl.(g) + H,00) = HCl(aq) + HClO(aq)

Determination of the solubility of a sparingly soluble gas


This is usually done by means of an absorption pipette (Figure 12.17). The
experiment will be described for oxygen in water.
The gas pipette, P, has a three-way tap at E and an ordinary tap at R. Its
capacity (v, cm*) must be known. By suitably opening taps C and Etoa
source of oxygen connected at C, air is displaced from the copper (or lead)
tube D. E and C are then closed. By raising A, B can be filled with mercury
- (air being expelled at C) and then, by connecting the oxygen supply at-C,
oxygen can be passed into B as A is lowered. The volume of oxygen is read
off, a, cm, with levels of mercury equal in A and B. The pipette, P, full of
water, is then connected; A is raised and C, E and R opened so that water
runs out of P at R into a measuring cylinder, where the volume of water is
measured, v, cm®. The volume of water left in P is then (v, = v2) cm’,
P is placed in a thermostat at the temperature required and is shaken
repeatedly. As the gas dissolves, the tube A is adjusted to keep the levels
equal in A and B (i.e. to maintain atmospheric pressure). When the level in
B is constant, it is read off, a, cm®. re
The initial volume of oxygen (neglecting the volume of it in tube D,
which remains constant) is a; cm*. The final volume of oxygen is
164 Physical Chemistry
(ag + v2) cm*. That is, at the given temperature amis pressure,
(a, — dg — v2) cm® of oxygen are absorbed by (v; — v2) cm of water.
The water used in the experiment must be air-free. Tube D is made of
metal not rubber, because rubber is appreciably permeable to oxygen and
is liable to distortion and change of capacity when shaken.

mercury

Figure 12.17. Solubility of a gas

Definition of solubility of a gas in a liquid


The solubility of a gas in a liquid is best expressed by means of an
absorption coefficient; this can be defined as follows:
The absorption coefficient of a gas, A, ina liquid, B, is the number of
cm? of A, expressed at s.t.p., which saturate 1cm? ofthe liquid, B, at the
temperature stated and at a pressure of one atmosphere.
Some absorption coefficients are given below for the solvent water:

Gas Absorption coefficient


At273K At 293K

Hydrogen chloride 506 442


Carbon dioxide eel 0.88
Oxygen 0.049 0.028
Nitrogen 0.024 0.015
Phase Equilibria 165
It will be noticed that all these gases become Jess soluble with rise of
temperature. This is opposite to the behaviour of most solids. The
dissolution of a gas in water is generally an exothermic process; con-
sequently, if the temperature of a system of gas and liquid in equilibrium is
raised, the equilibrium must shift so as to tend to Jower it again (by Le
Chatelier’s Principle, page 204); that is, heat must be absorbed by expulsion
of the gas from solution.

Henry’s Law
Subject to certain qualifications mentioned below, Henry’s Law states the
effect of change of pressure on the solubility of a gas. The law takes the
form:

The mass of a given gas which saturates a constant volume of a given


solvent is directly proportional to the pressure ofthe gas at equilibrium,
provided that the temperature is constant and that the gas does not react
chemically with the solvent.

Notice that Henry’s Law does not apply to the aqueous behaviour of
very soluble gases, which invariably react chemically with the water in
which they ‘dissolve’. Slight inaccuracies also arise in Henry’s Law because
gases do not obey Boyle’s Law exactly.
If gaseous volume is kept constant at constant temperature, the mass of
gas contained in that volume is directly proportional to the pressure.
Consequently Henry’s Law can also be expressed by the statement that, at
constant temperature and provided that the gas does not react chemically
with the solvent, the volume of a given gas which saturates a given volume
of solvent is constant and independent of pressure.
The following figures illustrate the extent to which Henry’s Law is
obeyed by oxygen in water at 298 K. For perfect agreement with the law,
the final column should show constant figures:

A B B
Pressure Mass of oxygen (g) to se
(N m-?) saturate 1 dm* of water A

101 300 0.0408 5.31.


81 330 0.0325 5.33
55 200 0.0220 meh
40 000 0.0160 5.33
23 330 0.0095 5.43
rrr ——
166 Physical Chemistry
Solution of gases from mixtures
The volume of a gas dissolved by water from a mixture is directly pro-
portional to two factors and, therefore, to their product. These factors are:
1. The absorption coefficient of the gas.
2. The partial pressure of the gas in the mixture; this can be defined as
follows:
The partial pressure of a gas is that pressure which the gas would exert
if it alone occupied the whole volume actually occupied by the mixture,
temperature remaining constant.
From this it follows that the ratio of partial pressures in a mixture of gases
is the same as the ratio of volumes.
The operation of these factors can be illustrated by their application to
solution of air in water. Carbon dioxide is omitted because Henry’s Law
does not apply well to this gas.

Ne O2 Ar
Volume composition of air
per cent 78 21 1
Absorption coefficient of gas
at ous: 0.024 0.049 0.053
Vols. dissolved in water are . .
proportional to 78 x 0.024. 21 x 0.049 I x 0.053
==)1187 == £03 = 0.053
Total vol. of gas dissolved is
proportional to 1.87 + 1.03 + 0.053
= 2.95 (approx.)

Consequently, the volume composition of air dissolved into water, or


boiled out from it, is:

No O, Ar
1.87 1.03 0.053
—— x 100 per cen t
595 — x 100. per cent
595 ——
795 x 100 per cent

= 63.4 per cent = 34.9 per cent = 1.80 per cent

Determination of the solubility of a very soluble gas


The method of the gas-pipette is not suitable for this determination. A
pyknometer is used (Figure 12.18). The pyknometer is weighed. A suitable
volume of water. is then introduced into it. With the pyknometer placed in
a constant-temperature bath, the dry gas is passed until the solution is
Phase Equilibria 167

Figure 12.18 Pyknometer

saturated. The pyknometer is then sealed off by heating and drawing off its
ends in a blowpipe flame. After cooling, the whole (including any parts
drawn off during sealing) is weighed. This gives, by subtraction, the mass
of solution. The solution 1s then analysed. It will generally be either
alkaline or acidic, most of the very soluble gases being one or the other. If
the liquid is acidic, the container is broken under excess of standard
alkali and the excess is titrated. For alkaline liquids, excess standard acid
is the absorbent.

5. Mixtures of liquids
When considering a mixture of two liquids (two components) in equilib-
rium with the mixture of the two vapours (two phases), there are three
variables to consider: temperature, pressure, and the composition of either
the liquid mixture or the vapour mixture.
If two completely miscible liquids behave ideally, the total vapour
pressure of the mixture (at a given temperature) is equal to the sum of the
partial pressures of the two constituents. This is because both constituents
obey Raoult’s Law of Vapour Pressure. They do not enter into combination
with each other or cause dissociation or polymerization. This case is
shown in Figure 12.19. The vapour pressure T, of the mixture is the sum of
the vapour pressures PX and PY of the two constituents, and correspond-
ingly for each point on AB. This ideal position is rarely, if ever, achieved
in practice, but many cases occur in which it is approximately reached, so
that the vapour pressures of mixtures of two such liquid components
168 Physical Chemistry
always lie between the vapour pressures of the two constituents (at the
relevant temperature). No maximum or minimum of vapour pressure 1s
shown. Correspondingly, the boiling-point curve at constant pressure
shows no maximum or minimum (Figure 12.20).
A

2
2
© o
®
5] 5
85
fae.
3
a
= =
55
ee:
$8
ee > g
$
% re
-~vapour pressure of N

100% M P 100% N
Figure 12.19.

2
®
)z liquid =
3S 28
Es a=
se Ps
a3
Oc
32
Pome)
25 8
— oO
o

100% M 100% N 100% M 100% N


Figure 12.20.

A lower vapour pressure than expected (negative deviation from


Raoult’s Law) is due to bond formation between the components of the
mixture (see page 98). A higher vapour pressure than expected (positive
deviation from Raoult’s Law) indicates that bonds which existed between
the molecules of one component are stronger than bonds between mole-
cules of the different components.
A mixture of this sort can be completely separated into its constituents
by distillation. The dotted line in the boiling-point diagram represents the
composition of vapour in equilibrium with the liquid mixtures. In Figure
12.20, liquid A is in equilibrium with vapour B, which is richer in the more
volatile constituent, M, than A is. If mixture A is distilled, vapour B can
Phase Equilibria 169
be condensed to liquid C, which yields vapour D on distillation. This
condenses to liquid E, giving vapour F and so on. It is evident that a
succession of ideal distillations of this kind will finally yield pure M as
vapour, which can be condensed to liquid. Ideal conditions will leave pure
N in the distillation vessel and separation is theoretically complete.
By the use of a fractionating column, a fair approximation to the above
ideal state of distillation can be reached. A typical fractionating column is
shown in Figure 12.21 and consists of a central rod with discs at short

vapour of
=~ pure M
to condenser

vapour B —

Figure 12.21. Fractionating column


170 Physical Chemistry
intervals, the discs being almost as wide as the tube in which they are
placed. Ideally, the ascending vapour should come into equilibrium with
a liquid layer condensed on each disc. If so, the position corresponds to
that of Figure 12.20 in the manner shown. The thermometer will register
the boiling point of M until all of M has been delivered; then vapour of N
will distil and the thermometer will register the boiling point of N.
Examples of mixtures which can be separated into their constituents in
the above way are methanol—-water; benzene-toluene; benzene-hexane.

Mixtures of minimum boiling point


If the vapour-pressure curves of the two constituents are concave
upwards, the vapour-pressure curve (at constant temperature) of the

vapour pressure
of mixture

~N vapour pressure
~~ of M

pressure
vapour boiling
point
at
constant
pressure
temperature
constant
at

100% M 100% N 100% M 100% N


Figure 12.22.

mixture may show a maximum. Correspondingly (Figure 12.22), the


boiling-point curve (at constant pressure) will show a minimum. If liquid A
is distilled with a fractionating column, it will yield vapour, B, condensing
to liquid, C, vaporizing to vapour, D, condensing to liquid, E, and so on.
It is evident that the distillation will yield the mixture, G, which can be
condensed, i.e. the mixture of minimum boiling point. Thus, complete
separation of the constituents is not possible by distillation.
The mixture, G, is in equilibrium with vapour of its own composition
and will distil unchanged at the constant, minimum boiling point at
constant pressure. This mixture is said to be azeotropic. When the removal
of such a mixture has exhausted one of the constituents, the other (excess)
constituent will be left pure and can be distilled. A corresponding situation
is found if the original mixture has a composition on the other side of G.
Again, the azeotropic mixture distils, leaving the other constituent as
residue; notice, again, that separation of pure samples of both constituents
(from a given mixture) is not possible by distillation.
Phase Equilibria 171
The most important case of this kind is that of ethanol-water. The
azeotropic mixture (boiling point 351.1 K at 101 300 N m-? (760 mmHg))
has 95.6 per cent of the alcohol and this is the limit of purification of dilute
alcohol by distillation. Modern versions of the Coffey still can yield
alcohol of this composition in a continuous process starting from dilute
ee of alcohol (about 8 per cent) produced by fermentation of
starch.

Mixtures of maximum boiling point


If the vapour-pressure curves of the two constituents ofa liquid mixture
are convex upwards, the vapour-pressure curve may show a minimum

vapour pressure
§ of mixture ©
es
=] g wD
2
QE .=eS
25 \ 9.2,
a | vapours =
& € | pressure \ 2s
a5 of M 3 0)2
fom

$s 8
% v7vapour iy
7 pressure
ot gs of N

100% M 100% N 100% M 100% N

Figure 12.23.

(at constant temperature). Correspondingly (Figure 12.23), the boiling-


point curve, at constant pressure, may show a maximum. In this case, a
liquid with the composition of the azeotropic mixture distils unchanged at
the maximum boiling point, being in equilibrium with vapour of its own
composition. A mixture such as A distils vapour B, condensing to liquid C,
vaporizing to D, and so on, delivering finally a liquid which tends to be
pure M, or, at any rate, much richer in M than A was. The residual liquid
becomes more concentrated in N until it reaches the composition of the
azeotropic mixture, and then distils unchanged at the maximum boiling
point. Similar changes occur on the other side of the azeotropic mixture,
tending to deliver pure N as distillate and leave the azeotropic mixture as
residue. Here again, complete separation of the constituents from a given
mixture is impossible and any mixture, if distilled long enough, leaves the
mixture of maximum boiling point.
All the mineral acids show behaviour of this kind with water. For hydro-
chloric acid, sulphuric acid, and nitric acid, the data are:
2 Physical Chemistry

Boiling point
Maximum boiling- at 101 300 N m~2
point mixture (760 mmHg)

Hydrochloric acid 20.2% HCl 383 K


Sulphuric acid 98.79, H.SO; 611 K
Nitric acid 68.0% HNO; 394K

6. Distillation of non-miscible liquids


If two liquids are immiscible, their behaviour when distilled is quite
simple. They operate quite independently and there are three phases
present, and each liquid is in equilibrium with its own vapour. Each contri-
butes its own separate vapour pressure at a particular temperature and the
total vapour pressure of the mixture is the sum of the two contributions.
The mixture boils when the total vapour pressure reaches the prevailing
atmospheric pressure. The most important application of this situation is
found in steam distillation, which is employed mainly in the purification of
organic compounds.
The mixture for steam distillation is heated and steam is passed in from
a can in which water is boiled. The issuing vapour, which (idéally) contains
only water and the organic compound, is condensed by passage through a
Liebig’s condenser (see text-books of organic chemistry for details). The
composition of the distilled mixture is given by the following calculation.
Consider the case of a compound, X, insoluble in water and forming an
aqueous mixture which boils at 370K at 101 000 N m~? (757 mmHg)
pressure, the vapour pressure of water at 370K being 96000 N m-?
(720 mmHg). Let X have a relative molecular mass of 125. The vapour
pressure of X at the boiling point is (101 000 — 96000) Nm? or
5000 N m~?, From this, in the vapour which distils,

vol. of X ____ vapour pressure of X


vol. of steam vapour pressure of steam
__ 5000
~ 96.000

mass of X in distillate 5000 x (mol. mass of X)


mass of water in distillate 96000 x (mol. mass of steam)
et) Kal 2 tates He
~ 96000 x 18 2.76
Phase Equilibria 173
That is, the percentage by mass of X in the distillate is

1
376 x 100 = 26.6 per cent

The remainder is water.


As can be seen from the above calculation, to achieve a reasonable
proportion of the compound in the final distillate, the compound should
have the following characteristics:

1. It should have a considerable vapour pressure at temperatures near the


usual boiling point of water.
2. It should have a fairly high relative molecular mass (usually about
90-140).
Steam distillation can be particularly useful when a compound is liable
to decompose near its normal boiling point, as the process of steam
distillation, in a similar manner to distillation under reduced pressure,
results in the liquid boiling at a temperature below its normal boiling
point.

Three-component Systems
The Partition Law (or Distribution Law)
The Partition Law is concerned with the distribution, between two solvents,
ofa material which is soluble in both of them. It can be stated as follows:
If a solute X distributes itself between two immiscible solvents A and B
at constant temperature and X is in the same molecular condition in both
solvents,
concentration of X inA
—__—__—__—..,.._, = a constant
concentration of X in B

This constant is called the partition coefficient of X with respect to A and


B.

Experimental illustration of the Partition Law


Equal volumes of water and tetrachloromethane are shaken in a separating
funnel with a known mass of iodine. If the data are required not at room
temperature but at (say) 298 K the funnel is placed in a thermostat at this
temperature and left there for a considerable time, during which it is shaken
vigorously at frequent intervals. This should ensure an equilibrium
distribution of the iodine between the two solvents.
After time for settling, the lower (CCl,) layer is run off and the aqueous
layer (or a known fraction of it) is then titrated for iodine content by
174 Physical Chemistry
standard thiosulphate solution, say 0.01 M. The mass of iodine in the
aqueous layer is given by the relation.
I, = 2Na.S8203

2 Xm Zing 200 dm? 0.01 M soln.

and, from this and a knowledge of the volume of water used, the concentra-
tion of iodine in the aqueous layer is known. The total mass of iodine less
the iodine in the aqueous layer gives the mass of iodine in the tetra-
chloromethane layer, and the concentration of iodine in this layer can be
calculated.
The experiment can then be repeated with varying masses of iodine and
varying proportions of water to tetrachloromethane. In all cases, the ratio
of the iodine concentrations in tetrachloromethane and water should be
(approximately) the same.
For example, suppose 1.00 g of iodine was shaken with 50 cm® of each
solvent and 25.0 cm® of the aqueous layer required 4.50 cm® of 0.01 M
sodium thiosulphate
I, = 2Na,S8,03

DERN OA Gs 200 dm* 0.01 M


se cons 9.00
M ass off iodine
iod in thethe total
total aqueous ]layer == ((2 x
x 127) )xx ———
300000 ®

= 0.0114 g
Mass of iodine in tetrachloromethane = (1.00 — 0.0114) g
= 0.9886 g
.. Partition coefficient = Psu Eel=o
0.0114/50 1
Typical results obtained by varying the mass of iodine and the volume of
solvents are: 87.9, 88.3, 87.5.

Extraction of organic compounds by ethoxyethane


The extraction (usually of organic compounds) from aqueous mixtures by
ethoxyethane is probably the most important application of the Partition
Law. Ethoxyethane is not, in fact, completely insoluble in water (which
dissolves about 7 per cent of its mass of ethoxyethane at room tempera-
ture), but the law can be applied approximately to cases of extraction, in
calculations of which the following is typical:
Example A solution of 6g of a substance X in 50cm? of aqueous
solution is in equilibrium, at room temperature, with a solution of X in
ethoxyethane containing 108 g of X in 100 cm3. Calculate what mass of X
Phase Equilibria Ly:
could be extracted by shaking 100 cm? of an aqueous liquid containing
10 g of X with (a) 100 cm? of ethoxyethane, (b) 50 cm? of ethoxyethane
twice, at room temperature.

From the first data, partition coefficient of X _ 108/100


between ethoxyethane and water 6/50

_ 108 50Gama!
~ 100
_9
Pa
ltitini
In extraction (a), let x g of X pass into ethoxyethane, Then, » a
fo \S Rlq
x/100 ats ee:
(10 — x)/100 «1 /i
910—x)_ x |
100 ~—«:100
aued
10x = 90 4 ar) i Q
‘ ¢
= - oy ~
4 X hy aah ted = >
i.e. 9 g ofX pass into ethoxyethane.
In extraction (b), for the first extraction, let a g of ie — into.ethOxy-
ethane. Then
alS0__9
(10—a)/100 1
KIT —"a) = a
100 50

lla = 90

pew!
Beat1
= 8.2 g (approx.)

For the second extraction, 1.8 g of X remain in the water. Let b g of X pass
into ethoxyethane. Then
b/50 9
(unsss D100?
91.8 —b)_ b
100- .. 50
11b = 16.2
b = 1.5 g (approx-)

The total extraction of X into ethoxyethane is (8.2 + 1.5) g = 9.7 g


176 Physical Chemistry
It will be seen that the extraction is more efficient (9.7 g) if the same
ethoxyethane is used in two half-batches instead of all at once (9.0 g).
The above treatment of the Partition Law has assumed throughout that
there is no difference between the molecular conditions of the solute in the
two liquid layers. If such a difference arises, the situation becomes more
complex. Two of the simpler possibilities are mentioned below.

1. Solute ionized slightly in water and not at all in the other liquid
In this case, the Partition Law equilibrium is maintained between the
unionized molecules in both solvents. Since there are no ions in the non-
aqueous layer, the question of ionic equilibrium cannot arise as between
the two layers, but an ionic equilibrium is maintained, according to the
Dilution Law, in the aqueous layer.
If C, is the total concentration of weak electrolyte, X, in the aqueous
layer and the degree of ionization is «, the concentration of unionized
molecules in the aqueous layer is C,(1 — «). By the Partition Law,

CU Cy a) = K ((at constant temperature


————. Pp )

where C, is the total concentration of X in the non-aqueous layer.


A case of this sort arises in the distribution of a weak acid (e.g. succinic
acid or oxalic acid) between ether and water.

2. Solute polymerized in the non-aqueous layer and in unimolecular


condition in the aqueous layer

Suppose a solute X is in the unimolecular condition in water and


polymerized (entirely) as molecules X, in the non-aqueous solvent. The
results of experimental investigation show that in this case the Partition
Law takes the form

"V(Cone. of Xin non-aqueous layer) __ a constant (at constant


Conc. of X in water temperature)

The best-known cases of this are the equilibria between dimerized


ethanoic acid, (CH;CO2H)., or benzoic acid (benzenecarboxylic acid),
(C.sH; CO2H).2, in benzene and the unimolecular forms of these acids
in water. In these cases, n = 2. Figures for benzoic acid illustrate the
position ;
Phase Equilibria 177

Concentration Concentration
in water, in benzene, VC,
ce cs &

0.0145 0.252 0.252 _ 34


V0.252
Comes
0.0188 0375 v03750.375 3
oo1eg ~ 2°

0.0213 0.509 179:509 _ 335


0.0213

0.0275 0.817 0.817 _ 35


V0817
00275 7?
The figures in the final column are approximately constant.
lonization of the acid in water is neglected.

3. Solute in the form of a complex ion in the aqueous layer and in its normal
state in the non-aqueous layer
Partition may be used to determine the number of ammonia ligands
co-ordinating with a copper(II) ion in aqueous solution. If the aqueous
solution of the complex ion and excess ammonia is shaken with trichloro-
methane, equilibrium will be established and the total ammonia in each
layer may be determined by titration with acid. The ammonia in the
aqueous layer is made up of complex ammonia and free ammonia. If the
partition coefficient of ammonia between the two solvents is known, the
results of the analysis of the trichloromethane layer and the coefficient
value may be used to calculate the concentration of free ammonia in the
aqueous layer. Then by subtraction from the total ammonia in the aqueous
layer it is possible to determine the complexed ammonia and hence the
formula of the complex ion. 4
13. Energetics

All chemical reactions are accompanied by energy changes which are


usually observed as heat changes. The chemicals involved (known as the
system) may lose energy to the surroundings, such a reaction being known
as an exothermic reaction. When the system absorbs energy from the
surroundings it is known as an endothermic reaction.
The quantity of heat evolved or absorbed during a chemical reaction will
depend on the following factors.

1. The amounts of substances involved will affect the magnitude of the


heat change and it is usual to quote heat changes for specific molar
quantities of reactants or products.

2. The physical states of the reactants and products must also be specified,
as to convert a substance from one physicalstate to another will also
involve an energy change. It is usual to indicate the ‘state of each
substance by the appropriate letter placed in brackets after the formula:
(s) for solid, (1) for liquid and (g) for gas.

3. The temperature at which the experiment is carried out is either


specified or it is indicated that it is the standard temperature which is
taken as 298 K (room temperature, 25 °C).

4, The pressure at which the experiment is carried out is usually taken as


the standard atmospheric pressure, 101 300 N m~? (760 mmHg).

5. The magnitude of the heat change can also depend on whether the
experiment is carried out at constant pressure (represented by AH) or
at constant volume (represented by AU). It is usually more convenient
to carry out experiments at a constant pressure.

Standard Enthalpy Change


The heat change at constant pressure under the standard conditions of
298 K and 101 300 N m~? (760 mmHg) pressure is known as the standard
enthalpy change ‘and is represented by AH®.
Energetics 179
Sign convention
During an exothermic reaction energy is lost by the system and so the
products must be at a lower energy state than the reactants (Figure 13.1(a)).
The enthalpy change for such a reaction is said to be negative.

reactants

enthalpy seielpy
change

products

(a)

products

enthalpy shea!
change

reactants

(b)
Figure 13.1. (a) Exothermic reaction and
(b) endothermic reaction

For an endothermic reaction the products appear to be at a higher


energy state than the reactants and AH is positive as the system has
gained energy from the surroundings (Figure 13.1(b)).

Relationship Between AH and AU


Liquids and solids show so little volume change in connection with
thermochemical change that the issue of measurement at constant volume
and constant pressure does not arise to any significant extent. With gases,
180 Physical Chemistry
however, the case is different. If observations are carried out at constant
volume, no work can be done on, or by, the system. For an exothermic
reaction carried out at constant pressure, a diminution of volume means
that work is done on the system by the external pressure; that is, the heat
change appears greater than it would at constant volume by the thermal
equivalent of the work done. An increase of volume means that work is
done by the expanding system against the external pressure, so the heat
change appears less than it would at constant volume by the thermal
equivalent of the work done. In quantitative terms, the work done is
measured by the product pAV, where p is the constant pressure and AV the
change of volume. From the gas equation, pV = nRT, the work done is
measured by RT for every change of 1 mole unit of gaseous volume
(i.e. 22.4dm*), and is nRT for a change of n. R has the value of
8.3 JK~} mol~? in thermal terms and T is the absolute temperature of
the observations.
The relationship between AH and AU may be summarized by the equation
AH = AU + nRT

Consider the following example. 890 kJ are evolved when 1 mole of


methane is burned in oxygen under standard conditions (101 325 N m-?
and 298 K), i.e. AH = —890 kJ mol~?. How much heat would be evolved
if the same reaction were carried out at constant volume?
The equation for the reaction is .

CH,(g) + 202(g) > CO.(g) + 2H20(1)

The volume of the water can be considered as negligible and there is a


reduction from 3 moles of gas to 1 mole of gas, i.e. nm= —2. The quantity
of work done on the system when this contraction occurs is given by

NRL= — 2 8.3) x98


= —4947 J
—4.947 kJ
Using AH = AU + nRT
—890 = AU — 4.947
AU = —885 kJ mol7!

Experimental Determination of Enthalpy Changes


As previously indicated, when a value for an enthalpy change is quoted it
must be absolutely clear under what conditions such a value is obtained. It
is therefore necessary to have strict definitions for enthalpy terms which
are in common use.
The standard enthalpy change for a reaction, AH,$., is equal to the
enthalpy change when the molar quantities, represented by the equation,
Energetics 181
react at a pressure of 101 300 N m-? (760 mmHg) and a temperature of
298 K, the substances being in their normal physical states for these
conditions.

The standard enthalpy of formation, AH;,, is the enthalpy change which


occurs when | mole of a compound is formed from its elements, in their
normal states, under the standard conditions of pressure and temperature
(101 300 N m-? and 298 K).

The standard enthalpy of combustion, AH, is the enthalpy change when 1


mole of an element or compound is completely burned in oxygen under
standard conditions.

The enthalpy changes for reactions involving liquids are most easily
determined by insulating the reactants from the surroundings and record-
ing the temperature change of the reactants. If the reaction does not
involve organic solvents, a polystyrene beaker is a particularly convenient
container for such a determination as it is a good insulator and, due to its
small mass, it has a negligible heat capacity. The quantity of the heat
change is calculated by multiplying the temperature change by the mass of
the reactants and by their specific heat capacities. This value must then be
multiplied up for the required molar quantities.
The alternative approach is to conduct the heat away from the reactants
and to record the increase in temperature of the surroundings. This
method is particularly suitable for enthalpies of combustion.

Determination of heat of combustion


This determination is carried out in a bomb calorimeter, Figure 13.2. This
instrument was originally devised by Berthelot. A typical example of it is
made of steel. It is nickel plated on the outside and protected on the inside
by a coating of non-oxidizable material (enamel or, better, gold or
platinum). It has a tight-fitting, screw lid, working on to a lead washer.
Material is burnt in a small platinum cup, C, in the calorimeter. Insulated
platinum leads allow electric current to be passed through a thin spiral of
iron wire, W, which glows and then fires the material.
A known mass of the material under investigation is placed in the
platinum cup. Air is displaced by oxygen, which is allowed to reach a
pressure of 20-25 atmospheres. The bomb calorimeter is then closed by a
screw valve, and placed in water in a calorimeter fitted with a stirrer and an
accurate thermometer, and well lagged to minimize heat losses. With the
stirrer operating, temperature readings are taken at regular intervals of
time. Current is then passed. The iron wire becomes red hot and fires the
material, and combustion occurs. This brings a sudden rise of temperature,
182 Physical Chemistry

Figure 13.2. Bomb calorimeter

temperature

time
Figure 13.3.

which is registered on the temperature-time curve, Figure 13.3. Observa-


tions are continued until the system is cooling slowly and regularly, as in
the part of the curve CD. If E is the mid-point of BC and EF is parallel to
the temperature axis, the highest temperature reached can be taken as F
(or T). The heat capacity of the whole system—bomb calorimeter, water,
calorimeter, thermometer, stirrer, etc—must be known.
Then, if the total heat capacity of the system is W, and the rise of
temperature is ¢ K, the total evolution of heat is Wt J. If the iron wire in
Energetics 183
burning produces x J, the heat evolved in the combustion of the material
is (Wt — x) J. Then, if the mass of the material used was a g, and its
molecular mass M, the heat of combustion of the material is

aM — x)

Hess’s Law and Energy Cycles


This law is the essential basis of calculations in thermochemistry. It has
both experimental and theoretical justification, as shown below. Hess’s
Law takes the form:

The total energy change resulting from a chemical reaction is dependent


only on the initial and final states of the reactants and is independent of
the reaction route.

This law is an aspect of the general Law of Conservation of Energy or


the First Law of Thermodynamics.
Suppose that a given chemical system changes from the state A to the
state B by way of intermediate stages C and D, with liberation of x, y, and
z J respectively in these stages; let the system also change from A to B by
a single intermediate stage, E, with liberation of a and b J as follows;

S iv
ee = E

Hess’s Law requires that (a + b) = (x + y + 2). If not, let (@ + 5) be


greater than (x + y + z). It would then be possible to pass from A to B by
E and gain (a + b) J then reverse the process by D and C, giving back
(x + y +z) J. This would give a net gain on the circuit of (a + b) —
(x + y +z) J. This could be repeated indefinitely, yielding an unlimited
supply of energy from no corresponding source. This is contrary to the
Law of Conservation of Energy. Consequently, Hess’s Law must be true.
Experimental illustration of Hess’s Law is obtained by performing the
same chemical change in two alternative ways and showing the heat
change to be the same for both. The following is an example of this:
Conversion of 56 g of calcium oxide, CaO, into a dilute solution of the
corresponding mass (111 g) of calcium chloride by methods A and B.
184 Physical Chemistry
Method A Method B
Convert the CaO to Ca(OH)z. Add the CaO capes a ae*
Ca0(s) + H,O() > CAH) i
appropriate mass of dilu te :
H® = —63kJ CaO(s) + 2HCl(aq) >
: CaCl H,O(
Convert the Ca(OH), to a dilute aGlalad) nNbe ==1975k5
solution.

Ca(OH),(s) + Bey — Ca(OH),(aq)


AH® = —12.5kJ
Neutralize this with the correspond-
ing amount of dilute HCl.
Ca(OH).(aq) + 2HCl(aq) >
CaCl,(aq) + 2Hz0(1)
Nek ERS
Total enthalpy change = Total enthalpy change =
(—63 —12.5 —117) — 192.5 kJ
= —192.5 kJ

The above example may be more conveniently CE in the form of an


energy cycle.
t

AH,e
CaO(s) + 2HCl(aq) ——————-——> CaCl,(aq) + H2O())
AH§| +H,0() AH | +2HCi(aa)
Ca(OH),(s) 7 Sew cak ch2 Ca(OH)2(aq)
AH? = AH? + AH2 + AHS
Indirect Determination of Enthalpy Changes
The enthalpy changes for some reactions cannot be determined directly by
experiment, but, by the application of Hess’s Law, they can be calculated
from other experimental results. For example, it is possible to determine
experimentally the enthalpies of combusion AH of methane, carbon, and
hydrogen and these results may then be used to calculate the enthalpy of
formation of methane which cannot be determined directly.

Given that AH? for methane is —890 kJ mol-},


AH for carbon is —393 kJ mol~},
AHS for hydrogen is —286 kJ mol~},
Energetics 185
the equations associated with the above enthalpy changes are
CH,(g) + 202(g) — CO.(g) + 2H,O(1) AH® = —890 kJ mol-?
C (graphite) + O.(g) > CO,(g) AH*® = —393 kJ mol7?
H.(g) + 30.(g) > H,O() AH® = —286 kJ mol7}
The equation for the formation of methane from its elements is

C (graphite) + 2H.(g) > CH,(g)


The following energy cycle can be constructed:

Crepe) ee
oH CH Ig)
fee

CO.(g) + 2H,0(1)

According to Hess’s Law the total enthalpy change during the formation
of the combustion products by one route is equal to the total change when
they are formed by another route, i.e.
AH? + AH? = AHY (1)
where AH? = AZH¢@ the standard enthalpy of formation of methane,
AH? is the standard enthalpy of combustion of methane,
AH? is the enthalpy of combustion of carbon + 2 x the enthalpy
of combustion of hydrogen.
Substituting into equation (I),
AHy + (—890) = —393 + [2 x (—286)]
AH? for methane = AH? = —75 kJ mol-?

Standard Enthalpy of Atomization and Bond Energy Terms


The standard enthalpy of atomization of an element is the heat required to
convert the element, in its normal state and under standard conditions,
into 1 mole of atoms in the gas phase.
Note that enthalpies of atomization are quoted for 1 mole of gaseous
atoms produced and not for 1 mole of the element in its normal state. The
following example will emphasize this point.
Consider the conversion of hydrogen molecules to atoms.
H.(g) > 2H(g) AH? = + 436kJ
186 Physical Chemistry
436 kJ mol-! is known as the mean H—H bond energy, as it is the
energy required to break 1 mole of H—H bonds. However, the standard
enthalpy of atomization of hydrogen is 436/2 = +218 kJ mol~* as this is
the energy required to produce 1 mole of atoms in the gas phase.
Enthalpies of atomization can be used to calculate mean bond energies
for bonds involving unlike atoms.

Mean C—H bond energy


Given that the standard enthalpy of formation of methane is —74.8 kJ mol-*
and the enthalpies of atomization of graphite and hydrogen are +715
and +218kJ per mole of atoms in the gas phase, the following energy
cycle can be constructed.

, AH?
C (graphite) + 2H.(g) —————————_——>> CH, (g)
a
AHS AH
C(g) + 4H(g)
By Hess’s Law,
AH? = AH? + AHP
where AH? is the standard enthalpy of formation of methane,
AH, is the enthalpy of atomization of graphite + (4 x the
enthalpy of atomization of hydrogen),
AH? is the enthalpy of formation of 1 mole of methane from
gaseous atoms of carbon and hydrogen.
Substituting,
—74.8 = 715 + (4 x 218) + AH?
AH? = — 1661.8 kJ mol-!
This is the energy evolved during the formation of 4 moles of C—H
bonds from gaseous atoms of carbon and hydrogen. Therefore the mean
C—H bond energy is
—1661.8
m = —415.4 kJ mol7?

Mean C—C bond energy


The standard enthalpy of formation of ethane is —84.6 kJ mol-}. Using
this value and the standard enthalpies of atomization of hydrogen and
Energetics 187
graphite from the previous example, it is possible to calculate the mean
C—C bond energy from the following energy cycle.

AH?
2C (graphite) + 3H2(g) —————————+ C.H,(g)
AH? AHs
2C(g) + 6H(g)
By Hess’s Law,
AH? = AH? + AH®
where AH? is the standard enthalpy of formation of ethane,
AHP is (2 x standard enthalpy of atomization of graphite) +
(6 x standard enthalpy of atomization of hydrogen),
AH? is the standard enthalpy of formation of ethane from gaseous
atoms of carbon and hydrogen.

Substituting,

—84.6 = (2 x 715) + (6 x 218) + AH


AH® = —2822.6 kJ mol-?
This is the energy evolved during the formation of 1 mole of C—C bonds
and 6 moles of C—H bonds, i.e.
—2822.6 = 1(C—C bond energy) + (6 x —415)
This gives a value for the mean C—C bond energy of —333 kJ mol™?.

By calculations similar to those employed above it is possible to build up


a table of mean bond energies. These bond energies strictly refer to the
bond in a particular molecular environment (e.g. the C—H in methane as
opposed to the C—H in trichloromethane) and care must be exercised when
using them as it is common for bond strengths to be affected by the rest of
an organic molecule. Despite these limitations, as will be seen in the
following section, bond energy terms can be very useful.

Prediction of Enthalpy Changes from Bond Energy Terms


The enthalpy change for a reaction may be considered as the difference
between the sum of the energies required to break bonds in the reactants
and the sum of the energies evolved when new bonds are formed to give the
products. Clearly a knowledge of bond energy terms makes it possible to
predict enthalpy changes for reactions, e.g. assuming the following mean
188 Physical Chemistry
bond energies it is possible to predict a value for the standard enthalpy of
hydrogenation of ethene (ethylene).
C—C = 345 kJ mol=?
C=C = 610 kJ mol=?
C—H = 415 kJ mol-?
H—H = 435 kJ mol?
The equation for the hydrogenation of ethene is

sat yA: ii i
hae (g) + Ha (g) > ag fora(g)
H H H H
This change could be considered to take place by the breaking of 1 mole
of C=C bonds and 1 mole of H—H bonds which would require:
610 + 435 = 1045 kJ
This is then followed by the forming of 1 mole of C—C bonds and 2 moles
of C—H bonds. The heat evolved during the formation of these bonds
would be
345 + (2 x 415) = 1175 kJ
The standard enthalpy of hydrogenation is likely to be
1045—1175 = —130 kJ mol-*

Delocalization Energy :
Using the principles employed in the previous section it is possible to
predict a value for the standard enthalpy of hydrogenation of benzene if
its structure is assumed to consist of alternate double and single carbon-
carbon bonds. Using the following mean bond energies:
C=C = 610 kJ mol-?
C—C = 345 kJ mol7?
C—H = 415 kJ mol=}
H—H = 435kJ mol=}
The reaction may be represented by the equation
H H H
| rere

Ri
aN-C—H eee
H—c C—H
ae eae 3H, (2)>
H, (g) ox) dyee (1)
cf yA Nees Se
| aey
H H H
Energetics 189
This reaction could be considered to involve the breaking of 3 moles of
C=C bonds and 3 moles of H—H bonds which would require
(3 x 610) + (3 x 435) = 3135 kJ
This is then followed by the forming of 3 moles of C—C bonds and 6 moles
of C—H bonds. The heat evolved during the formation of these bonds
would be
(3 x 345) + (6 x 415) = 3525 kJ
This gives a value of
3135 — 3525 = —390 kJ mol=?
for the standard enthalpy of hydrogenation of benzene.
The experimentally determined value is approximately —210 kJ mol7?.
The difference between these two values indicates that the actual structure
of benzene is more stable than the alternate double and single bond
structure by a factor equivalent to (390-210) = 180 kJ mol-+. This extra
stability is attributed to the delocalization of the bonding electrons over
all six carbon atoms (see page 83).

Born-Haber Cycle and Lattice Energies


The Born-Haber energy cycle considers the energy changes which occur
during the formation of an ionic crystal. The Born-Haber cycle for the
formation of sodium chloride is given below.
AH?
Na(s) + $Cl.(g) ——————__——> NaCC|(s)
AH, AHS Aue
Na(g) + Cl(g) ano Cl-(g) + Na*(g)
AH?

Starting with the elements in their standard states, the steps in the cycle and
the enthalpy changes associated with them are:
1. The atomization of the elements to form gaseous atoms. These changes
require energy and so the values for AH? and AH? will be positive.
Remember standard enthalpies of atomization are quoted for 1 mole
of gaseous atoms and it is not necessary to halve the chlorine value.
2. The gaseous metal atom is converted into a positive ion. This step needs
energy and AH?, which is the ionization energy for sodium, has a
positive value.
3. The gaseous non-metal atom forms a negative ion. This usually results
in the evolution of heat (O~ to O?~ being an exception) and AHP,
which is known as the electron affinity of chlorine, is negative.
190 ...° Physical Chemistry
4. The gaseous ions come together to form the crystal lattice. This step
will be accompanied by evolution of heat and AH?, which is the lattice
energy, will be negative.
The lattice energy is defined as the standard enthalpy of formation of
1 mole of the crystal lattice from its ions in the gas phase.
The other half of the cycle is AH?, which is the standard enthalpy of
- formation of the sodium chloride crystal, i.e. the enthalpy change which
occurs when 1 mole of the compound is formed from its elements in their
standard states.
Applying Hess’s Law to the cycle:
AH? = AH? + AH? + AH? + AHY + AHP
All of these enthalpy changes except AH? can be determined either
directly or indirectly from experimental results and hence the cycle can be
used to calculate a value for the lattice energy (AH).
In the case of sodium chloride,
AH? = AH? of NaCl (s) = —411 kJ mol-?
AH? = standard enthalpy of atomization of sodium = +108 kJ mol-?
AH? = standard enthalpy of atomization of chlorine = +121 kJ mol~+
AH? = first ionization energy of sodium = +493 kJ mol?
AH? = electron affinity of chlorine —364 kJ mol-?
AH? = lattice energy for sodium chloride. t

Na*(g) + Cl(g)

+493 kJ
—364 kJ

Na*(g) + CI-(g)

Na(g) + Cl(g)
energy
—770 kJ
+229 kJ.

Na(s) + 3 Cl,(g)

—ANkJ

NaCl(s)
Figure 13.4.
Energetics
Substituting,
—411 = 108 + 121 + 493 —
AHS = —769kJ mol“? f >
The lattice energy of sodium chloride is 769 kJ mol. It is worths it
while considering this energy cycle in the form of an-energy level diagram;
Figure 13.4. {oo VY a 4
If the ionic crystal is to be energetically stable with respect tovits,’
constituent elements in their standard states, then AH must be negative.
It is clear from Figure 13.4 that the main energy terms which determine the
sign of AH? are the ionization energy of the metal atom and thelattice
energy. It is mainly the relative magnitude of these two values which will
determine whether or not the compound is formed exothermally from its
elements in their standard states.

Theoretical Lattice Energies


Lattice energies determined by the Born-Haber cycle are based directly or
indirectly on experimental results. They may be considered as the energy
released when the ions in the gas phase come together to form 1 mole of
the crystal, irrespective of the type of bonding in the resulting crystal.
In other words, lattice energies can be calculated from the Born-Haber
cycle even if the resulting lattice is not 100 per cent ionic and does involve
some electron sharing.
If the lattice is considered to be 100 per cent ionic then it is possible to
calculate a theoretical value for the lattice energy. This calculation must
take into account the magnitude of the charges, their coming together from
infinity, and therefore the variation of forces of attraction with distance of
separation (inverse square law). In addition, using the geometry of the
crystal, it is necessary to take into account both attractions between
oppositely charged ions and repulsions between ions with like charges.
In the cases of salts such as sodium chloride there is good agreement
between the theoretical and experimental values, indicating the validity of
the assumption that they are 100 per cent ionic. Where there is an obvious
discrepancy, as in the cases of silver chloride and zinc sulphide, it suggests
that the bonding in the lattice is not entirely ionic and there is appreciable
sharing of electrons (covalent bonding).

Standard Enthalpy of Solution and Standard Enthalpy of Hydration


The dissolving of an ionic compound in water can be either an exothermic
or an endothermic process, which suggests that there are at least two energy
terms involved. In fact, from the energy point of view, the process can
be thought of as occurring in two stages. The first is the separation of the
ions which will require energy equal in magnitude but opposite in sign to
192 Physical Chemistry
the lattice energy. The second stage is the hydration of the ions, which will
be an exothermic process.
If the lattice energy is greater than the enthalpy of hydration, the
enthalpy change associated with the dissolving of the compound to form
a dilute solution will be endothermic. If the enthalpy of hydration is the
greater, then the dissolving process will be exothermic.
For sodium chloride the following energy cycle may be drawn:

AH? Mt
NaCl(s) —————————— Na
*(aq) + Cl (aq)

AHS AH
Nat (g) + Cl-(g)
By Hess’s Law,
AH? = AH? + AH?
where AH? = the standard enthalpy of solution,
AH?2 = — (lattice energy),
AH? = the sum of the enthalpies of hydration of the two ions.
The standard enthalpy of solution of sodium chloride is +4 kJ mol? and
the lattice energy is —770 kJ mol~?. This means the enthalpy of hydration
is —766 kJ mol7}. ::
The balance between lattice energies and enthalpies of hydration accounts
for the observation that certain hydrated compounds dissolve endotherm- —
ally whereas the corresponding anhydrous compounds dissolve exother-
mally. The ions in the crystals are already partially hydrated so that when
they are dissolved the lattice energy easily outweighs the enthalpy change
associated with any further hydration. For example, anhydrous copper(II)
sulphate dissolves to form a dilute solution with the evolution of
66.3 kJ mol~*, whereas 1 mole of the hydrated copper(II) sulphate
absorbs 11.4 kJ mol~? when it is dissolved to form a dilute solution. The
following energy cycle may be constructed;

CuSO,(s) ——————> Cu?* (aq) + $02-(aq)

aoe . 5H,0(s)
From this it is possible to calculate that the enthalpy change associated
with the formation of the hydrated crystals from the anhydrous salt is
— 66.3 — 11.4 = —77.7kJ mol-!
ie. CuSO,(s) + SH,O() + CuSO,,5H,O(s) AH® = —77.7kJ mol7?
Energetics 193
Endothermic Reactions, Entropy, and Free Energy
When a reaction, from the overall energy change point of view, is feasible,
it is sometimes described as being energetically favourable. An exothermic
reaction would appear to be energetically favourable as the system is losing
heat energy to the surroundings and hence is moving to a lower energy
state. However, this approach raises the question - Why do endothermic
reactions occur? The existence of endothermic reactions suggests that
enthalpy changes are not the sole factor to be considered. The other
factor is given the name entropy and the change in entropy, AS, represents
the change in the degree of disorder. It appears to be the natural tendency
of systems to move to a more disordered or chaotic state. For example, if
marbles of two different colours are shaken in a tray, one would expect a
random, chaotic distribution of the two colours (high entropy) to result
rather than the more ordered state of marbles of one colour in one half of
the tray and marbles of the other colour in the other half (low entropy).
Similarly, if two gases which do not react are placed in a container one
would expect the gases to mix and achieve the highest state of entropy. In
the context of a chemical reaction, if, for example, the number of moles of
gas increases as the reaction proceeds, there will be an increase in entropy. In
general the greater the number of particles which can move independently,
the greater the entropy.
The overall term which incorporates both enthalpy change and entropy
change is the change in Gibb’s free energy. The standard free energy
change, AG®, is related to the standard enthalpy change and the standard
entropy change by the equation
AG? = AH® — TAS®
Thus the overall factor which determines whether or not a reaction is
energetically feasible is the change in free energy. A negative free energy
change for the complete reaction indicates an energetically favourable
reaction. If the entropy change is small, for example when two moles of
liquid produce two moles of two other, but fairly similar, liquids, then the
enthalpy change may provide a good indication of the feasibility of the
reaction, but it must be noted that the importance of the entropy change (as
it is multiplied by the temperature on the Kelvin scale) increases with
temperature.
Also it is important to note that overall free energy changes indicate
energetically favourable reactions but give no indication of the rate of the
reaction (see page 256).
14. Chemical Equilibria I:
Molecular Equilibria

Reversible Reactions and the State of Dynamic Equilibrium


A reversible process, whether physical or chemical, in a closed system
(i.e. one which will not allow any of the components to escape) and at a
constant temperature, will eventually reach a state of equilibrium. The
equilibrium will be dynamic and may be considered as a state at which the
rate of the process in one direction (see page 254) is exactly balanced by the
rate in the reverse direction. When a chemical reaction reaches this state of
dynamic equilibrium, the equilibrium concentrations of the reactants and
products will remain constant as long as the conditions are not changed.
Consider the homogeneous system
wA + xB=yC+zD
then, when the system is in a state of equilibrium, the equilibrium molar
concentrations of the reactants and products are found to be related in the
following way,
[CP DI Ly
uy 2

[A]” [B]?
K., for a fixed temperature, is found to have a constant value and is
known as the equilibrium constant. The square brackets indicate equilibrium
concentrations measured in mol dm~ °. The relationship itself is known as
the Equilibrium Law.

Application of the Equilibrium Law to Homogeneous Molecular


Equilibria
1. The reaction between ethanol and ethanoic acid
The reaction is represented by the equation

CH;CO,H()) + C,H;OH() = CH;CO.C.H,(l) + H,0()


Suppose that a moles of ethanoic acid and b moles of the alcohol are
mixed and allowed to come to equilibrium at constant temperature, so
that, at equilibrium, x moles of ethyl ethanoate and water are formed. Then
Chemical Equilibria I: Molecular Equilibria 195
(a — x) moles of acid and (b — x) moles of alcohol must remain. If the
volume of the liquid is V dm3, the molar concentrations of the reagents are

ethanoic acid: er»)

alchohol: Vee)
V

ester and water: each .

It is advisable to employ a systematic method of setting out this informa-


tion.
CH;CO.H(I) + C,.H;OH(1) = CH;CO.C-H;(1) + H,Od)

Initial
amounts (mol) a b 0 0
Equilibrium
amounts (mol) a-x b—x x x
Equilibrium
concentrations
(mol dm-°) a-—x b-—x
N N Se &
NI
The equilibrium law for this reaction is

_ [CHsCO,H] [C2H;OH]
* [CHsCO2CHs] [H20]
Substituting,
“— x2 /V2

° @— x6 — xP?

‘a (a — x)(b — x)
A value for K, for this reaction at a particular temperature may be deter-
mined by the following procedure.
Ethanoic acid and ethanol, both in a pure and dry state, are mixed to
make a known volume, in the proportions of 1 mole of each (that is,
60 g of ethanoic acid to 46 g of ethanol). The containing flask is corked
and kept in a thermostat at 298 K. At intervals, 2 cm® of the liquid are
extracted and titrated at once by barium hydroxide solution (or carbonate-
free sodium hydroxide solution) with phenolphthalein as indicator. This
196 Physical Chemistry
determines the ethanoic acid in the mixture. It will be found that the
proportion of ethanoic acid present falls as esterification proceeds, finally
reaching a constant minimum when equilibrium is attained. At this point,
experiment shows that one-third of the original mole of ethanoic acid
remains, i.e. two-thirds have been esterified. Applying these figures to the
above equation, a = b = 1 and x = ; that is,

=a
From these figures, the equilibrium constant of the reaction at 298 K is 4.
The equilibrium law states that for this particular temperature K, for the
ethanoic acid-ethanol reaction will always be 4. The implication of this is
that if other initial concentrations of reactants are used the position of the
dynamic equilibrium finally reached will be such that the equilibrium
concentrations, when substituted into the equilibrium law, will still give a
value of 4 for K,. Suppose, now, that a mixture of 1 mole of ethanoic acid
and 0.5 moles of ethanol is used. If the equilibrium law is correct, and if
equilibrium is attained when n moles of ester (and water) have been
formed,
n2
K,=4
G=n05—7”)
This is a quadratic in n, which can be solved for nas follows:
r?= 4(1 — n)(0:5 — n) .
= 4(0.5 — 1.5” + n°)
= 2 — 6n + 4n?
i.e.
3n? — 6n +2=0
From this,
na Ot VG6 — 4 x 3 x 2)
6
_6+ yi2
6
_6+2y3
6
_6—2x 1.732
6
= 0.42
; (The alternative value of n, taking the positive value of 24/3, gives an
impossible answer.) From this, equilibrium should be reached with
Chemical Equilibria I: Molecular Equilibria 197
0.42 moles of ester and water, (1 — 0.42) or 0.58 moles of ethanoic acid
and (0.50 — 0.42), or 0.08 moles of ethanol present. This can be tested by
performing the actual experiment with initial proportions of 1 mole of
ethanoic acid and 0.5 moles of ethanol. The order of agreement reached
in experiments by Berthelot and St Gilles is shown in the table below.
The good agreement between the observed and calculated values for the
ester justifies acceptance of the Equilibrium Law for this reaction.

Original proportions Ester produced at equilibrium


(moles) (moles)
Ethanoic acid Ethanol | Observed by expt. Calculated from K, = 4

1.00 0.080 0.078 0.078


1.00 0.280 0.226 0.232
1.00 0.330 0.293 0.311
1.00 0.500 0.414 0.423
1.00 0.670 0.519 0.528
1.00 1.500 0.819 0.785
1.00 2.000 0.858 0.845

2. The reaction between hydrogen and iodine


The classical example of the application of the Equilibrium Law to a
reaction involving entirely gaseous reagents is Bodenstein’s investigation
of the synthesis of hydrogen iodide from its elements:
H.(g) + I.(g) = 2HI(g)
There is no change of total volume in this reaction, so that the results are
independent of pressure (a fact which was first verified by Bodenstein). In
the experimental work, he took a known mass of iodine (say 0.319 moles)
and placed it in a glass bulb which was then sealed containing a known
mass (say 0.206 moles) of hydrogen, actually under considerable pressure,
which is convenient because of the very low density of the gas. The bulb
was then heated at a constant temperature in the region of 723 K till
equilibrium was attained. The bulb was then cooled rapidly to room
temperature. The time of cooling was so short that no appreciable change
could take place in the equilibrium as the temperature was falling; at room
temperature, the velocity of reaction is so slow that the mixture can be
kept indefinitely without appreciable change. This device of rapid cooling
to a temperature at which reaction velocity is negligible is called freezing
the equilibrium and is often resorted to on both the laboratory and
industrial scales. In the present case, the ‘frozen’ equilibrium mixture was
analysed for iodine and hydrogen iodide by absorption in excess of
198 Physical Chemistry
standard potassium hydroxide solution and the application of the usual
methods of analysis. The iodine was found to be 0.134 mole and the
hydrogen iodide 0.370 mole; from these figures, the hydrogen present at
equilibrium was
0.206 — (0.319 — 0.134) = 0.021 moles
Applying the Equilibrium Law,
H.(g) + I.(g) = 2HI(g)
Initially: a b — moles
At equilibrium: (a—x) (b—x) 2x moles
If the volume at equilibrium is V dm°,

ee
K, ~ [@—x
KOS ns
VIG —x/V] (@—x)6—)
Notice that K, is independent of V (and, therefore, of pressure).
Applying the above figures of experiment,
(0.370)?
as eee 87
Ke = Toni x 0.134
By using different proportions of hydrogen and iodine, other values of
K, can be calculated and should, if the Equilibrium Law is valid, be
constant. For Bodenstein’s figures, K, shows some variation; but, in
view of the experimental difficulties involved, the results are considered
reasonably satisfactory (K, = 54 +9 approximately). The following
calculation further illustrates the operation of the Equilibrium Law in
this case.

Example In an experiment, 0.206 moles of hydrogen and 0.144 moles of


iodine were heated (at 723 K) to equilibrium in the reaction Hz, + I, =
2HI. 0.258 mole of hydrogen iodide was formed. Calculate the equilibrium
constant of the reaction to the nearest integer. Hence calculate the molar
composition (at 723 K) of the equilibrium mixture obtained from heating
0.515 moles of hydrogen with 0.360 moles of iodine.

H.(g) + I.(g) = 2HI(g)


As 0.258 moles of hydrogen iodide were formed, 0.258/2 = 0.129 moles of
both hydrogen and iodine must have been used up. The iodine in the
equilibrium mixture must, therefore, be (0.144 — 0.129) = 0.015 moles,
and the hydrogen (0.206 — 0.129) = 0.077 moles.
joe ETP 2
00.25 tinephcg
2

[H2][I2] 0.077 x 0.015


Chemical Equilibria I: Molecular Equilibria 199
In the second part of the calculation, let there be 2x moles of hydrogen
iodide produced at equilibrium. Then the number of moles of iodine in
the equilibrium mixture is (0.360 — x) and of hydrogen (0.515 — x)
From this,

58 3 (2x)?
(0.515 — x)(0.360 — x)
4x? = 58(0.515 — x)(0.360 — x)
13.5x? — 12.7x + 2.68 = 0
x = 27+ VIC2.7? — 4 x 13.5 x 2.68]
2 XA3.5
= 0.32
From this, the equilibrium mixture is
Hydrogen iodide 0.64 moles
Iodine (0.360 — 0.32) = 0.04 moles
Hydrogen (0.515 — 0.32) = 0.195 moles

3. Dissociation of phosphorus pentachloride (phosphorus(V) chloride)


Phosphorus pentachloride dissociates according to the equation
PCI,(g) = PCla(g) + Cla(g)
Suppose that a mole of the pentachloride are originally present and
dissociate to form b mole each of PCI; and Cl,, and to leave (a — b) mole
of PCI;, all gaseous and in a volume V dm® at equilibrium at some definite
temperature, t K. Applying the Equilibrium Law,
7 reIV DIKshee008?
~~ (a=b|[Ve- (a— bv
In this case, K, involves V and so is not independent of pressure. If V
increases, K, must remain constant at constant temperature; consequently
b? must increase and (a — b) decrease to maintain the value of K,. That is,
increase of volume increases the dissociation of the pentachloride. This is the
same as the assertion that decrease of pressure (at constant temperature)
increases the dissociation, and vice versa.
Suppose that, to the above system in equilibrium, one of the products,
say chlorine, is added with no change of volume or temperature. If the
concentration ofchlorine is raised to (b + c) moles in V dm®,,this quantity
must enter the numerator of the value of K, above. To maintain the
value of K,, the b factor for PCl; must decrease and the (a — b) factor
for PCI, increase, i.e. the introduction of chlorine tends to suppress the
dissociation of the pentachloride. The same is true if phosphorus trichloride
is added in similar conditions.
200 Physical Chemistry
When all the materials concerned are gaseous, their concentrations can
be expressed in terms of pressure. For example, in the dissociation of phos-
phorus pentachloride:

PCl,(g) = PCla(g) + Cla(g)


partial pressures at equilibrium: P2 Ps

the equilibrium constant can have the alternative representations

_ PCIe} [Cle]
; [PCl,]
Ie, — P2 x P3

P1

The sum (p, + pz + ps) is the total pressure of the mixed vapours.
In terms of the total pressure, p, of the system, the partial pressure of a
reactant is given by p X the mole fraction of the reactant present at
equilibrium.
In the case of the dissociation of phosphorus pentachloride, the total
number of moles present at equilibrium is

(a—b)+b+b=(a+b) ;

Therefore the partial pressures are given by

Pliny psn.
(a
+ 6)

(a + b)

, (a + b)
Substituting into the K, form of the Equilibrium Law,

_ _pb?|(a + b)?
2 p(a — b)[(a + b)
eS sive Pbiy Nis
(a + b\(a — b)
This expression can be used to make quantitative predictions of equilibrium
concentrations for various total pressures.
Chemical Equilibria I: Molecular Equilibria 201
4. The equilibrium in the Haber synthesis of ammonia
In the Haber process, ammonia is synthesized from its elements

N.(g) + 3H.(g) = 2NH,(g)


Initially: 1 3 sureties

At equilibrium: (a— x) (6 — 3x) 2x mole


Pt Pa Ps partial pressures
From this,

cates alist, sein:


teefNa} Ha}? 2127? py: xpe
As in the above examples, K, can be expressed in terms of a, b, x, and the
total volume of the system, V. K, can be expressed in terms of a, b, x, and
Pp, the total pressure of the system.

Application of the Equilibrium Law to Heterogeneous Systems


In a general way, the Equilibrium Law can be applied only in homogeneous
systems. With some appropriate assumptions, however, it can be applied
to certain heterogeneous systems. Some of these are mentioned below.

1. Action of heat on calcium carbonate


When heated, calcium carbonate dissociates in the reversible reaction
CaCO,(s) = CaO(s) + CO.(g)

If the assumption is made that the solids CaCO; and CaO have definite
vapour pressures and so participate in the equilibrium in the vapour phase
(which is homogeneous), the Equilibrium Law can be applied, to the vapour
phase, in the form
(partial pressure of CaO) x (partial pressure of CO.)
= a constant
(partial pressure of CaCOs3)
The further assumption is then made that, so Jong as any of the solid is
present at all, no matter how little, its vapour pressure (at a given tempera-
ture) can be taken as constant. Consequently, for a system in which
CaCO, (solid), CaO (solid), and CO, are all present together in equilibrium,
the partial pressures of both CaO and CaCO, can be taken as constant
and the above expression becomes:
(partial pressure of CO.) = a constant = K,
This means that, by a deduction from the Equilibrium Law applied in
this special way, we should expect that if calcium carbonate is heated in an
202 Physical Chemistry
originally evacuated container so that CaCO, CaO, and CO, are always
present together, there is a definite pressure of carbon dioxide, (known as
the dissociation pressure for calcium carbonate) for each temperature at
which equilibrium is reached. This is, in fact, the case. Up to about
873 K, this equilibrium pressure of carbon dioxide is quite minute and the
rate of decomposition of calcium carbonate is so slow (at atmospheric
pressure) as to be negligible. At bright red heat, the equilibrium pressure
of carbon dioxide is about one atmosphere and decomposition of calcium
carbonate is rapid.
In a closed vessel of such a size that the equilibrium pressure of carbon
dioxide can be reached with some calcium carbonate still left, complete
decomposition of calcium carbonate cannot be achieved. In the open air,
the partial pressure of carbon dioxide is negligibly small, the reverse action
is negligible, and virtually complete decomposition of the carbonate is
achieved.

2. Reaction between iron and steam


At red heat, iron reacts with steam, in a reversible reaction, to form its
black oxide, iron(II) diiron(III) oxide, and hydrogen.
3Fe(s) + 4H2O0(g) = Fe,0.(s) + 4H.(g)
Assuming, as in case 1, that the solids Fe and FesO, have definite vapour
pressures and participate in the equilibrium in the vapour phase, the
Equilibrium Law can be applied to the vapour phase in the form
(partial pressure of FezO,) x (partial pressure of H.)*
(partial pressure of Fe)® x (partial pressure of H,O)* Be
Assuming again that so long as any of the solid is present its vapour
pressure is constant, the partial pressures of Fe and Fe3O, can be taken as
constant in a system in equilibrium (at constant temperature) in which
all four reagents are present together. The above expression can then be
written
(partial pressure of H,)* _ 4 K
(partial pressure of H,O)* — a constant
or

(partial pressure of H.)


(partial pressure of H,O) = a constant (Kj)

This means that, if iron and steam are heated together in an initially
evacuated vessel of such a size that equilibrium can be reached with Fe,
H,O, Fe,O,, and Hg all present together, there is a definite total equili-
brium pressure for each temperature and it is made up of steam pressure
and hydrogen pressure in a ratio characteristic of that temperature, and
Chemical Equilibria I: Molecular Equilibria 203
constant for it. Also, an exactly similar equilibrium position will be
reached, for the same temperature, starting from the materials Fe,O,
and Hg, similar conditions applying.
If, however, steam is passed over iron (at red heat) in a continuous
stream in an open tube, hydrogen is continually swept away by steam
pressure as fast as it forms. Its effective concentration (or partial pressure)
tends to zero, the reverse action is negligible and iron can, for practical
purposes, be fully converted to its black oxide. (Notice, however, that the
conversion of steam to hydrogen is far from complete.) If hydrogen is
similarly passed over black iron oxide at red heat, the relations just stated
for hydrogen and steam are reversed and the black oxide can be reduced,
virtually quantitatively, to iron.

Qualitative Illustrations of Reversible Reactions and Chemical


Equilibria
A reversible reaction may be defined as a reaction which will proceed in
either direction if conditions are arranged appropriately.
The following are examples of reversible reactions which can be used to
give qualitative illustration of chemical equilibrium.

1. The bismuth chloride reaction


BiCl,(aq) + H,O() = BiOC\(s) + 2HCl(aq)
If bismuth carbonate is treated carefully with dilute hydrochloric acid
with shaking, it can be converted to a colourless liquid, bismuth chloride
solution, containing a certain excess of hydrochloric acid.
If this solution is then poured into a moderate excess of water and stirred,
a white precipitate will appear. This is bismuth oxychloride. In this case,
the excess of water has pushed the equilibrium to the right, producing
enough oxychloride to cause precipitation. asd
If the ‘milky’ liquid is then stirred while concentrated hydrochloric acid
is added drop by drop, the white precipitate will eventually dissolve. In
this case, the increased concentration of acid has pushed the equilibrium to
the /eft (in the above equation), producing soluble bismuth chloride. By
suitable alternate additions of water and acid the reaction can usually be
reversed several times, though at greater dilutions the precipitate may be
slow to appear.

2. The thiocyanate reaction


The reaction between iron(II) chloride and ammonium thiocyanate (both
in solution) is represented by
Fe®+(aq) + CNS~(aq) = [Fe(CNS)}?*(aq))
204 Physical Chemistry
[Fe(CNS)]?* is very strongly coloured (blood-red) and the equilibrium
position can be followed by observation of this colour.
Prepare a solution of ammonium thiocyanate (1g in 100 cm*) and of
iron(II) chloride (1g of the hydrate, FeCl;.6H,O, in water, 95 cm®
and concentrated HCl, 5 cm). Add 2.5 cm® of each solution to 250 cm
of water in each of four beakers and stir; this should give a solution of a
suitable intensity of red colour.
Keep one of the beakers for comparison of colour, adding 50 cm® of
water. To the others, add:
To beaker (a), 50 cm® of the iron(II) chloride solution.
To beaker (b), 50 cm® of the ammonium thiocyanate solution.
To beaker (c), 50 cm® of water and 20 g of ammonium chloride and stir.
The ammonium chloride will quickly dissolve. It will be found that
in (a) and (b) the red colour will intensify; i.e. increased concentration
of either of the reagents on the left-hand side of the equation forces the
equilibrium position to the right; in (c) the red colour will weaken, i.e.
addition of the (colourless) right-hand side reagent forces the equilibrium
position to the /eft.

Le Chatelier’s Principle
The principle of Le Chatelier (sometimes ascribed to Le Chatelier and
Braun jointly) is universal in its application in the field ofjphysical events;
here it will be treated mainly in its chemical applications.
Le Chatelier’s Principle can be stated in the following way:
If a system is in equilibrium and one of the factors affecting the
equilibrium is changed, the equilibrium will shift so as to annul, or tend
to annul, the effect of the change.
The principle is so general in its application that no better can be done
than illustrate its working by characteristic examples, as below.

1. Effects of temperature change


Suppose the reversible reaction
A+ B=C+D with heat absorbed

operates (as shown) endothermally from left to right. Suppose that a


system involving A, B, C, and D is in equilibrium at a certain temperature
and the temperature is then raised. Le Chatelier’s Principle requires that
the equilibrium shall move in such a way as to tend to reduce the tempera-
ture again; that is, to absorb heat. This requires a shift of equilibrium to
the right, producing higher concentrations of C and D. Thus, in all cases,
an endothermic reaction is favoured by rise of temperature.
Chemical Equilibria I: Molecular Equilibria 205
Incidentally, the rise of temperature also increases the rate of reaction.
Consequently it is usual to operate an endothermic reaction at the highest
possible temperature, e.g. that of the electric arc, and not to employ a
catalyst. Examples of endothermic reactions promoted by high temperature
are:
N,(s) + 0,(s) = 2NO(l) AH + 180
C(s) + 2S(s) =CS,(l) AH? + ulreading from
30.(g) = 20,(g) AH® + 285 kJ left to right

Reading the reaction in the reverse direction,


D+C=A-+B
with heat evolved, it is exothermic when operating from left to right. If the
temperature of an equilibrium system containing A, B, C and D is Jowered,
Le Chatelier’s Principle requires the equilibrium to shift so as to tend to
raise the temperature again; that is, to evolve heat. This requires a shift of
equilibrium to the right as the equation is written above. Thus, an exo-
thermic reaction is favoured by Jowering of temperature.
Incidentally, the fall of temperature reduces the reaction rate, which is a
disadvantage to industrial production. For this reason, a catalyst is usually
employed in industrial exothermic reactions. Then a temperature can be
used which is low enough, with the help of the catalyst, to give both a
reasonably favourable equilibrium and a reasonably rapid rate of attain-
ment of the equilibrium. Important examples of this are:
Catalyst Temperature
N.(g) + 3H.(g) = 2NH;(g) AH = — 92 kJ Tron 723 K
2S80.(g) + O.(g) =2S0,(g) AH = —778kJ V.O,orPt 723K
Since the great majority of industrial reactions are exothermic, catalysts
are very frequently used.

2. Effects of pressure change


Since liquids and solids are only very slightly compressible, the effects of
pressure change are significant only in the case of gaseous reagents.
Consider a reaction
nA + mB=C
in which A, B, and C are all gaseous and C is produced with a diminution
in the number of molecules, that is, a decrease of volume. Let the system
containing A, B, and C be in equilibrium (at constant temperature), and
then let the pressure on the system be increased. Le Chatelier’s Principle
requires the equilibrium to shift so as to tend to decrease the pressure
again. This requires the gaseous system to take up a smaller volume and so
exert lower pressure; that is, the formation of C is favoured, and the
206 Physical Chemistry
equilibrium moves from left to right as the reaction is written above. In
general, a reaction which proceeds with decrease of volume is favoured by
high pressure. The converse of this is also true. If, however, a reaction
proceeds with no volume change, the associated equilibria are independent
of pressure.
An important industrial case, in which very high pressure (at least
200 atmospheres) is employed, is the synthesis of ammonia (Haber’s
process).
N.(g) + 3H2(g) = 2NH2(g)
l.vol— 3'vol: 2vol. (at constant temp. and pressure)
Sulphur dioxide is converted to trioxide with diminution of volume:
2S0,(g) + O2(g) = 2S03(g)
2 vol. lvol. 2vol. (at constant temp. and pressure)
This oxidation is favoured by high pressure, but, in practice, the yield at
atmospheric pressure is such that higher pressure is not necessary.
Absorptions of gas to form a solid product are favoured by high pressure,
C8.
NaOH(s) + CO(g) —> HCO,Na(s)
Vols. negligible 22.4dm° negligible (at s.t.p.)
Polymerizations of gases are favoured by high pressure because the
polymerization reduces the number of molecules and, further, the polymer
may occupy relatively negligible volume as liquid or solid, e.g.
3C,H.2(g) ae C.H,(1)
ethyne benzene

nCzH.(g) —- (CsH,),(s)
ethene polyethene

Reactions in which there is no volume change, and in which the equilibria


are independent of pressure, include
N.(g) + O.(g) = 2NO(g)
H.(g) + I.(g) = 2HI(g)

3. Changes of concentration
Consider a reversible homogeneous reaction:
A+B=C+D
Suppose the system is in equilibrium with A, B, C, and D all present. Then
let the concentration of A be increased by addition of A from outside the
system. Le Chatelier’s Principle requires the equilibrium to shift so that
the concentration of A is reduced; i.e. the reaction will move from left to
right, reducing the concentration of B and increasing that of C and D.
Corresponding effects are produced by an increase in the concentration of
Chemical Equilibria I: Molecular Equilibria 207
B; increase of concentration of C or D favours the reverse reaction by
similar reasoning. Examples of these principles in operation were provided
by the bismuth oxychloride and iron(II) thiocyanate reactions earlier in
this chapter.

Le Chatelier’s Principle and solubility


Most solids which dissolve in water do so with absorption of heat. If an
aqueous solution is in equilibrium with the corresponding solid, and the
temperature is then raised, Le Chatelier’s Principle requires that the equilib-
rium must shift so as to tend to Jower the temperature, i.e. to absorb heat.
To do this, more solid must dissolve. This explains why most solids show
increased solubility with rise of temperature.
Gases, however, usually dissolve in water with evolution of heat and, by
the reverse of the above reasoning, should become /ess soluble with rise of
temperature. This, in fact, they do. If a gas is in equilibrium with its
aqueous solution (at constant temperature) and the pressure is then raised,
Le Chatelier’s Principle requires the equilibrium to shift so as to tend to
reduce the pressure. To do this, the gas must tend to occupy a smaller
volume, i.e. to dissolve further in the water. This, in fact, it does, the mass
of gas dissolved in a given mass of water at constant temperature being
directly proportional to the pressure (Henry’s Law, see page 165).

Le Chatelier’s Principle and the Equilibrium Law


Le Chatelier’s Principle can be used to make qualitative predictions about
the change in position of an equilibrium when any of the conditions
temperature, pressure, or concentration are varied. The Equilibrium Law
makes both qualitative and quantitative predictions about the effect of
changing either the pressures or the concentrations but it cannot be used
alone to predict the effect of changes in temperature. The sign of the
enthalpy change, together with Le Chatelier’s Principle, will give some
qualitative indication of the effect of temperature. The quantitative
relationship between the equilibrium constant for a reaction and tempera-
ture can be shown experimentally to be that the logarithm of the equi-
librium constant varies inversely with temperature. That is a graph of In K
against 1/Tis a straight line. The slope of the straight line can be shown
to be —AH®/R and the relationship between K and T can be written as:
ce

In kK = a + constant.

This relationship can be used to predict values of equilibrium constants


for different temperatures and hence make quantitative predictions about
the position of equilibrium. It is important to realise that such predictions
refer to only the position of equilibrium when it is attained and tell us
nothing about the rate of attainment of equilibrium.
15. Chemical Equilibria I:
Tonic Equilibria

The Dilution Law of Ostwald is an attempt to apply the Equilibrium Law to


ionic solutions and so to obtain a mathematical expression that will state
accurately the behaviour of a given electrolyte in solutions of varying
concentration at constant temperature.
Consider ethanoic acid as a typical weak, binary electrolyte (i.e. an
electrolyte ionizing only slightly and into two ions per molecule). The
ionic situation is
CH;CO,H(aq) = CHgCO,.7 (aq) + H*(aq)
Originally, before
ionization: 1 mole — —
At equilibrium in
V dm? of solution: (1 — «) moles med » a moles
where « is the degree of ionization of the acid.
By the Equilibrium Law at constant temperature.

[CH,CO,7 ][H*]
a
[CH3CO2H]
Substituting
K a arly?

‘(l= ®/V
a2

~ (—aV
where K, is the dissociation constant of the electrolyte.
(By convention, the concentrations of the materials on the right-hand
side of the ionic equation are put into the numerator of the final fraction.
V is the volume of solution in dm* which contains 1 mole of the original
electrolyte.)
This final fraction is the mathematical expression of the Dilution Law
and, so far, rests on pure theory. To test it, the degree of ionization of
ethanoic acid must be determined over a wide range of dilutions; the
Chemical Equilibria IT: Ionic Equilibria 209
results must be applied to the left-hand side of the Dilution Law fraction
above to see if they do, in fact, give a constant value as the Dilution Law
requires. Experimental results obtained for ethanoic acid are shown in the
table below.
The degree of ionization is determined by electrical conductance. It will
be observed that the final column shows reasonably constant figures after
V
has reached a value of about 2 dm® and the solution is 0.5 M, or less, in
concentration. This general position is found to apply for all weak elec-
trolytes, i.e. the Dilution Law is a satisfactory expression of the ionization
behaviour of weak electrolytes in dilute solution. K is used to represent the
dissociation constant of weak electrolytes in general, whereas the use of
K, is confined to the dissociation of weak acids.

At 298 K throughout
Vol. V (dm®), containing Degree of ionization a?
1 mole of ethanoic acid e (1
—«)V

1.977 0.00570 7.652%. 10°


5.374 0.00981 vBieXe10~*
10.753 0.0138 1.80 x 10-®
24.875 0.0216 1.92 xv10-*
63.26 0.0336 1.85°¢:1078

For very weak electrolytes, i.e. those with very little ionization, an
approximate form of the Dilution Law can be used. In this case, « is so
small that (1 — «) does not differ appreciably from unity and can be taken
as unity with no significant error. So, for very weak electrolytes, the
expression
ao?
ee eee K
(l — a)V
can be written
a2
at ET
V

and this is the Dilution Law in approximate form.


From this form, it follows that «2 = KV, or « = »/Ky/V. Since the
square root of K is a constant, it follows that, for very weak electrolytes
the degree of ionization is directly proportional to the square root of the
volume of solution containing 1 mole of the electrolyte. If preferred, the
equation for the dissociation of weak electrolytes may be determined using
210 Physical Chemistry
concentrations rather than dilutions. For example if the initial concentra-
tion of ethanoic acid is c (mol dm~*) and the degree of dissociation is «,
CH,CO,H(aq) — CH;CO,." (aq) + H*(aq)
Before
ionization: c 0 0 mol dm=-?
At equilibrium: c(1 — «) Ca CH mol dm=*
Substituting in the Equilibrium Law,

irae

“7-9
ca?

The following example shows a characteristic use of the Dilution Law.


Example Ethanoic acid has a degree of ionization of 1.4 per cent in
decimolar solution at 298 K. Calculate its degree of ionization in one-
hundredth molar solution at the same temperature.
CH;CO.H(aq) = CHs,CO; (aq) + H*(aq)
At equilibrium: (1 — a) m a moles
If the volume containing 1 mole of acid is V dm, :
a2
ih a
(1 — a)V
Applying the data for decimolar solution,
2
Co) = 0.000 019 9
a (= 01014) 10
= 1.99 x 10-5 mol dm-?
If «, is the degree of ionization of ethanoic acid in one-hundredth molar
solution,

ae Cle = 0.000 019 9


(1 — «,)-100
Rearranging this expression, we have
a? + 0.001 99a, — 0.001 99 = 0
%, = 0.0487
That is, the degree of ionization of ethanoic acid in 0.01 M solution at
298 K is 0.0487 or 4.87 per cent.
Chemical Equilibria II: Ionic Equilibria 211
Using the approximate form of the Dilution Law,

Bae 20
% 4/100
a= «4/10

= 0.014 x 3.16
= 0.044,
We have seen that the Dilution Law expresses quite well the behaviour
of weak electrolytes in dilute solution. It breaks down completely, however,
when applied to strong electrolytes. The case of potassium chloride
illustrates this fact.

All measurements at 291 K

Vol. (dm*) containing Degree of ionization a?


1 mole of KCI apparent (1—a)V

1 0.757 PIES)
5 0.831 0.815
50 0.923 0.222
200 0.958 0.108
1000 0.980 0.0485

It will then be seen that the required ‘constant’ of the third column varies
progressively from 2.35 to less than 0.05, falling to about one-fiftieth of the
initial value. Several attempts have been made, notably by Debye and
Hiickel, to produce alternative equations for the behaviour of strong
electrolytes, but they lie outside the scope of this book.

Acidity, Alkalinity, and Neutralization


The ionic state of water
Water has a solvent action for a very wide variety of materials and is, in
consequence, difficult to purify. Water is purified by running it through
columns containing cation and anion exchange resins. During this process
the cations in the water are replaced by hydrogen ions and the anions by
hydroxide ions. The H* ions and the OH™~ ions then combine to form
water.
The progress of the purification can be followed by measurement of the
electrical conductance of the water. Starting at a comparatively high level
(caused by impurity), it will be found to fall gradually to a very low value,
which cannot be reduced by further efforts at purification. This constant
212 Physical Chemistry
electrical conductance is believed to be genuinely that of water itself and
to be the result of an ionization which is usually represented as
H,0() = H*(aq) + OH (aq)
though the hydrogen ion is actually hydrated (as the hydroxonium ion
H,0+; see also page 75) and the ionization is more properly written as
2H,0(1) = H3;0* (aq) + OH~ (aq)
Except for special reasons, the simpler equation will generally be used.
Applying the Equilibrium Law,

iK= [H*] [OH7]


[HO]
Due to the ionization being very slight, the equilibrium concentration of
the water does not differ appreciably from the original concentration and
so it is regarded as being a constant and is incorporated in the equilibrium
constant. The expression now becomes
K, = [H*] [OH]
The very small conductance of pure water (conductivity 3.84 x 10-8
ohm-! cm-! at 294 K) can be shown to correspond to the following
concentrations of hydrogen ion H*, and hydroxide ion, OH~, in the
water: :

[H*] = [OH-] =1 x 10-7 at 298K


(The square brackets indicate concentrations in mol dm~%, the usual con-
vention.) From these figures, the product of the hydrogen and hydroxide
ion concentrations is given by
at 298K [H*][OH-} = (1 x 10-7) = (1x 10~)
=] < 105**imol? dm>-
= é

This quantity, K, of the value | x 10714, is called the ionic product of


water and is a very important constant. It is always maintained in an
aqueous liquid at 298 K. In pure water, it is maintained by equal con-
centrations of H+ and OH7 ions, both being 1 x 10-7 mol dm~ and this
situation defines a neutral solution. That is, a neutral aqueous solution is
one in which the concentrations of hydrogen ion, H+, and hydroxide ion,
OH, are equal at the value of 1 x 10-7 mol dm~§ at 298 K.
At any other temperature K,, will have a different value and the H+ and
OH~ concentrations of a neutral solution will be different. In fact, as the
reaction between hydrogen ions and hydroxide ions is exothermic, a rise in
temperature will cause an increase in the concentration of the hydrogen
ions in a neutral solution.
Chemical Equilibria II: Ionic Equilibria 213
However, at 298 K the value of K,, has to be maintained at 1 x 10-14, so
that any variation in the concentration of one of the ions must be com-
pensated for by a change in the concentration of the other ion. For example,
suppose pure water is acidified by hydrogen chloride till the concen-
tration of HCl is 10~? mol dm~°. At this concentration it can be considered
as fully ionized, so that [H*] = 10-?. This H+ participates in the equilib-
rium of water:
{H*](OH-] = K, = 1 x 1072* mol? dm-6
Consequently [OH™] falls to 10-?# mol dm~*. The product K,, is still
maintained, but by unequal concentrations of H+ and OH™ ion. The
preponderance of H* is characteristic of an acidic solution. Similarly, if
pure water is made one-thousandth molar in potassium hydroxide
[OH-] = 10-3, K, (1 x 10714) is maintained by a fall of H+ concentra-
tion to 107+ mol dm~%. The preponderance of OH™ is characteristic
of an alkaline solution.
It is important to notice that, however acidic an aqueous solution may be,
it still retains enough OH to satisfy the requirement K,, = [H*][OH-] =
1 x 10~+*. Similarly, however alkaline an aqueous solution may be, it
still retains enough H* to satisfy this requirement.

The nature of neutralization


This is best illustrated by an example. Suppose 0.2 M hydrochloric acid
and 0.2 M sodium hydroxide solution are mixed, perfectly and instantane-
ously, in equal volumes so that both become 0.1 M. At the moment of
mixing, assuming complete ionization of the two strong electrolytes,
[H*] = [OH-] = 10-1. That is, for the moment, the product of
{[H*+][OH~] = 10-2. This is very much in excess of the permissible value of
K,,, the ionic product of water, which is 1 x 107-14. Consequently, OH~
and H* are at once withdrawn from the solution as unionized molecules
of water until the value of K,, is again attained. This process of readjusting
the disturbed ionic product of water is the process of neutralization.
H*(aq) + OH~(aq) = H,0())
The reaction is reversible but lies so far to the right at equilibrium that for
all practical measurements neutralization is quantitatively complete; this is
the assumption made in acid-alkali titrations in the laboratory.

The hydrogen ion index, pH


The hydrogen ion index is a device which has been adopted for con-
venience in stating the nature of a solution with respect to acidity. It is
defined in the following way:
If the concentration of hydrogen ion, H*, in a solution is 10-* mol dm= °,
then the pH of the solution is x.
214 Physical Chemistry
re:
pH = —lg [H*] (or —logio [H*])
This leads to the following conclusions.
Neutrality Pure water shows the situation (at 298 K),
[H+] = 10-7 (Also [OH-] = 10~")
From this, pH in pure water is 7. That is, a neutral aqueous solution at
298 K has a pH of 7.

Acidity Consider, say, 0.01 M HCl. The acid may be taken as fully
ionized, i.e. [H*] = 10-2. From this, the pH of the solution is 2. Similarly,
all other acidic solutions have a pH less than 7.

Alkalinity Since alkalinity is associated with the ion, OH~, it might be


thought that it would be indicated by a hydroxide ion index, pOH.
However, we have seen that an aqueous liquid always shows the relation
(at 298 K):
K, = [H*][OH-];= lx 107+*mol dmz
From this, it follows that pH + pOH = 14; consequently, pOH is
always related to pH in the same solution by the constant relation
pOH = 14 — pH. ‘
For this reason, it is quite convenient to state both acidity and alkalinity
in terms of the one index, pH. Consider 0.001 M potassium hydroxide
solution. If the alkali is fully ionized, [OH~] = 10~%. That is, pOH is 3 and
pH is (14 — 3) or 11 for the solution. From this, we see that an alkaline
solution has a pH greater than 7.

pH of aqueous liquid Nature ofliquid

less than 7 acidic


neutral
greater than 7 alkaline

It is very important to be quite clear about the following position. Since


the value of pH is a logarithmic index with the negative sign omitted, an
increase in hydrogen ion concentration (i.e. in acidity) is accompanied by
a decrease in the pH of the solution. For example, an increase in hydrogen
ion concentration from 10-® to 10-2 mol dm~® means a decrease in pH
from 5 to 2. Consequently, low values of pH indicate high acidity. Con-
versely, high values of pH mean low acidity and values of PH greater than
7 indicate alkaline solution, increasing in alkalinity as the pH rises.
Chemical Equilibria IT: Ionic Equilibria 215
The value pH = 0, for example, indicates a solution with a concentra-
tion of 10° mole of H+ dm~®. Since 10° = 1, such a solution is a molar
acid with the acid fully ionized. Correspondingly, pH = 14 indicates a
solution in which pOH = 0, i.e. a molar solution of a fully ionized strong
alkali.
The following examples illustrate the calculation of the pH of a solution.

Example 1 Calculate the pH of a decimolar solution of ethanoic acid, in


which the acid is 1.4 per cent ionized.
If the decimolar acid is fully ionized, the H* concentration is
10-1 mol dm~%. But as the degree of ionization is 0.014, the actual Ht
concentration is 10-1? x 0.014 = 0.0014 mol dm~§, ice.
[H+]
= 0.0014
= 10°48 (ig 1.4 = 0.146)
ne 10-3+0-146
= 10)-2-854
i.e. pH = 2.85 (to 3 sig. figs.)
(Notice that the logarithmic index has to be made completely negative by
manipulation of the negative characteristic and the positive mantissa of the
logarithm.)

Example2 Calculate the pH of a hundredth-molar solution of ammonium


hydroxide, in which the degree of ionization of the electrolyte is 0.043.
If the alkali is fully ionized, the OH ion concentration is 10-? moldm~§8,
But, as the degree of ionization is 0.043, the actual OH~ concentration is
10-2 x 0.043 = 0.00043 mol dm~°, ice.
[OH-] = 0.00043
= 104-694 (Ig 4.3 = 0.634)
107-440-834
10- 3-366

From this,
pOH = 3.37 (to 3 sig. figs.)
pH = 14 — 3.37 = 10.63

Acids and Bases


Development of the idea of acidity
Some vague notion of the more obvious properties of acids has probably
existed for a very long time because of their occurrence in sour and unripe
fruits and the production of ethanoic acid by the souring of wine (the
oxidation of ethanol). At any rate, by the middle of the seventeenth
century the properties of acids were recognized as including a sour taste,
corrosive action, ability to change the colours of certain vegetable products
216 Physical Chemistry
(i.e. indicators, such as litmus), ability to precipitate sulphur from potas-
sium sulphide, and ability to react with alkalis to produce neutral substances.
Alkalis were recognized as possessing detergent properties, the ability to
dissolve fats and oils (saponification) and also sulphur, to reverse the
action on indicators (above) and ‘neutralize’ acids. Later, the ability of
certain (strong) acids to dissolve metals with liberation of hydrogen was
recognized.
Towards the end of the eighteenth century, Lavoisier produced the
oxygen theory of acids. This was an unjustified generalization from the
fact that many acids can be produced by the combination of non-metallic
oxides with water, e.g. H2SO,, H.SO3, HPO3, HgPO., H3AsOg, and so
contain oxygen. Lavoisier extended this experience wrongly to produce the
theory that oxygen is the essential acid-producing element. The name
oxygen is, in fact, from oxus (sour) and gennao (to produce). Davy’s proof
that chlorine is an element, containing no oxygen, and that hydrochloric
acid is produced by the combination of hydrogen with chlorine finally
disproved the oxygen theory, though Berthollet had earlier recognized the
absence of oxygen in the acids HCN and H,S.
In the early nineteenth century, it was gradually recognized that the
element essential to an acid is hydrogen, but that the hydrogen must be of a
particular type, that is, hydrogen which is capable of being replaced by a
metal (with formation of a salt), e.g.

Zn + H,SO,— ZnSO, + He (direct replacement)

CuO + H,SO, — CuSO, + H.O (indirect replacement)

At this time, an acid could have been roughly defined as a compound


which turns blue litmus to red and contains hydrogen which can be
replaced, directly or indirectly, by a metal.
With the introduction of ionic ideas (about 1880) replaceable hydrogen
was recognized as hydrogen which is capable of ionizing, so that an acid
could be defined as a compound which, in water, yields hydrogen ions, e.g.
HCl= H++c-
H,SO, = 2H* + SO2-
It is now recognized that the simple hydrogen ion (or proton) does not
exist in aqueous solution but is solvated to give the hydroxonium ion,
H,0*. Subject to the discussion which follows, we may say that the
ability to produce this ion in water is recognized as the essential property
of an acid, e.g.

HCl + H,O =H,0* +Cl-


H,SO, + 2H,0 = 2H;0+* + SO2-
CH;CO.H + H,0 =H,0* + CH;CO;
Chemical Equilibria II: Ionic Equilibria 217
Essential nature of acid and base
From the ionic point of view an acid can be regarded as a substance which
has a tendency to lose one or more protons (i.e. hydrogen ions) per mole-
cule. Conversely, a base can be regarded as a substance which has a
tendency to gain one or more protons per molecule or ion. This definition
of acids and bases was first put forward by Bronsted and Lowry in 1922.
The relation can be expressed in the equations
acid = H* + base
H* + base = acid
That is, for every acid, there exists a base, which is produced when the
acid loses a proton (or H*). The acid and base which stand in this relation
to one another are said to be conjugates. (For examples, see below.) An
acid which loses a proton readily from its molecule is said to be a strong
acid; an acid which loses a proton with difficulty from its molecule is said
to be a weak acid. It is obvious that a strong acid must have a weak
conjugate base, and vice versa; for example:
Strong acid Weak base
HCl= H*+cC>-
H,SO, = 2H* + SO2-
HNO; = H* + NOZ

Weak acid Strong base


CH;CO,H = H* + CH;COz
H.O = Ht + 0OH-
H,C,0, = 2H* a G0

In practice, however, the free proton never exists in solution but is always
solvated. When water is the solvent, the H* ion (or proton) is hydrated to
H,O+, which is known as the hydroxonium ion. The water molecule which
hydrates the proton in this way is acting as a proton acceptor, i.e. as a
base, so the essential relation can be expressed in equations such as
HCl + H,O =H,0+* + Cl- (1)
CH,CO.H + H,O0 =H,;0* + CH3;CO;z (2)
acid, + base, = acid, + base,
In equation (1), HCl is a very strong acid, losing a proton with ease; the
conjugate base, Cl~, is very weak. In dilute solution, the equilibrium lies
so far to the right, that for practical purposes ionization is complete.
In equation (2), ethanoic acid is weak, losing a proton with difficulty ;
the conjugate base, CH;COz, is strong. In decimolar solution, the
~ equilibrium lies so much to the left that only fourteen molecules of ethanoic
acid per thousand are ionized. The ionization increases with dilution
(according to the Dilution Law) so that with dilution the acid becomes
stronger.
218 Physical Chemistry
This raises the point that the terms strong and concentrated are now used,
in connection with acids and bases, with quite distinct meanings.
A strong acid is one which very readily loses a proton from
its molecule and so (in dilute solution at least) tends to be highly
ionized.
A concentrated acid is one in which the proportion of acid
to water is very high.
The opposite of a strong acid is a weak acid, and this term is
applied to an acid which loses a proton with difficulty and which,
at any significant concentration, is only slightly ionized.
The opposite of a concentrated acid is a dilute acid, and the
term Is applied to a liquid in which the proportion of acid to
water is low.
An example of a concentrated, weak acid is glacial ethanoic acid. The
percentage of water in it is almost nil and the acid loses a proton with
difficulty. An example of a dilute, strong acid is bench hydrochloric acid.
The proportion of hydrochloric acid is low (about 7 per cent) and the acid
loses a proton very readily. These terms are used correspondingly with
bases.

Water in relation to acidic and basic behaviour


Water can show basic behaviour by accepting a proton and does this
when acids ionize in aqueous solution, e.g. ;
HCl + H,O = H;0* + Cl-
The molecule H,O accepts a proton to become the hydroxonium ion,
H;0+.
Water can also show acidic behaviour by losing a proton (H*) to
produce the base, OH~. This occurs, for example, when ammonia gas
reacts with water producing an alkaline solution.
NH; + H,O = NH; + OH-
Water supplies a proton to the NH; molecule to produce the ion, NH}.
Having proton-accepting (protophilic) and proton-donating (protogenic)
properties, water is called an amphiprotic solvent.
It should be noted that the hydroxyl bases must now be considered as
only one group of bases among many, all of which show, in varying
degrees, the property of combining with protons. Some of these bases
were shown a little earlier and included materials such as Cl-, SO2-,
NO;, C,07° and CH;CO; as well as OH-. From the modern point of
view, the principal distinguishing feature about the hydroxyl base OH° is
its very great strength as a base. The equilibrium position of its proton-
accepting reaction
OH” +H* =H,0 (or OH- + H;0* = 2H,0)
Chemical Equilibria II: Ionic Equilibria 219
lies so far to the right (as written above) that for practical purposes of
measurement in titration it is regarded as irreversible. This reaction is the
very familiar phenomenon of acid-alkali neutralization when read from
left to right, but is merely one special case of its kind, i.e.
base + proton = acid
The following definitions can now be formally stated:
An acid is a substance which shows a tendency to lose a proton;
if the tendency is marked, the acid is said to be strong; if the
tendency is slight, the acid is said to be weak.
A base is a substance which shows a tendency to gain a proton;
if the tendency is marked, the base is said to be strong; if the
tendency is slight, the base is said to be weak.
An alkali is a substance which is soluble in water and produces
the hydroxide base, OH-, in the solution.
Neutralization is the special case of the reaction
base + proton = acid
in which the hydroxyl base, OH-, is involved, i.e.
OH- (aq) + H+ (aq) = H,O(l)
It must be noted that in an alkaline solution the hydroxyl base, OH-, is
always accompanied by an equivalent concentration of a cation, such as
Nat, K*, Ca?*, or NHj. When the base undergoes neutralization,
combining with protons supplied by the participating acid, the conjugate
base of the acid (an acidic ion such as Cl~, SO2-, or NO;) is left in
solution, together with the cation associated with the base. This aggrega-
tion of ions in the liquid constitutes the salt, which is normally regarded as
a product of neutralization. For example,
Na+ + OH- +H*++Cl—Na?*+Cl- + H,O
The effect of this reaction is to substitute a cation for some (or all) of the
hydrogen of an acid which is capable of being lost as protons. In this sense
a salt is a by-product of neutralization, the essential of which is proton-gain
by the base, OH~. A salt can be defined in the following way:
A salt is a compound produced when a metallic cation is
substituted for some (or all) of the hydrogen of an acid which Is
capable of being lost as protons. When the substitution is
complete, the salt is said to be normal; when incomplete, the
salt is said to be an acid salt; for example:

Acid Acid salts Normal salts

H.SO, Na*+HSO; (Nat),SO2-


H,CO, 7 HCO: (K*),CO2-
H.SO, |NatHSO; (Na*),SO3-
220 Physical Chemistry
One important aspect of the Bronsted-Lowry view of acids and bases is
that it can be applied to non-aqueous solvents. For instance, the dissolving
of hydrogen chloride in liquid ammonia may be compared to its dissolving
in water.
HCl + NH, = NH} + Cl-
HCl + H,O = H,0* + Cl-
In fact a solution of ammonium chloride in liquid ammonia does have
acidic properties. In addition, a solution of sodium amide in liquid
ammonia exhibits basic properties in a similar manner to a solution of
sodium hydroxide in water.
NaNH, — Nat + NH;
NaOH — Nat + OH
So sodium amide neutralizes ammonium chloride in liquid ammonia
solution.

Comparison of the strengths of acids


It has already been mentioned that a strong acid is one which readily loses
a proton and is highly ionized in dilute aqueous solution. A weak acid is
one which loses a proton with difficulty and is never more than slightly
ionized in any solution of a usual working concentration, say from 2M to
0.1 M. The great majority of acids are weak, e.g. ethanoic, ethanedioic,
tartaric, citric, and other organic acids. The common strong acids are the
three mineral acids, hydrochloric, sulphuric, and nitric. Hydrobromic
acid, HBr, and perchloric acid are also strong. The division between
strong and weak acids is not rigid. Certain acids exist, e.g. sulphurous acid,
H.SOs3, which are moderately ionized at the usual working dilutions and
cannot be classified as definitely strong or weak.
It must be remembered that no absolute comparison of strengths of
acids can be made. The relative strengths of two given acids in aqueous
solution may vary with dilution and with temperature. For example, at
decimolar concentration and room temperature, hydrochloric acid is
almost completely ionized, while ethanoic acid is about 1.4 per cent
ionized. At 0.001 M, hydrochloric acid scarcely alters its strength at all
because ionization was already almost complete, but ethanoic acid alters
its ionization to about 14 per cent, so becoming considerably stronger.
This means that comparisons of strength can be made only at stated
dilutions. Since ionization also changes with temperature, this condition
should also be specified.
As ionization of a weak acid increases with dilution, it is obvious that
the acid becomes stronger as it becomes more dilute; in theory, at infinite
dilution all acids become equally strong because all become fully ionized.
Chemical Equilibria IT: Ionic Equilibria 221
Methods for comparing strengths of acids
oa following are some of the methods available for comparing strengths
of acids:

1. On a logarithmic scale
The dissociation constant of a weak electrolyte is sometimes expressed on
a logarithmic scale, which for an acid is denoted by pK, = —lg K,, and
for a base, pK, = —lg K,. The advantage of this scale is that the rather
cumbersome numbers are converted to relatively simple ones. For example,
for ethanoic acid,
K, = 1.99 x 1075 mol dm-
lg Kz = 5.2989
= —5 + 0.2989
= —4.7011
pk, = —lg K, = 4.7
The pK, values of some common’ weak acids are given below:
chloroethanoic acid pkKp = 29
methanoic (formic) acid pK, = 3.8
ethanoic acid pK, = 4.7
propanoic acid pK; = 49
carbonic acid
Ist dissociation pK, = 6.4
2nd dissociation pk, = 10:3

A large pK, value (which means a small K, value) indicates a weak acid.
Similar terms for bases, K, and pK,, may be used to give an indication of
their comparative strengths.
The explanation of why a particular acid is stronger than another is
related to the composition and structure of the acids concerned. It is
profitable to consider two acids which are only slightly different in struc-
ture and attempt to relate this difference to a comparison of their pK,
values. Ethanoic and chloroethanoic acid only differ by the substitution of
a chlorine atom for a hydrogen atom.
He 840 re
hye és glw4
H—C—C Cl—C—C
i Non H O-H
The difference in structure results in a significant difference in pK,, which
can be explained in terms of the more electronegative chlorine atom
transmitting its attraction for electrons throughout the molecule and so
222 Physical Chemistry
increasing the bond polarity (see page 97) of the O—H bond and hence the
ease of ionization of the hydrogen.

2. By observations ofstandard enthalpy of neutralization


Let two acids, of which the strengths are to be compared, be A and B.
The enthalpy of neutralization of 1 mole of hydrogen ions from the acid
A at a given dilution by 1 mole of hydroxide ions from sodium hydroxide
in suitable dilution is measured (or obtained from records). Let this be
x (J). Similar data is obtained from acid B. Let this be y (J).
Then the heat given out by mixing A and B in the same solution, so
that they each provide 1 mole of hydrogen ions, with sufficient sodium
hydroxide solution to provide 1 mole of hydroxide ions is measured. Let
this be z(J). In this experiment the acids may be considered as competing
for the base. If acid A neutralizes n mole of the hydroxide ions and acid B
neutralizes (1 — n) mole, the following relation must hold:

nx +(l—n)y =z

From this n can be calculated. Then:

Strength of A “Sapa
Strengthof B l—wz

This method is unsuitable for acids of roughly similar strengths because


for such acids, the heat changes are usually similar.

Constant values of enthalpy of neutralization for strong acids and bases


When accurate thermal observations began to be made, it was noticed that
the enthalpy of neutralization of 1 mole of H*ions from strong acid by
1 mole of OH~ions from any strong base, both being in dilute solution,
was almost constant at a value close to — 57.3 kJ mol-1. Writing typical
equations in the molecular manner:

NaOH(aq) + HCl(aq) — NaCl(aq) + H,O() AH = —57.3kJ mol-}


KOH(aq) +-HNO,(aq)—> KNO,(aq) + H,O(l) AH = —57.3 kJ mol-!
NaOH(aq) + HBr(aq) — NaBr(aq) + H,O(l) AH = —57.7kJ mol-!

This result was very surprising because these reactions, in molecular


presentation, appear quite different. The ionic theory, however, requires
this result for the following reasons. The acid and base, being ‘strong’,
are both fully ionized (or very nearly so) in dilute solution. The salt
produced is also a strong electrolyte (whatever its individual identity) and
so is fully ionized. Consequently the only change occurring in all these
Chemical Equilibria II: Ionic Equilibria 223
neutralizations is the formation of 1 mole of water from its ions. This is a
constant change, requiring a constant evolution of heat; as:
Na*(aq) + OH~(aq) + H*(aq) + Cl-(aq)
— Nat(aq) + Cl-(aq) + H,O() AH = —57.3 kJ mol-!
K* (aq) + OH-(aq) + H*(aq) + NO; (aq)
— K*(aq) + NO; (aq) + H.Od) AH = —57.3 kJ mol-!
In both cases the real changeisH+ + OH- +H2O0 AH = —57.3kJ mol-!.
If, however, the acid concerned is weak, i.e. only slightly ionized, the
enthalpy of neutralization differs from AH = —57.3 kJ mol-!. The
reason is that the process of neutralization takes place in two stages:

1. The completion of the ionization of the weak acid, which is accom-


panied by a heat change, e.g. for ethanoic acid:
CH;CO.H(aq) = CH3CO; (aq) + H*(aq) AH = XkJ mol!

2. The neutralization proper, as with sodium hydroxide:

Na*(aq) + OH~(aq) + H*(aq) + CH3CO; (aq)


—> CH; CO; (aq) + Na*(aq) + H200) AH = —57.3kJ mol“!
The net heat change is, therefore, (—57.3+X) kJ mol-!. For sodium
hydroxide+ethanoic acid, the experimental figure is —56.1 kJ mol-}.
That is
—57.3+ X¥ = —56.1
so that X = +1.2kJ mol}
Similar considerations apply for a weak base.

Salt Hydrolysis
Salt hydrolysis is essentially the reversal of neutralization. Consider the
reaction between an acid, HA, and a base, BOH, neither being strong.
The equilibrium between them can be written

H+ + A- + B+ + OH- =B* + A- +H,0


—~_———" MNn“/_—_"’
1 IL Ir
HA BOH Hitter OH
base salt water

Read from left to right, this process is neutralization; read from right to
left, it is salt hydrolysis. The following discussion will show that salt
hydrolysis is appreciable except when the salt is derived from a strong acid
and a strong base. There are four possible cases as below. One example of
each will be given in detail and others in outline only.
224 Physical Chemistry

1. Salt of a weak acid and a strong base


(a) Sodium ethanoate This salt is derived from the weak acid, ethanoic,
and the strong base, sodium hydroxide. In sodium ethanoate solution, the
equilibria set up are
CH,;CO; (aq) Na*(aq)
aa

H,O()= H*(aq) + OH (aq)


tt
CH,CO.H(aq)
The salt (a strong electrolyte) can be considered fully ionized. Since sodium
hydroxide is a strong base, it can be considered as remaining completely
ionized, with no formation of ‘molecules’ of NaOH. Ethanoic acid,
however, is weak. As shown above, it must form some unionized molecules,
using H* derived from water. This disturbs the equilibrium of water
((H*] {(OH-] = K,.=1 x 10°** at 298 K). To restore -K,,, more water
ionizes. This puts OH™ into excess, causes the pH of the solution to rise
above 7 and the solution to react alkaline.
This is a general situation.
A solution of a salt derived from a weak acid and a strong base always
contains unionized molecules of the acid and reacts alkaline.
Further examples of this are given below and can be fully argued on the
same lines as for sodium ethanoate.
(b) Sodium carbonate
2Na*(aq) COz" (aq)
a

2H,O(l) = 20H-(aq) + 2H*(aq)


It
H,CO,(aq)
(c) Potassium cyanide
K* (aq) CN™ (aq)
+

H2,0() = OH (aq) + sah


)
HCN(aq)
It (A)
HCN(g)
The general situation here is similar to that of sodium ethanoate solution:
there is, however, the additional factor that hydrogen cyanide, HCN, is
extremely volatile. Consequently, if the solution is boiled, the vapour of
hydrogen cyanide is expelled into the air. The equilibrium (A) tends to be
Chemical Equilibria IT: Ionic Equilibria 225
displaced continually downwards. This causes further hydrolysis to replace
the hydrogen cyanide in the solution, which is again expelled. This con-
tinues until eventually only K+ and OH" ions remain (with the trace of Ht
to maintain K,,); that is, the final solution is one of potassium hydroxide.
Corresponding reactions occur with sodium cyanide.

2. Salt of a strong acid and a weak base


(a) Ammonium chloride The salt is derived from the strong acid, hydro-
chloric, and the weak base, ammonium hydroxide. In ammonium chloride
solution, the situation is
NH}(aq) —-CI-(aq)
4.

H,O() = OH-(aq) + H*(aq)


t
NH,OH(aq)
Like all salts, ammonium chloride is a strong electrolyte and can be taken
as fully ionized. Hydrochloric acid is a strong acid, so there is no formation
of HCI molecules. Ammonium hydroxide, however, is a weak base so that
molecules of NH,OH are formed, as shown, by use of the OH™ ion from
water. Withdrawal of OH™ ion disturbs the ionic equilibrium of water.
(ii*|[OH-], = K,, = 1x 10-** at 298 K) and, to. restore XK,, water
ionizes further. This puts H* into excess, the pH of the solution falls below
7 and it reacts acidic.
This is a general situation.
A solution of a salt from a strong acid and a weak base always contains
unionized molecules ofthe base and reacts acidic.
Further examples are stated in outline below and can be fully argued as in
the case of ammonium chloride.
(b) Ammonium sulphate
2NH;} (aq) SO2-(aq)
+

2H,0(0) = 20H-(aq) + 2H*(aq)


i)
2NH,OH(aq)
Salts of organic bases, such as aniline, behave like ammonium salts.
Taking aniline hydrochloride as the example, the situation is
C.H,NHi(aq) Cl-(aq)
+

H,O()) = GaKe + H*(aq)


C,H;NH;,OH(aq)
226 Physical Chemistry
(c) Iron(II1) chloride (and aluminium chloride)
Fe? + (aq) 3Cl- (aq)
+

3H,O(l)= 30H-(aq) + 3H*(aq)


It
Fe(OH).,(aq)
The case of aluminium chloride is similar, substituting Al for Fe.
In these cases, hydrolysis may be so marked that the solution will dissolve
magnesium with liberation of hydrogen.
Mg(s) + 2H*(aq) > Mg**(aq) + H2(g)
(d) Copper(I1) sulphate
Cu?*(aq) SO,?7 (aq)
a

2H,0(1) = 2O0H-(aq) + 2H*(aq)


It
Cu(OH)2(aq)
The above explanation of the hydrolysis of iron(II) chloride, aluminium
chloride, and copper(II) sulphate is consistent with that used forammonium
salts but it is somewhat simplified as it ignores the hydration of the ions.
The cations in these solutions are heavily hydrated and exist as
[Fe(H,O).]**, [AI(H,0).]°*, and [Cu(H,O),]?*. It is the reaction of
these hydrated ions with water which causes their solutions to be acidic. In
the case of iron(III) chloride the following sequence of reactions occurs,
culminating in the precipitation of colloidal hydrated iron(III) hydroxide.
Fe(H,0)3+ + H,O = Fe(H,0),OH?* + H,O*
Fe(H,0);0H2* + HO = Fe(H,0),(OH); + H,0+
Fe(H,0),(OH)+ + H,O = Fe(H,0),(OH), + H,O*
Similar equilibria are considered to be set up in aqueous solutions of
aluminium chloride and copper(II) sulphate.

3. Salt of a weak acid and a weak base


(a) Ammonium ethanoate This salt is derived from the weak base,
ammonium hydroxide, and the weak acid, ethanoic acid. The salt is a
strong electrolyte and can be taken as fully ionized.
NH}(aq) CH;CO; (aq)
+ +
H,O(1)) = OE te + H*(aq)
)
NH,OH (aq) CH,CO,H(aq)
Chemical Equilibria II: Ionic Equilibria 227
Since the acid and base are both weak, unionized molecules of both must
be formed as shown, by utilizing the ions H+ and OH™ derived from
water. The removal of these ions disturbs the ionic product of water
((H*][OH-] = K, = 1 x 10-+* at 298K. To restore K,,, water ionizes
further. In this particular case, the acid and base are about equally weak
(Kis about 2 x 10~° for both) so that the concentrations of H+ and OH-
remain about equal in the solution in spite of hydrolysis. So ammonium
ethanoate is strongly hydrolysed in solution, but the solution remains
almost neutral.
In general, if the acid from which the salt is derived is stronger than the
base, the solution tends to react acidic, and vice versa.
(b) Aluminium sulphide This salt, Al,S3, can be made by heating the
two dry elements together. The corresponding acid, H.S, and base,
Al(OH)3, are, however, so weak that the salt is completely hydrolysed in
water, with liberation of hydrogen sulphide and precipitation of aluminium
hydroxide.
2Al?+(aq) —-3S?-(aq)
+ +
6H,O = 6OH-(aq) + 6H*(aq)
I I
2A1(OH),(s) 3H.S(g)
Because of this hydrolysis, the mixing of solutions of a soluble sulphide and
an aluminium salt precipitates aluminium hydroxide, not aluminium
sulphide as might be expected.
Because of similar hydrolysis effects, aluminium carbonate is unknown.
The mixing of sodium carbonate solution and a solution of an aluminium
salt precipitates aluminium hydroxide and liberates carbon dioxide.
2 Al®*(aq) 3CO3" (aq)
+ +
6H,0 = ae. + 6H*(aq)
dt
2 Al(OH),(s) | 3H2CO,(aq) = 3H,O()) + 3CO,(g)
The same is true for a iron(III) salt and sodium carbonate. The reactions
are similar, substituting Fe for Al.

4. Salt of a strong acid and a strong base


A typical case of this kind is sodium chloride. The situation in a solution
of this salt is
Nat(aq) Cl-(aq)
H,O(1) = OH~(aq) + H*(aq)
The acid and base are both strong so that no formation of molecules,
NaOH or HCl, occurs. The ionic equilibrium of water remains undisturbed,
228 Physical Chemistry
there is no hydrolysis, and the solution remains neutral. The salts of
sodium and potassium hydroxide with any of the three mineral acids show
this situation.

Qualitative detection of hydrolysis in a salt solution


A simple method of detecting hydrolysis in a solution of a salt is to employ
a universal indicator. This is a mixture of indicators in solution so chosen
that it will register a definite change of colour for each change of one unit
(sometimes half a unit) of pH. The indicator is added to the salt solution
in a proportion corresponding to the maker’s instructions (usually 4
drops to 20 cm*.) The pH can then be estimated by comparison with a set
of standard tubes of indicator in solutions of known pH. Any departure
from pH 7 indicates hydrolysis, acid hydrolysis for pH less than 7, alkaline
hydrolysis for pH greater than 7. This method cannot be used in coloured
salt solutions.

Quantitative treatment of salt hydrolysis


The general reaction for the hydrolysis of a salt of a weak acid, HA, anda
strong base is
A“ (aq) + H,0() =OH“ (aq) + HA(aq)
The hydrolysis constant, K,, for this case is written as

_ [OH “][HA]
$85 any[ADHieg
The concentration of H,O is virtually constant. By introducing the
quantity [H*] in both numerator and denominator, we can write
[HA]
OFTEN * Teer
= f[OH-][Ht a

1
Ve
Ska
Ke
K
where K,, is the ionic product of water and K, is the dissociation constant
of the acid.
In a similar way, the hydrolysis constant for the hydrolysis of the salt of
a weak base and a strong acid is
Chemical Equilibria IT: Ionic Equilibria 229
It can also be shown that if the salt is derived from a weak base and a
weak acid, its hydrolysis constant is given by

Nature and Choice of Indicators


Nature of an indicator
An indicator is a very weak acid or a very weak alkali. Since the great
majority of indicators are in the former class, the main discussion of
indicators will be stated in these terms. As a very weak acid, the indicator
is of the general type, HA, and ionizes as HA(aq) = H*(aq) + A~(aq)
with the balance of the equilibrium very much to the /eft of the equation.
It is a vital feature of an acidic indicator that there shall be a marked
difference of colour between the undissociated molecule, HA, and the
anion, A~. This colour change is often associated with a change (accom-
panying ionization) from a benzene-type organic structure to a quinone-
type structure, i.e.

HC CH

from <)- to ee Sal Os yO fs


sxacece

HG CH

As the indicator is a very weak electrolyte, its ionization is governed by


the Dilution Law, so that

[H*JIA7] _
[HA] °
If the indicator is in an acidic solution, a relatively high concentration of
H?* is also present in the solution and participates in the above equilib-
rium. The high concentration of H* tends to increase K,. To maintain the
value of K,, [A~] must decrease and [HA] must increase. That is, the
colour of the undissociated molecule is seen, e.g. colourless for the
indicator, phenolphthalein.
If the indicator is placed in an alkaline solution (say sodium hydroxide
solution), it will form its salt, which is highly ionized.
HA + Na*(aq) + OH~(aq) — Nat(aq) + A~(aq) + H,O
Consequently, the colour of the solution changes to that of the anion,
A-, e.g. purple for phenolphthalein.
230 Physical Chemistry
A few indicators are basic, ie. of the type, BOH(aq), ionizing very
slightly as BOH(aq) = B*(aq) + OH~(agq). In this case, there must be a
marked colour difference between the undissociated molecule, BOH, and
the cation, B*. By use of the Dilution Law,

[B*][OH-] _
[BOH]
and an argument as above, it can be shown that, in alkaline solution
(excess of OH~), the colour of BOH predominates. In acidic solution, the
colour of B* is seen as a result of salt formation, e.g.
BOH(aq) + H*(aq) + Cl (aq) > B*(aq) + Cl (aq) + H200)

Change point of indicators


Rearrangement of the equilibrium expression for an indicator which is
a weak acid gives
K, [HA]
[H*] = ——
[A7]
The intermediate colour of a two colour indicator will occur when

[HA] = [A7]

i.e. when
[HA]
Sa
[Av]
and [Hee Ke

Each indicator has its own particular K, value. Consequently, different


indicators change colour at different [H*] concentrations (and pH
values).
In general, it is found that a change of about two units of pH is necessary
to produce the full colour change in a given indicator. The change-points of
a few of the indicators which are in most frequent use are given in the
table on page 231.
The pH values are stated to the nearest integer. It will be seen that these
indicators cover the whole range of pH values from 10 to 3, i.e. the whole
range which is significant for ordinary acid—alkali titration. Litmus is little
used because, being a natural product, it is variable in quality. Bromo-
cresol purple (change-point, pH 5-7) covers roughly the same range as
litmus near the true neutral point of pH 7.
It is important to note that the ‘neutral’ point as shown by most
indicators, i.e. the mid-point of their pH range for complete colour change,
is not the true neutral point of pH 7. For example, the ‘neutral’ of phenol-
phthalein is about pH 9 and, in the absolute sense, is appreciably alkaline;
Chemical Equilibria IT: Ionic Equilibria 231
2
ee ee e
“a elIndicblah
ator
aT SR Mt Colo
i i
urs pH
alain aciainay

phenolphthalein patehi 8
ee eee tee. ib ARR WL VIG
litmus alu
Sie
l an :
SsSay
e

methyl red ane ;

methyl orange ae :

the ‘neutral’ of methyl orange is about pH 4 and is appreciably acidic.


Thus a saturated solution of carbon dioxide or carbonic acid, pH about
6.3, is quite strongly ‘acidic’ to phenolphthalein (change-point pH 10-8)
but is ‘alkaline’ to methyl orange (change-point pH 5-3). These varying
change-points are a valuable feature of indicators, making for flexibility in
their use.

Choice of indicators for titration


The graphs in Figure 15.1 show approximately the change of pH in
solutions when acid-alkali titration is taking place. The graphs cover all
the four possible cases—strong and weak acids and strong and weak
alkalis. The change-points of the three most important indicators are
included for reference.
An efficient indicator is one which, in the titration for which it is
employed, will change colour to give a sharp end-point. That is, it will
move over its full range of colour for the addition of a single drop of the
titrating acid (or alkali) from the burette. With the axes disposed as in
Figure 15.1 this means that the change-point of the indicator must appear
on a vertical part of the curve of pH change, i.e. a part where a maximum
change of pH occurs on the vertical axis for a minimum change of titrating
liquid on the horizontal axis. The end-point shown by the indicator will
then be sharp. This consideration dictates the choice of indicator for each
of the four possible cases stated below.
1. Strong acid-strong alkali Inspection of Figure 15.1 shows that the com-
bined strong acid-strong alkali curves are almost vertical for the entire
pH range 3-11. This range covers the change-points of all three indica-
tors shown. All of these are suitable for use and will show almost
232 Physical Chemistry
identical titration figures at the end-point. This covers cases like
HCI-NaOH, HNO,;-KOH, H,SO,-NaOH.
2. Strong acid-weak alkali Inspection of Figure 15.1 shows that the com-
bined strong acid—weak alkali curve is almost vertical over the range
pH 3-7, only. The indicators, methyl orange and methyl red, have their
change-points on this range and either is suitable, preferably methyl
red. The most important case here is that of HCI-NH,OH.

pH
strong alkali

weak alkali= ~.—~


ca
phenolphthalein< 9

25.0 cm? of acid 25.0 cm® of base


((H*] =01M) 5 10 5 ({OH ]=01M)
oe <—

addition of 25.0 cm? of base ¥ addition of 25.0 cm? of acid


g

5 methyl red
=
- ae .
--~ weak acid 4 methyl orange

strong acid

Figure 15.1. Change of pH in neutralization

3. Weak acid-strong alkali Figure 15.1 shows that the combined weak
acid-strong alkali curves are almost vertical over the pH range 7-11,
only. Phenolphthalein has its change-point on this range and is a
suitable indicator. This covers the titration of weak organic acids,
such as ethanoic or oxalic, with sodium or potassium hydroxide.
4. Weak acid-weak alkali Figure 15.1 shows that no part of the combined
weak acid—weak alkali curve is vertical. Consequently, no indicator will
give a sharp end-point and titration of weak acid by weak alkali is not
possible with an indicator depending on colour change. The graph
Chemical Equilibria IT: Ionic Equilibria 233
shows that the pH is changing most rapidly near the neutral point of
pH 7. But even here the necessary change of pH by two units, which
covers the colour range of an indicator, will require an addition of
titrating liquid over the range A-B. This is several cm? and at no point
is the colour change sharp. Titration of pairs like ethanoic acid-
ammonia, or oxalic acid-ammonia is, therefore, impossible with any
acid—alkali indicator.

Titration of salts of weak acids by strong acids


The most important case of this kind is the titration of a soluble carbonate
by mineral acid, usually sodium carbonate solution by standard hydro-
chloric acid. This is not really a neutralization of the type
H+(aq) + OH-(aq) = H,O(1)
except in so far as the carbonate is hydrolysed. It is mainly the displacement
of the weak acid, carbonic acid, from its salt by the strong mineral acid,
with subsequent decomposition of the weak acid to carbon dioxide and
water.
CO3- (aq) + 2H*(aq) —-> HzCO,(aq)
dt
H,O() + CO.(g)

Consequently, as long as any carbonate remains to react with the hydro-


chloric acid, the highest acidity attainable is that of a saturated solution of
carbon dioxide, i.e. about pH 6.3. If methyl orange is used as indicator
(change-point, pH 5-3), the solution appears to be alkaline and the
indicator remains yellow. As soon as all the carbonate has reacted and a
slight excess of hydrochloric acid appears, pH decreases rapidly, covering
the change-point of methyl orange, which alters its colour to orange (or
pink) to give the end-point.
A similar case is the titration of a soluble borate by dilute hydrochloric
acid. Here, the weak acid is boric acid and the reaction is

B07" (aq) + 2H*(aq) + SH2O(I) > 4H3BO,(aq)

Buffer solutions
A buffer solution is one which is made up to have a particular hydrogen
ion index, pH, and to retain that value of pH in spite of possible accidental
contamination by acid or alkali.
A typical buffer solution for acidic values of pH, say 4-7, contains a
weak acid and its sodium or potassium salt. A mixture of ethanoic acid
and sodium ethanoate is commonly used. (Suitable quantities are quoted
later.) This buffer solution operates in the following way. The sodium
234 Physical Chemistry
ethanoate which, like all salts, is a strong electrolyte, can be taken as fully
ionized, while the weak ethanoic acid is slightly ionized. In the buffer
solution the ionization of the acid is partially suppressed by the common
ion effect of ethanoate ion, so it appears weaker than in a corresponding,
purely aqueous, solution. Water is also ionized in the usual way
(K, = 1 x 10-14 at 298K), but its ions play no considerable part in
this case. The ionic situation is
From sodium ethanoate: Nat(aq) CHs,CO; (aq)
CH;CO,H(aq) = H*(aq) + CH3CO; (aq)
The salt, sodium ethanoate, provides a large reserve of ethanoate ion. The
weak acid, ethanoic, provides a large potential reserve of hydrogen ion,
which it can realize by ionizing as required. If the solution is contaminated
by acid, i.e. additional H*, the reserve ethanoate ion immediately reduces
the effective concentration of the H* to negligible proportions by com-
bining with it to form ethanoic acid molecules.
H*(aq) + CH;CO;(aq) = CHsCO,H(aq)
In the conditions of the buffer solution, the equilibrium position in this
reaction lies overwhelmingly to the right.
If the solution is contaminated by alkali, i.e. additional OH™, it is
immediately combined with H+ from the ethanoic acid,
H*(aq) + OH-(aq) = H,O()
and reduced to negligible concentration. Reserve ethanoic acid can then
ionize to restore the situation to a point different to a negligible extent from
the original one. For example, the addition of 1 cm® of 0.01 M HCl to
1 dm® of a sodium ethanoate-ethanoic acid buffer solution of pH 3.70
produces no change in the pH which can be expressed in three significant
figures. The same addition of acid to water alters the pH of the liquid from
7.0 to 5.0.
Correspondingly, a buffer solution on the alkaline side (pH 7-11) can
be made with a weak base and one of its salts. A mixture of ammonium
hydroxide and ammonium chloride illustrates the principle, but is notmuch
used in practice because of the volatility of ammonia. The salt provides a
large source of the ion, NH;, and the base a large potential source of
OH7 ion.
From NH,Cl NH} (aq) = Cl7(aq)
NH,OH(aq) = NH}; (aq) + OH~(aq)
Contamination by alkali, ie. OH~ ion, is taken up by formation of
molecules of the weak base.
OH-(aq) + NH} (aq) = NH,OH(aq)
Chemical Equilibria II: Ionic Equilibria 235
Contamination by acid, i.e. H* ion, is taken up by the OH™ ion from the
base, which then ionizes to restore the original situation almost exactly.
H*(aq) + OH" (aq) = H,00)
The following approximate calculations illustrate the position quantita-
tively. Suppose a buffer solution contains 0.02 mole of ethanoic acid
ae = 1.8 x 107°) and 0.2 mole of sodium ethanoate per dm*. For the
acid,
CH;CO;][H*
K,= ORC = 1.8 x 10-° mol dm-® (by the Dilution Law)
2

:
ie. +] 2
= 1.8 x 107-5 CH;CO.H]
[CH3CO.H]
a *< [CHCO3]
taking the sodium ethanoate as fully ionized, [CH;CO;z] = 0.2, and
ethanoate ion from the acid as relatively negligible. The ionization of the
acid is slight in the presence of its salt by common ion effect, so that,
approximately, [CH;CO,H] = 0.02. Substituting these figures,

a
+] = 1.8 x 10-8 *x 0.02
202
02
= 1.8 x 10-°
= 10°78 x 10-8 (since lg 1.8 = 0.26 approx.)
= 10-5-74
That is pH = 5.74
This combination (0.2 mole of sodium ethanoate and 0.02 mole of ethanoic
acid per dm®) gives, therefore, a buffer solution of pH 5.74. Other mixtures
used as buffer solutions are citric acid and its sodium salt, sodium carbonate
and hydrogencarbonate, boric acid and borax, the two sodium phosphates,
Na ,HPO, and NaH2PO,.

Sparingly Soluble Electrolytes


Solubility product
This very important concept concerning saturated solutions of sparingly
soluble electrolytes will be discussed first in relation to the special case of
silver chloride, and then generalized later.
Suppose a quantity of pure water is taken at some definite temperature
(usually 298 K). If silver chloride is added to the water, with stirring, and
plenty of time and unlimited silver chloride are available, a point will
ultimately be reached at which the solution becomes saturated with silver
chloride so that no more can dissolve. An equilibrium is then reached of
the type
AgCl(s) = Ag* (aq) + Cl- (aq)
236 Physical Chemistry
Applying the Equilibrium Law,
x = As @ql [Cl @a)]
[AgCl(s)]
The AgCli being a solid means that [AgCl(s)] can be regarded as constant,
so that
K,, = [Ag*(aq)] [Cl @q)]
In this expression, K,, is a constant and is called the so/ubility product of
silver chloride. Note that the ionic concentrations in this solubility product
equation relate to a saturated solution at a definite temperature.
The important factor about the solubility product of a compound is that
it defines the point at which the compound is about to precipitate, because
it relates to conditions in a saturated solution. In the case of silver chloride
at 298K experimental results show that 1 dm® of saturated solution
contains 1 x 10-5 mole of the salt. Since the mole produces 1 mole of
Ag* and Cl-, it follows that
[AgCl] = JAg*] =[@l-]' = 1 x 1022 moldm-*
From this, the solubility product of silver chloride is
{[Ag*][Cl-] = 1 x 107*°*mol*dm-°

This value of the solubility product is significant in the following way.


Suppose equal volumes of 0.05 M silver nitrate solution and 0.05M
sodium chloride solution are mixed thoroughly and instantaneously, so
that each of the salts momentarily becomes 0.01 M by mutual dilution.
Then, for a moment only,
[Agt] = [Cl-] = 1 x 10-2 mol dm=3
so that the product [Ag*][Cl-] is 1 x 10~*. This is greatly in excess of the
solubility product (1 x 1077°) of silver chloride. At once, Ag* and Cl-
are withdrawn from solution as a precipitate, and precipitation continues
till the amounts of these ions left in solution are just sufficient to reach the
solubility product of silver chloride. This is so small that precipitation is
almost complete.
Notice, however, that it is the product of the ionic concentrations which
is significant. Suppose we have a saturated solution of silver chloride,
in which the situation (at 298 K) is [Ag]+ = [Cl-] = 1 x 10-5, so that
[Ag*][Cl-] = K,, = 1 x 10-?° mol? dm-®
Then let hydrogen chloride be dissolved in the liquid up to a concentration
of, say, 0.01 M, so that [Cl-] = 1 x 10~?. This chioride ion participates in
the solubility product relation of silver chloride though it is not derived
from this salt. The chloride ion concentration from silver chloride being
Chemical Equilibria IT: Ionic Equilibria 237
relatively negligible, the ionic concentrations in solution must satisfy the
requirement
[Ag*][Cl-] = [Ag*][10-7] = K,, = 1 x 10-1°
That is, at equilibrium, [Ag*] = 1 x 10-8; so, by dissolving the stated
amount of hydrogen chloride in the solution, the concentration of Ag* is
reduced from 1 x 10-® to 1 x 10-® mol dm~® by precipitation of the
corresponding quantity of silver chloride. This is an aspect of the common
ion effect. The mass of silver chloride precipitated per dm? is (10-5 — 10-°)
[AgCl]. Since 10-8 is relatively negligible, this mass is 10-° x 143.5 g, ice.
0.001435 g of silver chloride per dm*.

General expression for solubility product


Suppose a compound A,,B,, ionizes in the following way:
AmB, = mA"* + nB™-
Each ion operates separately in the solubility product relation, so that the
solubility product of A,,B, is given by
[A™*]™ [BES = Ke

and the square brackets indicate equilibrium concentrations measured in


mol dm-~%,
Examples are:
Ag,CrO,(s) = 2Ag*(aq) + CrOZ-(aq) Ksp = [Ag*}[CrOz7]
PbCl.,(s) = Pb?*(aq) + 2Cl- (aq) K,, = [Pb?*][Cl-}?
Some typical examples of solubility product values are given in the
following table:

Compound Solubility product


at 298 K

AgCl Te Oa
Ag.CrO, ron Fate
ZnS 1x 10774
CoS 2 640577
PbS 4 x 10-28
Mg(OH), 1.2
x 10-22 (291 K)

Applications of the concept of solubility product


The first two applications are concerned with a scheme of qualitative
analysis in which metal ions are separated first into groups. These groups
238 Physical Chemistry
are known as analytical groups and their numbers bear no relation to the
numbers of the groups in the Periodic Table.

1. Precipitation of metallic sulphides in qualitative analysis


Two groups of insoluble sulphides occur—those of Group II, which are
precipitated from a solution acidified with hydrochloric acid, and those of
Group IV, which are precipitated by hydrogen sulphide in the presence of
ammonia, i.e. by ammonium sulphide. The important members of these
groups are:
Group II Group IV
HgS, PbS, CuS, CdS, Bi,S3 NiS, CoS, MnS, ZnS
Differentiation into these groups depends on the following facts.
Hydrogen sulphide is a very weak electrolyte. Considering its ionization,
H.S(aq) = 2H*(aq) + S?-(aq)
the situation is governed by the Dilution Law which requires that
[H*]? [S*7] = K,
[HS]
where K, is the dissociation constant of the electrolyte and has a value in
the neighbourhood of 1 x 10-2. It is obvious from this that the con-
centration of S?~ ion in a solution of hydrogen sulphide is always very low.
In Group II conditions, hydrochloric acid is present, supplying a large
concentration of H*. This participates in the above equilibrium, tending
to raise the value of K,. To maintain K, at its correct level, the concentra-
tion of S?- must fall, and of H,S must rise, i.e. the ionization of HS is
considerably suppressed, so that the concentration of S?2~- is reduced much
below its already low value in solution. This is an example of the ‘common
ion effect’.
The sulphides of Groups II and IV all have low solubility products, but
the solubility products of the Group II sulphides are lower than those of
the Group IV sulphides. Typical cases are:
Group II Group IV
Copper(II) sulphide Zinc sulphide
[Cu?*][S?-] = 3 x 10-48 mol?dm-® [Zn?*][S?-] =1 x 10-24 mol? dm=-®
In Group II conditions (HCI present), the concentration of S2~ is so low
that the higher solubility products of Group IV sulphides cannot be
reached and these sulphides do not precipitate. The much lower solubility
products of Group II sulphides can, however, be reached and (momen-
tarily) exceeded; consequently, the Group II sulphides precipitate.
In Group IV conditions (NH,OH present) the H* concentration is
reduced due to its reaction with the OH™ ions. This results in increased
Chemical Equilibria IT: Ionic Equilibria 239
ionization of the hydrogen sulphide. The concentration of S27 is relatively
high. Consequently the solubility products of Group IV sulphides are
reached and (momentarily) passed and the Group IV sulphides precipitate.
Cadmium sulphide is a borderline case. It is allowed for in the analysis
tables as a Group II sulphide. If, however, the concentration of hydro-
chloric acid is too high, the concentration of S?- may be reduced to a
point at which the solubility product, [Cd?*] [S?-] = K,, cannot be
reached. In this case, cadmium sulphide will not precipitate in Group II.
This is why a sample of the solution should be well diluted and subjected
to passage of H.S before it goes to Group III. The dilution ensures
precipitation of cadmium sulphide, which might otherwise be missed.
The differences between the sulphides of Groups IIA and JIB are
considered in connection with the chemistry of arsenic, antimony, bismuth,
and tin, in Jnorganic Chemistry, the companion volume to this book.

2. Precipitation of metallic hydroxides in qualitative analysis


After the precipitation of the Group II sulphides discussed above, the
remaining metals with ‘insoluble’ hydroxides include Fe, Al, Cr, Ni,
Co, Zn, and Mn. If, however, the precipitating agent used is ammonium
hydroxide in the presence of ammonium chloride, only the hydroxides
of Fe, Al, and Cr precipitate. They constitute Group III. This is explained
as follows.
Ammonium hydroxide is a weak base, i.e. only feebly ionized at the
usual dilution (2M-4M). By the Equilibrium Law, applied to the
ionization NH,OH(aq) = NH;(aq) + OH~(aq),

[NH7][OH™] = K, = 2 x 10-° mol dm~8 (at room temperature)


[NH,OH]
If the salt NH,Cl is present, it is fully ionized and provides a high con-
centration of the ion, NH}. This participates in the above equilibrium.
To maintain the value of K,, the concentration of OH~ must fall and the
concentration of NH,OH must rise. That is, the weak base becomes
weaker and the concentration of OH™~, small at any time, is depressed to
a very low value, an example of the ‘common ion effect’.
The ‘insoluble’ hydroxides named above all have low solubility products.
The Group III hydroxides, however, are much less soluble than the rest.
Consequently, the low concentration of OH~ available in the presence of
ammonium chloride suffices to reach the solubility products of these
hydroxides and they precipitate. The higher solubility products of the
hydroxides of Ni, Co, Zn, and Mn are not attained and these hydroxides
do not precipitate. Manganese hydroxide is a borderline case. If the
concentration of ammonium chloride is rather low, this hydroxide may
precipitate partially in Group III. Manganese will then follow the same
240 Physical Chemistry
course as iron in Group III. Some manganese will, however, always go
forward and appear as sulphide in Group IV.

Solubility of salts of weak acids in dilute strong acids


There are many cases in which the salt of a weak acid is only very slightly
soluble in water, but dissolves readily in a dilute mineral acid. Calcium
ethanedioate is a typical case and will be used in illustration.
If calcium ethanedioate is stirred in water, it will dissolve to the point at
which, in the solution, [Ca?*] [C,0?-] = K,,, where K,, is the solubility
product of the salt at the temperature in question. K,, is very small so that,
of a moderate quantity of calcium ethanedioate, only a very little will dis-
solve and the rest will remain in suspension. When hydrochloric acid is
added, the ionic situation is
From calcium ethanedioate: Ca?* (aq) C027 (aq)
ei
2HCl(aq) = 2Cl-(aq) + £2H*(aq)
tt
H2C20,(aq)
Ethanedioic acid is weak ;consequently unionized molecules of theacid are
produced by utilizing oxalate ion of the salt and hydrogen ion from
hydrochloric acid. The withdrawal of C,02- ion in this way reduces the
product [Ca2+][C20%-] momentarily below the value, K,,. To restore
this value, calcium ethanedioate passes from the suspended material
into solution. More formation of Hz€2O4 molecules occurs, drawing
more calcium ethanedioate into solution. If sufficient hydrochloric acid is
available, this continues till all the calcium ethanedioate is dissolved. The
free C,O03- ion is then insufficient to reach the solubility product of
calcium ethanedioate.
Calcium phosphate provides a similar case, being only sparingly soluble
in water and readily soluble in dilute hydrochloric acid. The relevant
equations are
[Ca?*]° [PO?-]}? = K,, (very small)
From calcium phosphate: 3Ca?* (aq) 2PO$~ (aq)
a

6HCl(aq) = 6Cl-(aq) + 6H*(aq)


I
2H3PO,(aq)
(a weak acid)
Another application of solubility product ideas occurs in the following
connection. Many cases are found in which materials (usually salts) are
almost insoluble in water but pass readily into solution in water containing
certain other dissolved chemicals. One of these will be discussed in detail;
others will be stated in outline which can be filled in by similar development.
Chemical Equilibria IT: Ionic Equilibria 241
Why silver chloride is almost insoluble in water but readily soluble in ammonia
solution
Ifsilver chloride (about 5 g) is put into water (about 100 cm‘), the salt
will dissolve until the product of the ionic concentrations in solution
reaches the solubility product of silver chloride, i.e. until

[Ag*][Cl-] = Kp = 1 x 1072 mol? dm-® at 298 K


This value is so small that almost all the silver chloride will remain as a
solid. If, however, ammonia is added, the solid material will quickly
‘dissolve’. The reason for this is as follows.
When ammonia is added, a complex cation is formed.

Ag* (aq) + 2NH,(aq) = [Ag(NHs)z] * (aq)


The silver ion for this complex ion formation is taken from the solution
and no longer participates in the solubility product equilibrium of silver
chloride. Consequently, K,, is no longer attained in the solution. To
restore the value of K,,, silver chloride passes from the precipitate into
solution as Ag* and Cl ions. The ion, Ag*, at once forms a more complex
ion with the ammonia, which in turn causes more silver chloride to dissolve
in the attempt to restore the value of K,,. This continues (if sufficient
ammonia is present) until all the silver chloride precipitate has dissolved
and the concentration of Ag* in the solution is so small that the solubility
product of silver chloride cannot be reached. The silver chloride is then
said to have ‘dissolved’ in ammonia though, in fact, a definite chemical
reaction has occurred. The final ‘solution’ contains the silver almost
entirely as the ion, [Ag(NHs)2]*.
Most of the cases in which a material is almost insoluble in water but
‘soluble’ when the water contains another chemical are explained by
complex ion formation as above. The complex is sometimes on the cation,
sometimes on the anion. The following is a selection of such cases stated in
outline. Each can be fully argued as in the case of ammonia and silver
chloride above.

1. Silver cyanide: almost insoluble in water; readily soluble in potassium


cyanide solution
The solubility of silver cyanide in water reaches its limit when [Ag*][CN™]
= K,, (very small). When potassium cyanide is present, a complex anion
is formed:
Ag*(aq)
ik
+ CN-(aq) + CN-(aq) = [Ag(CN)2]“
S tS = See
(aq)
from AgCN from KCN
242 Physical Chemistry
2. Copper(II) hydroxide: almost insoluble in water; readily soluble in excess
of aqueous ammonia
The solubility of copper hydroxide reaches its limit in water when
[Cu?*][OH-]? = K,, (very small). When ammonia is present, a complex
cation is formed:
Cu?*(aq) + 4NH,(aq) = [Cu(NHs)«]?*
(aq)

3. Mercury(II) iodide: almost insoluble in water ;readily soluble in excess of


potassium iodide solution
The solubility of Mercury(ID iodide in water reaches its limit when [Hg?*]
[I-]? = K,p (very small). When potassium iodide is present, a complex
anion is formed:
Hg’* (aq) + 21" (aq) + 21" (aq) = [Hgl.]?" (aq)
from Hgl, from KI

4. Iodine: sparingly soluble in water ; readily soluble ina concentrated solution


of potassium iodide
Iodine dissolves in water till the equilibrium
I,(s) = I,(aq)
is satisfied. This equilibrium lies very much to the left and very little iodine
dissolves (relative to the amount of water). Note that no solubility product
issue arises here; the concept of solubility product is not applicable to
an element.
When potassium iodide is present, a complex anion is formed.
I,(aq) + I” (aq) = I; (aq)
5. Lead chloride: sparingly soluble in cold water; considerably more soluble
in concentrated acid
The solubility of lead chloride in water reaches its limit when [Pb?*][C1-]?
= K,, (small in cold water). In concentrated hydrochloric acid, a complex
anion is formed:
Pb?* (aq) + 2Cl-(aq) + 2Cl-(aq) = [PbCl,]?~ (aq)
Cee
from PbCl2

Adequate presentation of each of these cases requires a fully developed


argument on the lines shown for the case of silver chloride and ammonia
earlier.
16. Chemical Equilibria II: Redox

Originally the term oxidation was applied solely to cases in which oxygen
was gained, and reduction to those in which oxygen was lost, e.g.
Oxidation: PbS(s) + 20.(g) > PbSO,(s)
Reduction: 2KCIO,(s) > 2KCI(s) + 30.(g)
The two processes could occur together, e.g.
CuO(s) + H.(g) — Cu(s) + H,0()
The copper(II) oxide was reduced; the hydrogen was oxidized.
With the coming of electrolysis in the early nineteenth century, oxygen
was frequently found as an anode product, opposed to hydrogen as a
cathode product, as in the electrolysis of dilute sulphuric acid or sodium
hydroxide solution. For this reason, oxygen and hydrogen appeared to be
rather special chemical opposites, so that loss of hydrogen came to be
considered as comparable to gain of oxygen. Loss of hydrogen was then
classed as oxidation and gain of hydrogen as reduction. For example, in
the reaction
HS(g) + Cla(g) > 2HCl(g) + S(s)
the hydrogen sulphide was said to be oxidized by loss of hydrogen and
the chlorine to be reduced by gain of hydrogen. No oxygen is involved.
With the further growth of chemical knowledge, many cases were
recognized in which a metallic element exercised two valencies (possibly
more) and formed two (or more) series of compounds. Iron, forming the
iron(II) series (with iron divalent) and the iron(III) series (with iron
trivalent), is a well-known case. Here the conversion of iron(II) oxide,
FeO, to iron(II) oxide, Fe,O3, is a clear case of oxidation by gain of
oxygen. Since all iron(II) compounds correspond to iron(II) oxide and
show iron with valency 2, and all iron(III) compounds correspond to
iron(II) oxide and show iron with a valency 3, the conversion of any
iron(II) compound to any iron(III) compound came to be regarded as
oxidation (and vice versa for reduction), e.g.
2FeCl,(s) + Cl.(g) > 2FeClg(s)
GrondD chloride oxidized to iron(ID chloride)

Fe,(SO,)s(aq) + H2S(g) > 2FeSO,(aq) + H2SO,(aq) + S(s)


Giron(ID sulphate reduced to irondD sulphate)
244 Physical Chemistry
In such cases, oxidation involves an increase (and reduction a decrease)
in the valency of the metal. A similar case is the tin(II), tin(IV) inter-
relation.
oxidation
tin(I) compound ———~ tin(IV) compound
reduction

In 1880, ionic ideas entered chemistry. It was then quickly seen that in
cases like the iron(II)-iron(III) relation, oxidation involved an increase in
the proportion of the electronegative constituent of a compound, i.e. the
anion. This can be seen from the examples
iron(II) chloride iron(II) chloride
Fe? + 2cl- oxidation Fe? + 3Cl-

tin(II) chloride ‘eduction | tin(IV) chloride


Snz22Gle Sn**4Cl-

In the late nineteenth century, therefore, oxidation may be said to have


included the following ideas:

1. Increase of oxygen content.


2. Decrease of hydrogen content.
3. Increase in the proportion of the anion, with increase in the valency of
the metal present.
Since the establishment of theories of atomic structure, oxidation and
reduction have come to be regarded mainly as electronic phenomena in
the following sense:
Oxidation is the process of electron-loss.
Reduction is the process of electron-gain.
These two processes are complementary and must occur together.
An oxidizing agent is an electron-acceptor.
A reducing agent is an electron-donor.

As the following discussion shows, this much simplified concept includes


all the older ideas. In the light of it, consider the oxidation of iron(II)
chloride to iron(III) chloride by chlorine.

As a molecular equation: 2FeCl,(aq) + Cl.(aq) > 2FeCl,(aq)


Tonically: 2Fe?* (aq) + Cl.(aq) 2Fe**(aq) + 2Cl-(aq)
This ionic equation summarizes two processes.

1. Each iron(II) ion Joses one electron. It is oxidized and so acts as a


reducing agent. The valency of iron increases from 2 to 3.
2Fe?* (aq) — 2e~ > 2Fe°* (aq)
Chemical Equilibria III: Redox 245
2. The electrons are accepted by chlorine atoms. The chlorine is reduced
and so acts as an oxidizing agent. The proportion of the anion, Cl-, is
increased.
Cl,(aq) + 2e- > 2CI-(aq)
These ideas are quite general. The following are common oxidizing agents.
Notice how, in each case, they operate by accepting electrons (supplied by
a reducing agent).

Oxidizing agents
Chlorine: Cl.(aq) + 2e~ — 2Cl- (aq)
Bromine: Br.(aq) + 2e- — 2Br~ (aq)
Iodine: I,(aq) + 2e7 — 21" (aq)
Potassium manganate(VIJ) (in acidic solution):
MnO; (aq) + 8H*(aq) + 5e~ > Mn?*(aq) + 4H,O(1)
Potassium dichromate(VI) (in acidic solution):
Cr,02-(aq) + 14H*(aq) + 6e7 — 2Cr?*(aq) + 7H,O(1)
Iron(III) salts: Fe®* (aq) + e~ — Fe?*(aq)
Mercury(II) salts (two stages of reduction—to mercury(I) salts and
to mercury)
2Hg?* (aq) + 2e7 —> Hg,”* (aq)
Hg3*(aq) + 2e7 > 2Hg(I)
Hydrogen ions: 2H*(aq) + 2e~ > H.(g)
Hydrogen peroxide: H,O,(l) + 2H*(aq) + 2e~ > 2H,0()
Manganese(IV) oxide (in the presence of acid):
Mn0O,(s) + 4H*(aq) + 2e- > Mn?*(aq) + 2H,0()
The following are common reducing agents. Notice how, in each
case, they operate by donating electrons (which are accepted by an
oxidizing agent).

Reducing agents
Iron(II) salts: Fe?* (aq) —> Fe**(aq) + e7
Tin(D salts: Sn?*(aq) — Sn*t (aq) + 2e7
Metals: X(s) > X"*(aq) + ne~

Hydrogen sulphide: H,S(aq) = 2H*(aq) + S*”(aq);


alae hil ? S?-(aq) —> S(s) + 2e7
Sulphurous acid: SO2-(aq) + H,O(1) > SOf (aq) + 2H*(aq) + 2e7
246 Physical Chemistry
Iodides: 21- (aq) — I,(aq) + 2e7
Ethanedioates: C,02- (aq) > 2CO,(aq) + 2e7
Sodium thiosulphate: 28,027 (aq) — 8,027 (aq) + 2e7

These oxidizing and reducing agents do not necessarily all interact with
each other. However, when observations indicate that such a reaction has
occurred, or redox potentials (see page 248) are used to predict that a
reaction is likely to occur, the complete ionic equation for the reaction is
obtained by multiplying the two half equations by appropriate numbers
and then adding them together so that the electrons cancel out. For
example, when hydrogen sulphide is bubbled into a solution of an iron(IID)
salt, a yellowish precipitate is observed. This indicates that the sulphide
ions have been oxidized to sulphur and it is likely that the iron(IIJ) ions
have been reduced to iron(II) ions. The two ion-electron half equations are

Fe®*(aq) + e~ — Fe?*(aq)
S?- (aq) —> S(s) + 2e7

The complete ionic equation is obtained by multiplying the first of these


equations by two and then adding it to the second equation. The equation
for the reaction is

2Fe** (aq) + S?- (aq) > 2Fe?*(aq) + S(s)


Oxidation Numbers
As an alternative to the ion-electron half equation method of interpreting
redox reactions, a system based on oxidation numbers may be used. It has
the advantage that it can also be applied to substances which do not exist
in the form of ions.
The oxidation number of an element indicates the oxidation state of that
element in a particular compound. An element in an uncombined state has
an oxidation number of zero. When in the form of a simple ion in a
compound its oxidation number is equal to the charge on the ion.
In the reaction

2Na(s) + Clo(g) — 2NaCl(s)

the oxidation number of sodium changes from 0 to +1 and that of chlorine


changes from 0 to —1. The increase in the oxidation number of sodium
when it is oxidized is exactly balanced by the decrease in oxidation number
of chlorine when it is reduced.
When assigning oxidation numbers to elements in covalent compounds
or more complex ions, it is necessary to fix arbitrarily the oxidation
numbers of certain elements.
Chemical Equilibria IIT: Redox 247
Rules for assigning and using oxidation numbers
1, The oxidation number of an uncombined element is zero.
2. The algebraic sum of the oxidation numbers of the constituents of a
compound is zero.
3. The algebraic sum of the oxidation numbers of the constituents of
an ion is equal to the charge on the ion.
4. In order to apply the system to covalent compounds and more complex
ions the following invariable oxidation numbers are adopted:
F is —1,
O is —2 except when combined with fluorine and in peroxides,
H is +1 except in metal hydrides.
5. The total change in oxidation number during a redox reaction is zero.
The increase in oxidation number of the species being oxidized is
exactly balanced by the decrease in oxidation number of the species
being reduced.
The following examples will serve to illustrate the system.

Compound or ion Oxidation numbers of constituent elements

CaCl, Ca +2 cl —1
MgO Mg +2 O —2
P,O,, P +5 O —2
H,O H +1 oO —2
ClO; cl +5 O —2
H,O, H +1 Oo —-1
F,O, F —1 O +1
NaH Na +1 H —1

Oxidation numbers may be used to balance redox equations. An acidified


solution containing dichromate ions will oxidize iron(II) ions to iron(II),
the dichromate itself being reduced to chromium(III). The oxidation
number of the chromium is being reduced and that of iron(II) is being
increased. The total increase must be equal to the total reduction.
The oxidation number of The oxidation number of
chromium in (i) Cr,02- is +6 iron in (i) Fe?* is +2
(ii) Cr+ = is +3 (ii) Fe** is +3
Total reduction = 6 Increase = 1
Therefore, for every mole of Cr.02~ which is reduced, 6 mole of Fe?* must
be oxidized and the balanced equation is
Cr,02-(aq) + 6Fe?*(aq) + 14H*(aq) > 2Cr°* (aq) +
ae 6Fe®* (aq) + 7H,0()
248 Physical Chemistry

The oxidation numbers of H and O do not change during the reaction.


Oxidation numbers may also be used to predict the oxidation state of a
product. For example, it can be shown by experiment that 1 mole of
bromate ions will oxidize acidified iodide ions to produce 3 mole of iodine
molecules. The increase in oxidation number of the iodine must be balanced
by the decrease in oxidation number of the bromine. When 6I- changes
to 31,, the change in oxidation number is from —1 to 0 for each I, which
gives a total increase of 6. In BrO; the bromine has an oxidation number
of --5 and if it is to decrease by 6 it must change to an oxidation state of
—i, ie. to Br~. The complete equation is

BrO; (aq) + 61~(aq) + 6H*(aq) — 31,(aq) + Br~(aq) + 3H,0()

Electrode Potentia!s
A metal dipping into a solution containing its ions establishes an equilib-
rium with its ions. For zinc,

Zn(s) = Zn?*(aq) + 2e-

The electrons remain on the metal and so there is a difference of potential


between the metal and the solution. It is known as the electrode potential of
the metal-metal ion system and is a measure of the tendency of the metal
to provide electrons, i.e. to act as a reducing agent. Any attempt to
measure this potential difference will entail the introduction of another
electrode which will exert its own electrode potential. It is not possible to
determine absolute values of electrode potentials, but by employing a
standard reference electrode it is possible to establish an order of oxidizing
and reducing tendency for electrode systems.

Hydrogen Electrode and Standard Redox Potentials


The hydrogen electrode is used as the reference electrode. It consists of
hydrogen gas, at one atmosphere pressure, bubbling over platinum coated
in electrically deposited platinum and dipping into a solution which is 1 M
with respect to hydrogen ions.
Cells are constructed with the hydrogen electrode as one half and a
metal dipping into a 1 M solution of its ions as the other half. The two
halves are connected by a salt bridge which is made from a piece of filter
paper soaked in potassium nitrate solution (Figure 16.1).
The maximum potential difference, known as the e.m_f., of a cell is then
determined by means of a potentiometer or a high-resistance voltmeter.
The e.m_f. is the algebraic sum of the electrode potentials of the two half-
cells. As the hydrogen half-cell is the reference system, its electrode poten-
tial is taken to be zero and the e.m.f. of the cell is said to be the standard
Chemical Equilibria IIT: Redox 249
electrode potential of the other half-cell. In the cell shown in Figure 16.1
the two equilibria involved are, in the hydrogen half-cell
H.(g) = 2H*(aq) + 2e7 (1)
and in the zinc half-cell

Zn(s) = Zn?*(aq) + 2e7 (2)


The zinc has the greater tendency to form ions. Therefore, equilibrium (2)
moves to the right and electrons pass round the external circuit from the
zinc to the platinum and equilibrium (1) absorbs electrons by moving to

hydrogen ——=—

1M solution
of H* (aq)

platinum
Figure 16.1.

the left. The e.m.f. of the cell is found to be 0.76 V. Conventionally the
current is said to flow from the positive terminal of a cell but in fact
electrons flow in the opposite direction, that is, from the negative. The
zinc is the negative in the above cell and the zinc-zinc ion system is said to
have a standard electrode potential, E®°, of —0.76 V.
When a copper half-cell is substituted for the zinc, the electrons flow
from the hydrogen half-cell and the copper is the positive half-cell. The
e.m.f. of the cell is 0.34 V and the copper-copper(I]) ion system is said to
have a standard electrode potential of +0.34 V.
All the half-cell reactions are redox reactions and so it is common for
standard electrode potentials to be referred to as standard redox potentials.
A table of standard redox potentials is given on page 250. The order in
which the systems appear in this table is known as the electrochemical
series.
250 Physical Chemistry
To avoid drawing diagrams or writing lengthy descriptions, cells are
represented by inserting the appropriate formulae in the following scheme.
electrode|solution|solution|electrode
salt
bridge
The sign of the e.m.f. of a cell is taken as the sign of the right-hand
electrode as it is written. The two cells which have been considered so far
are represented as follows:
Pt, H.(g)|2H* (aq)|Zn?*(aq)|Zn(s) E® for this cell = —0.76 V
Pt, H,(g)|2H* (aq)|Cu?*(aq)|Cu(s) £° for this cell = +0.34 V

Standard Redox Potentials ES(V)

Oxidized Reduced
species species Half-cell reaction E°(V)

K* K K* (aq) + e~ = K(s) —2.92


Cath Ca Ca?*(aq) + 2e7 = Ca(s) — 2.87
Na* Na Na*t(aq) + e~ =Na(s —2.71
Mg?+ Mg Mg?*(aq) + 2e7 = Mg(s — 2.38
A+ Al Al®*+(aq) + 3e7 = Al(s) — 1.67
Ae Zn Zn?*(aq) + 2e— = Zn(s) , — 0.76
Fe?+ Fe Fe?+(aq) + 2e7 = Fe(s) : — 0.44
Sn?* Sn Sn?*+(aq) + 2e— = Sn(s) —0.14
Pbaa Pb = Pb?* (aq) + 2e- = Pb(s) — 0.13
Ht H, 2Ht(aq) + 2e— = H,(g) 0.00
Sn?#* Sn?*+ Sn*+(aq) + 2e7 = Sn?* (aq) + 0.15
(ip Cu Cu?t(aq) + 2e- = Cu(s) + 0.34
lo I- — Ig(aq) + 2e7 = 217 (aq) + 0.54
Oz H2O2 O2(g) + 2H* (aq) + 2e~ = H,O2(aq) + 0.68
Fe?+ Fe?+ Fe®+(aq) + e~ = Fe?*(aq + 0.77
Hg?* Hg2* 2Hg?*(aq) + 2e— = Hg2* (aq) + 0.91
Bra Br-_ _ Bra(aq) + 2e~ = 2Br- (aq) + 1.09
IOs Ip | 210g (aq) + 12H
*(aq) + 10e~ = Ia(aq) + 6H20(I) +1.19
Cr,07- Cr?* Cr077 (aq) + 14H* (aq) + 6e7 = 2Cr3+* (aq) + 7H2O(I) + 1.33
aes pe Mie any eH aa ae + 1.36
nO; n nO, (aq) + aq) + 5e~- = Mn?*(aq) + 4H,O(I :
F, F- Fa(g) + 2e- = 2F- (aq) isa pi +284
lon-ion Systems
A half-cell may be constructed from an inert electrode (platinum) dipping
into a solution containing two types of ions, one of which is the reduced
form and the other the oxidized form of a redox equilibrium. For example
the system represented by the equilibrium ;
Fe?* (aq) = Fe*+(aq) + e7
is capable of providing or absorbing electrons and can be used in a half-
cell. By putting such half-cells with the hydrogen half-cell it is possible to
Chemical Equilibria III: Redox 251
determine a set of values of standard redox potentials for ion-ion systems
(see table on page 250).
As previously stated, standard electrode potentials refer to the potentials
obtained when molar solutions of ions are used. If the concentrations of
the ions concerned are not molar, then the electrode potential of that
half-cell (Z) will not be equal to the standard electrode potential (E°). The
relationship between the two electrode potentials is given by the Nernst
equation, which is
RT #3
E= E+ + —
oF In [ion]

for a metal—metal ion half-cell and for an ion—-ion half-cell

parr epee Leeed ante]


zF [reduc species
,
is the gas
is theRchange
zwhere
constant in J K~? mol", Tis the temperature in Kelvin
in charge involved, F is the Faraday constant in coulombs
mol-!, and the concentrations are measured in mol dm-?.

Applications of Standard Redox Potentials


1. Predicting an e.m.f. value of a cell
The positions, with respect to hydrogen, of two half-cells in the series
may be used to predict the e.m.f. of a cell. For example, the systems
Zn?*(aq)|Zn(s) and Pb?*(aq)|Pb(s) are 0.63 V apart in the series. A cell
constructed from these two half-cells will be capable of producing an
e.m.f. of 0.63 V. The systems Zn?*(aq)|Zn(s) and Cu?*(aq)|Cu(s), being
on opposite sides of the hydrogen system, are 1.1 V apart in the series.
As previously mentioned the sign of the e.m.f. of the cell is that of the
right-hand electrode as the cell is written, thus
Zn(s)|Zn?*(aq)|Pb?*(aq)|Pb(s) E£° = +0.63 V
whereas
Pb(s)|Pb?*(aq)|Zn?*(aq)|Zn(s) E° = —0.63 V
The above discussion may be summarized by the statement:
eo - pe _ re
Eve = EGight-hand half-cell) Evett-hand half-celD

2. Predicting whether or not a redox reaction is likely to occur


Simple displacement reactions show that a metal which is higher in the
series will displace a metal which is lower in the series from a solution
containing its ions, Such observations are consistent with redox potentials;
252 Physical Chemistry
a system which is high in the series (high negative E®) is a strong reducing
agent and one which is low (high positive E®) will be a strong oxidizing
agent. Thus, by referring to the series, one might predict that the reactions
Zn(s) + Cu?*(aq) > Zn?*(aq) + Cu(s)

and
Cl,(g) + 2Br~ (aq) > Br2(aq) + 2CI(aq)
are likely to occur and that the reaction
Bro(aq) + 2F- (aq) > F.(g) + 2Br- (aq)
is not likely to occur.
In general terms, if the reduced form of a system which is higher in the
series is mixed with the oxidized form of a system which is lower in the
series then a reaction is likely to occur.
As mentioned on pages 193 and 207 the fundamental factor which
determines whether or not a reaction is energetically favourable is the
change in free energy. This is related to the e.m.f. of the cell under standard
conditions by the equation,
AG? = —zFE?

where z is the number of moles of electrons transferred and F is the


Faraday constant. As with all predictions of feasibility based on energy
changes for the complete reaction, no indication of the rate is implied. The
following are some reactions which could be predicted as feasible from the
standard redox potentials of the half reactions and which give observable
results in a short time.
(a) Chlorine and bromine oxidize iron(ID salts in solution.
2Fe** (aq) + Cl,(aq) or Br.(aq) — 2Fe**(aq) + 2Cl-(aq) or 2Br~(aq)
The halogen colour is discharged and the green solution changes to
yellow.
(b) Iodine (in potassium iodide solution) oxidizes tin(II) chloride (in
moderately concentrated hydrochloric acid).
The brown colour of iodine is discharged, leaving a colourless liquid.
I,(aq) + Sn?*(aq) — 21-(aq) + Sn**(aq)
(c) The metals with more negative E® values are oxidized by the hydrogen
ion of dilute mineral acids.
Zn(s) + 2H*(aq) > Zn?*(aq) + H4.(g) ) Ditute sulphuric and
Mg(s) + 2H*(aq) > Mg?*(aq) + H.(g) | hydrochloric acid
2Al(s) + 6H*(aq) > 2A1°*(aq) + 3H2(g)} Dilute hydrochloric acid only
Chemical Equilibria III: Redox 253
(d) Potassium manganate(VII) in acidic solution oxidizes iron(II) salts in
acidic solution.
5Fe?* (aq) + MnO; (aq) + 8H*(aq) > 5Fe®+(aq) + Mn?+(aq)
+ 4H,O()
The purple colour of the manganate(VII) is discharged and the green
iron(II) solution turns to a yellow iron(III) solution.
(e) Potassium dichromate(VI) in acidic solution oxidizes iron(II) salts in
acidic solution.
6Fe?* (aq) + Cr.0?- (aq) + 14H*(aq) — 6Fe?* (aq) + 2Cr?* (aq)
+ 7H,0()
The orange-yellow dichromate(VI) solution is reduced to a green
solution of a chromium(III) salt. The green iron(II) solution turns
yellow, but this change will be hidden by the other.
(f) Tin) chloride (in moderately concentrated hydrochloric acid)
reduces an iron(III) salt in acidic solution.
Sn?*(aq) + 2Fe** (aq) — Sn**(aq) + 2Fe**(aq)
The solution changes colour from yellow to green.
17. Rates of Chemical Reactions

The rate of a reaction may be expressed in terms of the rate of decrease in


concentration of a reactant or the rate of increase in concentration of a
product. As will be discussed later, investigations into rates of reactions
may yield useful information which can, for example, help to elucidate
reaction mechanisms.

Experimental Methods
1. Recording the time for a reaction to reach a certain stage
Examples of this method are the classic iodine clock experiment, in which
the time for the iodine-starch colour to appear is recorded, and also the
acid-thiosulphate reaction, in which the time for the sulphur precipitate to
reach a certain density is recorded. Such experiments can give roughly
quantitative information on how rates vary with temperature and con-
centrations of reactants, but they are of limited value as they only provide
an average rate for the period of time over which the experiment has run.
Techniques which provide values for initial rates of reactions are of more
value.

2. The progress of a reaction is followed by chemical analysis


Known volumes are withdrawn from the reaction mixture at regular
intervals of time. Each sample is quickly run into an excess of a reagent
which will arrest the reaction or alternatively the reaction is stopped by
rapid cooling. The sample is then analysed, usually by titration, to
determine the concentration of either a reactant or a product. Examples
of this technique are:

(a) Acid catalysed hydrolysis of an ester


CH3CO2C2Hs(aq) + H,O() => CH,;CO-,H(aq) + C,H;OH(aq)

The concentration of the ethanoic acid can be determined by titration with


standard alkali, suitable allowance being made for the mineral acid which
was added to catalyse the reaction.
Rates of Chemical Reactions 255

(b) Reaction between iodine and propanone (acetone)


I, + CHz;COCH; — CH,;COCH,I + H+ + I-
In this case samples can be withdrawn and, after neutralizing the
hydrogen ions, each sample can be titrated with standard sodium thio-
sulphate solution in order to estimate the iodine left.

3. The progress of a reaction is followed by a physical method


This group of methods has the advantage that they do not require the
removal of portions of the reaction mixture. Examples are:
(a) The volume of a gaseous product is continuously recorded by collecting
the gas in a graduated syringe. The decomposition of hydrogen
peroxide can be investigated by this method.

2H,0.(1) >,2H,0() + O.(g)

(b) The loss in mass of a reaction mixture due to the loss of a gaseous
product.

CaCO,(s) + 2H*(aq) > Ca?*(aq) + H,O(l) + CO.(g)


(c) The rate of change in concentration of a coloured reactant can be
followed by a colorimeter which, by means of a light-sensitive cell,
translates the intensity of light transmitted by a coloured solution into
an electrical measurement. The instrument is firstly calibrated by
finding the ammeter readings for solutions containing known con-
centrations of the coloured reactant. The progress of the reaction may
then be followed by the ammeter readings. The reaction between iodine
and propanone, and that between permanganate and oxalate are both
suitable for this method.

(d) The change in conductivity of a liquid, due to the change in the number
of ions present, may be used to follow a reaction. For example, during
the hydrolysis of a halogenoalkane the conductance of the reaction mix-
ture will increase.
CH3CBrCHs(aq) + H,O(1) ~ CH; COHCH,(l) + H*(aq)
| + Br-(aq)
CHs3 CH;

In addition to the above physical methods, optical activity, refractive


index and volume changes of liquids have been used for appropriate
reactions.
256 Physical Chemistry

Factors Which can Affect the Rate of a Chemical Reaction

These are:
. concentrations of reactants,
. pressure,
. temperature,
. physical state of reactants,
. catalysts.
AWD

In order to find out how the rate of a reaction depends on one of these
conditions all of the other conditions must be kept constant and a series of
experiments carried out in which the condition under investigation is
varied.

Treatment of Experimental Results


1. Concentrations
When the progress of a reaction is followed by determining the concen-
tration of a product at regular intervals of time and the concentration of
the product is plotted against time, a graph similar to that in Figure 17.1
will be obtained.

of
concentration
product

(0) time

Figure 17.1.

The initial rate of reaction, given by the gradient at ¢ = 0, is the most


useful as it is only at this point that the concentrations of all the reactants
are known. At any time after t =0, some of the reactants will have
changed to products. In the general case of
A + B= products
Rates of Chemical Reactions 257
the dependence of the rate on the concentrations of A and B would be
determined by separate investigations. Firstly the concentration of B would
be kept constant and a series of experiments carried out with varying
concentrations of A. A graph would be plotted for each experiment and
the initial rate determined for each starting concentration of A.
Let us say the rate is related to the concentration of A by
rate oc [A]*
rate = k [A]*
Taking logarithms,
lg rate = x Ig[A] + lgk
This is the equation for a straight line and, when lg[A] is plotted against
lg (rate), the gradient will be x and the intercept lg k.
x is known as the order of the reaction with respect to A. That is, if
rate oc[A] it is first-order with respect to A, or if rate oc[A]? it is second-
order with respect to A; k is the velocity or rate constant.
If another series of experiments reveal that the rate is related to the
concentration of B by
rate oc[B]”
then the order of the reaction with respect to B is y. The overall rate
equation is
rate = k[A]” [B]”
and the overall order is x + y.

First-order reactions
It is not always necessary to carry out a series of experiments to show that
a particular reaction is first order. Consider the case of
A — products
Assume the concentration of A in the reaction mixture can be determined
at various times during the reaction. Let the initial concentration of A be
amol dm-* and the concentration at time t be (a — x) mol dm~%,
x mol dm~® having reacted. Then as the reaction is first-order, the rate
at which x increases at a particular time will depend on the value of
(a — x) at that time, i.e.
dx
ne k(a — x)
—=k(a-—x

where k is the rate constant. Re-arranged,


dx
=kdt
(asx)
258 Physical Chemistry
Integrating,
—-In(a—x)=kt+ec
where c is a constant, a value for which is found by considering the case
when ¢ = 0 and x = 0. Thus
c=-Ina
kt = —In(a—x)+ Ina
a
kt= ln——_
"G—2)
Converting to base 10,
a
kt = 2.303 lg———
°@—x)
Thus if a reaction is first-order, a plot of t against 1g [a/(a — x)] will bea
straight line, a value for k being determined by measuring the gradient of
the line.
As lg [a/(a — x)] is a ratio of concentrations, readings such as titration
volumes which are proportional to the concentrations may be used
instead of the actual concentrations.

Half-life of a first-order reaction


Determination of the time for the concentration of a reactant to fall to half
of a particular value (known as the half-life) affords a third method of
establishing that a reaction is first-order. The half-life, 4, is given when
(a — x) = 3a

—(2)

concentration
of
units
reactant
a

PP
W
Oo
WO
Omen
ODN
to 2034. By Gi) 20 hoes iat?
unitsof time
Figure 17.2.
Rates of Chemical Reactions 259
Substituting into

kt = 2,303 lg a

(a — x)
we get
kt, = 2.303 Ig 2
From this equation it can be seen that 4, is independent of the initial
concentration of A. Thus a reaction which has a constant half-life through-
out its path is a first-order reaction (Figure 17.2). Radioactive decay is an
example of a first-order reaction (see page 40).

Second-order reactions
Consider the case of

A + B— products
Let the initial concentrations of both A and B be a mol dm~°. Then at
time t, x mol dm~®* of both A and B will have reacted and the rate of the
reaction will be given by
dx
ai
—) == k(a _— x) 2

which on integration gives


af MEY,
a(a — x)
Thus for a second-order reaction, a plot of t against x/a(a — x) will be a
straight line.
Note that in the case of a second-order reaction, the half-life will depend
on the starting concentrations and hence will not be constant throughout
the reaction. The half-life t, is given when x = 4a. Hence
4a
Ae Tera)
o_o
1
tka

It must be stressed that the order of a reaction is an experimentally


determined quantity. Although it is usually 1 or 2, it is sometimes fractional
and, of course, if the rate of a reaction is independent of the concentration
of a particular reactant then the reaction is zero-order with respect to that
reactant. It is important to realize that the order of a reaction cannot be
deduced from the balanced chemical equation for the reaction. The
260 Physical Chemistry
overall order of a reaction can sometimes be modified by carrying out the
reaction with one or more reactants present in excess. The rate of the
reaction will be independent of changes in concentration of the reactant
which is in excess and hence the reaction will be zero-order with respect to
that reactant.
Most reactions take place in several stages, and if one of these stages is
slower than the others then a measurement of the overall rate of the
reaction will be a measurement of the rate of the slowest step. The slow
step in a reaction mechanism is known as the rate-determining step, The
number of molecules or ions involved in a particular reaction is known
as the molecularity of the reaction. Unlike the order of a reaction the
molecularity must be a whole number, it cannot be fractional or zero.
The order of a reaction provides an indication of the mechanism of the
reaction. The following example illustrates this point and may help to
clarify the distinction between molecularity and order.

Hydrolysis of halogenoalkanes by alkali


(a) Bromoethane
This reaction can be shown experimentally to be first-order with respect to
the bromoethane and first-order with respect to the hydroxide ion, thus
making the reaction overall second-order.
This result suggests that the rate-determining step involves both of the
reactants, i.e.
CH;CH,Br + OH-~ — CH;CH,OH + Br-
This is a bimolecular reaction and the molecularity and the order are
identical.
If the reaction was carried out with one reactant in a large excess, the
reaction would appear to be first-order although still remaining bi-
molecular.

(b) 2-bromo-2-methylpropane
In this case the reaction is first-order with respect to the halogenoalkane
but zero-order with respect to the hydroxide ion. This experimental
result indicates that the rate-determining step does not involve the OH~
ion and the likely mechanism is
(CH3)3CBr — (CH3)3C* + Br- (slow)
followed by
(CH3)3C* + OH-~ — (CH3)3COH (fast)
The first stage is the rate-determining step and the reaction is unimolecular.
Rates of Chemical Reactions 261

2. Temperature
It can be calculated that the rate at which collisions between molecules of
reactants occur is so large that if every collision resulted in a reaction, the
total reaction would be almost instantaneous. In fact, only those collisions
which involve molecules with greater than a certain critical energy result
in a reaction. This critical energy is known as the activation energy.
On average the rate of a reaction doubles for every 10 K rise in tempera-
ture. This large increase, which cannot be accounted for by the increase in
the number of collisions between the reacting molecules, is due to a
doubling of the number of molecules with more than the activation
energy. Figure 17.3 shows how the distribution of molecular kinetic
energies of a gas changes with a rise in temperature. It is clear that the
number of molecules with energy greater than the activation energy, E,,
increases rapidly.

TK

(7+10) K

number
molecules
of

kinetic energy a

Figure 17.3.

If the rate of a reaction is temperature-dependent, then the rate constant


must also vary with temperature. The relationship between the rate
constant and temperature, which is known as the Arrhenius equation, 1s
ke = Aer Balk?
Where A is a constant, R is the gas constant, and E, is the activation
energy which can also be taken as a constant. An alternative form of the
equation is:
E,
Ink Ss eee
Rr InA
262 Physical Chemistry

ey
E,
ea
or Ig k T3RT + lelg A

If a graph of lgk against 1/T is plotted, a straight line is obtained the


gradient of which is equal to —£,/2.3 R. Hence, if the rate constant can be
determined at various temperatures, the Arrhenius equation can be used to
find the activation energy of the reaction.

3. Catalysts
The path of a reaction and the accompanying energy changes may be
represented as in Figure 17.4.

energy

Figure 17.4.

The state of the reactants when they are at the point of maximum energy
is sometimes described as the activated complex. To catalyse a reaction it
is necessary to provide an alternative pathway in which the activated
complex requires a smaller activation energy. This enables a larger pro-
portion of the molecules to achieve the state of activated complex and
hence react.

Catalysis is defined as a change in the rate of achemical reaction brought


about by an agent (the patel which is left unchanged in mass and in
chemical nature at the end of the reaction.
Correspondingly, a catalyst is defined as an agent which alters the rate
of achemical reaction and is left unchanged in mass and in chemical nature
at the end of the reaction.

The following comments illustrate some points of importance which are


not expressly stated in the above definitions but are regarded as additional
criteria of catalysis.
Rates of Chemical Reactions 263
(a) The change in the rate of the reaction may be in either direction, i.e.,
an increase or a decrease. If it is an increase, the catalysis is said to be
positive; if a decrease, the catalysis is said to be negative. Negative
catalysis is also known as inhibition and a negative catalyst as an
inhibitor.

(b) While a catalyst is left unchanged in chemical nature after the reaction,
the definition does not exclude physical changes, which are, in fact,
quite frequent in catalysis. For example, coarsely powdered mang-
anese(IV) oxide used as catalyst in the decomposition of potassium
chlorate becomes much finer in grain.

(c) The definitions above allow intermediate chemical changes involving


the catalyst, provided that it is regenerated at the close of the catalytic
cycle.

(d) In general, a catalyst will operate even when present in very small
proportion to the reagents. For example, one ten-millionth of its mass
of finely divided platinum will’give a measurable increase in the rate of
decomposition of hydrogen peroxide. On the other hand, certain
agents are classified as catalysts when, to be effective, they need to be
present in relatively large amount. For example, in the Friedel-Crafts
reaction,

C,H,() + C2H;Cld) = CeHsC2Hs(l) + HCI(g)


the catalyst, anhydrous aluminium chloride, must reach about 30 per
cent of the mass of the benzene to be effective.
(e) There is some disagreement as to whether a catalyst can initiate a
reaction or merely alter the rate of an existing chemical change. The
general opinion favours the second view. On the other hand, this
opinion involves the acceptance of the idea that a chemical action must
be taken as proceeding even though it cannot be detected over a very
long period (perhaps years). For example, a mixture of gaseous hydro-
carbons and oxygen, which remains apparently unchanged almost
indefinitely at room temperature, can be brought to explosive speed by
the catalyst, platinum black, in a few seconds.

The catalyst and equilibrium


In general,a catalyst has no influence on the final state of equilibrium
reached by the chemical system. It merely alters the time required to reach
the equilibrium, usually shortening the time to a very marked extent. For
example, in the reaction
N,(g) + 3H2(g) = 2NH;(g)
264 Physical Chemistry
the state of equilibrium reached is determined by the choice of temperature,
pressure, and the relative proportions of the nitrogen and hydrogen. The
introduction of a catalyst, such as reduced iron, merely shortens the time
required to reach equilibrium. That is, from the industrial point of view,
the economy is in terms of time, not in terms of percentage yield from
materials. It follows from this experience that a given catalyst must
catalyse the forward and back reactions in a reversible system to an equal
extent. Otherwise a different equilibrium would be reached on introducing
the catalyst. This is also impossible because of energy considerations.
Suppose that, in a reversible reaction, a catalyst could move the equilibrium
in the direction of energy liberation. Then it would be possible, by alternate-
ly inserting and removing the catalyst, to obtain energy indefinitely. This
is contrary to the Law of Conservation of Energy.

Classification of catalysis
Two main forms of catalysis are recognized; these are (a) homogeneous
catalysis, (b) heterogeneous catalysis. A third type, mainly of biological
interest, is known as enzyme catalysis. Enzymes are complex organic
catalysts.

1. Homogeneous catalysis
All the reagents and the catalyst are in the same physical phase, that is, all
are gases or all are liquids (or in solution). Examples of this kind of
catalysis are:
(i) Catalysis of the oxidation of sulphur dioxide and steam to sulphuric
acid in the Lead Chamber process; the catalyst is nitrogen monoxide.
(ii) Catalysis of the esterification of ethanoic acid by ethanol; the catalyst
is dry hydrogen chloride or concentrated sulphuric acid.

2. Heterogeneous catalysis
The catalyst is in a different physical phase from the reagents. The most
common form of this occurs in contact catalysis in which the reagents are
in the gaseous phase and the catalyst is solid. Contact catalysis has great
industrial importance; the following are examples of it.
(i) N.(g) + 3H.(g) = 2NH,(g), catalysed by finely reduced iron at
about 720 K.
(ii) 2SO,(g) + O.(g) = 280;(g), catalysed by platinum (or vanadium
pentoxide) at about 720 K, as a stage in sulphuric acid manufacture.
(iti) 4NHs (g) + 502 (g) > 4NO (g) + 6H,0 (g) catalysed by platinum
gauze at red heat (or, recently, by iron(III) oxide and bismuth oxide) as a
stage in the manufacture of nitric acid.
Rates of Chemical Reactions 265
Theories of catalysis
There are two main theories of catalysis—the Intermediate Compound
Theory and the Adsorption Theory. It is, however, fairly certain that no
rigid division can be drawn between these two theories, the suppositions of
which may often operate together in the same catalysis.

1. The Intermediate Compound Theory


In its simplest presentation, this theory takes the following form. Suppose
a reaction, A + B — AB, is slow and is catalysed by a catalyst, C. The
theory supposes that C first combines with one of the reagents, A, to form
the compound, AC. This compound then reacts with the reagent, B, to
form the compound, AB, releasing the catalyst C, which may then start on
another round of the catalytic cycle, i.e.
A-AnC— AC
AC + B>AB+C
These two reactions together are much faster than the reaction A + B— AB,
to which, in sum, they correspond.
It is obvious that, in being alternately combined and released, the
catalyst probably passes through varying states of oxidation. Consequently,
elements of variable valency, capable of assuming varying oxidation
states, are frequently found in intermediate compound catalysts, especially
transition elements such as Mn, Co, and V. Examples of intermediate
compound catalysis are quoted below. They conform to the general
theory stated above, with, in some cases, rather more complication.
(a) Catalysis of the Lead Chamber process, by nitrogen monoxide.

NO(g) + 40.(g) — NO.(g)


H,O(l) + SO.(g) + NO2(g) > H2SO,(I) + NO(g)
Total result: H,O(l) + SO.(g) + 40.(g) — H,SO,(I)

(b) Catalysis of the decomposition of potassium chlorate, by man-


ganese(IV) oxide (according to McLeod).
2Mn0O,(s) + 2KCIO,(s) > 2KMnO,(s) + Cl,(g) + O2(g)
2KMnO,(s) > K,MnO,(s) + MnO,(s) + O2(g)
K,Mn0,(s) + Clo(g) > 2KCI(s) + MnO,(s) + O.(g)
Total result: 2KCIO,(s) > 2KCl(s) + 30.(g)

It must be remembered, however, that definite evidence of these mech-


anisms is very difficult to obtain. They are usually plausible rather than
proved. Even though each stage can be shown to occur as a single reaction,
it does not necessarily follow that the succession of reactions shown above
266 Physical Chemistry
is the actual mechanism of the catalysis. Actually, in case (b) above the
oxygen always contains a little chlorine and, if only a relatively little
manganese dioxide is used, it can be shown to contain some permanganate
after the reaction. This gives some support to the suggested mechanism.

2. The Adsorption Theory


This theory applies mainly to catalysis of a reaction between gaseous
reagents by a solid catalyst, i.e. to contact catalysis. The theory supposes
that the reagents are adsorbed on to the surface of the catalyst, i.e. brought
into some form of loose association with the outermost one or two
atomic or molecular layers of the catalyst. After the reaction, the products
are released (desorbed) from the catalyst surface, leaving it free to adsorb
more of the reacting substances.
The forces involved in adsorbing the reagents are probably not identical
with those concerned in ordinary chemical combination, but are similar to
those operating to hold together the atoms (or molecules) of the catalyst.
Contact catalysts of this kind are usually most reactive when in a very fine
state of division, a fact which points to the surface as the active part of the
catalyst.
The factors involved in adsorption catalysis are not fully understood.
The adsorption of reacting materials very close together on the catalytic
surface is equivalent to a marked concentration of them and accounts
partly for the catalytic effect. It is also known that adsofption catalysts
are not equally active over their entire surfaces. Spots of exceptional
activity occur and are often associated with discontinuities, e.g. cracks and
irregularities, and crystal boundaries. It should also be noted that adsorp-
tion catalysts also tend to be specific in their action so that the same
reagents give different products under the action of different catalysts. An
example of this is carbon monoxide and hydrogen:

CO(g) + H.(g) “+ HCHO@)


methanal

heated
CO(g) + 2H,2(g) incu lae CH;0H(g)
T203 methanol

These facts cannot be explained on a theoretical basis.


The phenomena of catalyst promotion and catalyst poisoning are well
known in connection with contact catalysts.
A catalyst promoter is an added material which enhances the activity of
a main catalyst, but does not necessarily catalyse the reaction if alone.
Most mixed catalysts consist of a main catalyst and a promoter. The action
of a promoter is not usually very spectacular, but it gives a useful increase
in the catalytic activity. For example, the activity of the main catalyst, iron,
in Haber’s process for ammonia synthesis is substantially promoted by
alumina and potassium oxide; also, the production of methanol from
Rates of Chemical Reactions 267
carbon monoxide is catalysed more effectively by a mixture of zinc and
chromium oxides than by either oxide separately.
There is no adequate theory of the action of catalyst promoters.
Possibly some part of their activity lies in the creation of the discontinuities
alluded to a little earlier.
Catalyst poisoning occurs when a small proportion of impurity in the
reagents gradually suppresses the activity of the catalyst. The impurity is
the catalyst poison. Well-known examples are the poisoning of platinum as
catalyst in the oxidation of sulphur dioxide by arsenic(III) oxide, and the
poisoning of the iron catalyst in Haber’s process by sulphur compounds,
such as HS. In some cases, the poisoning occurs because the poison is
adsorbed on to the catalyst surface and not released again, so that the
surface gradually becomes unavailable for adsorption of the reagents. The
poisoning of platinum by arsenic(III) oxide appears to be of this kind. In
other cases, the catalyst is chemically changed by reaction with the poison.
The poisoning of iron by sulphur compounds falls into this class. The chief
safeguard against catalyst poisoning is a thorough purification of the
reagents.

Negative catalysis
Catalysis is said to be negative when the catalyst reduces the rate of a
reaction, often to the point of virtual suppression. Negative catalysis is,
for obvious reasons, less useful in industry than positive catalysis, which
increases the rate of reaction, but it can be valuable for the suppression of
unwanted reactions. For example, about 2 per cent of ethanol in trichloro-
methane acts as negative catalyst to the oxidation of trichloromethane to
poisonous products by the air:
4CHCI,(1) + 30.(g) > 4COCI,(g) + 2H20() + 2Cl.(g)
Also a little benzene-1, 4-diol will suppress the oxidation of benzaldehyde
to benzoic acid by the air.
2C,H,;CHO() + 0.(g) > 2C,H;COOH()
In some cases at least, negative catalysts are believed to operate by
interfering with chain-reactions by which the overall reaction progresses.
For example, the combination of hydrogen and chlorine is believed to
proceed by the chain-reaction
Cl, — Cl- + Cl: (activated by light)
H, + Cl: > HCl + H-
H: + Cl, ~ HCl + Cl:
Nitrogen trichloride is negatively catalytic to the reaction by absorbing
free chlorine atoms:
NCI, + Cl: — $N2 + 2Cl,
268 Physical Chemistry
Autocatalysis
Autocatalysis is the catalysis of a reaction by one of its own products.
Some examples are:
1. catalysis of the reaction between an oxalate and potassium perman-
ganate by the manganese(II) ion, Mn?*.

2MnO; (aq) + 16H*(aq) + 5C,07° (aq)


—> 2Mn?+ (aq) + 8H,O() + 10CO,(g)
The earliest stages of the oxidation are slow, but the accumulation of
the catalyst, Mn?+, makes the later stages rapid.
2. catalysis of hydrolysis of an ester by H* derived from the acidic product
of the hydrolysis, e.g.
CH,CO,C,H,() + H,0() = CH,CO,H(aq) + C,H;OH (aq)
i
CH,CO;(aq) + H*(aq)
18. Electrolysis and Conductance

Electrolysis
Electrolytes and non-electrolytes
An electrolyte can be defined as a compound which, when molten or in
solution, can conduct electric current with decomposition.
A non-electrolyte is a compound which will not conduct electric current.

The mechanism of electrolysis


Electrolysis can be defined as the decomposition of a compound,
molten or in solution, by the passage of electric current.

For electrolysis, a source of direct current is required. The current is


conveyed to the electrolyte, from some source such as lead accumulators,
by electrodes. The electrode connected to the positive pole of the source of
current is called the anode; the electrode connected to the negative pole is
called the cathode. The conventional current enters the electrolyte by the
anode and leaves by the cathode.
The mechanism of electrolysis can be described by reference to the case
of molten sodium chloride. This salt is completely ionized even in the solid
state, and the ions arrange themselves into a rigid cubic lattice. If the
temperature of the salt is raised, the ions acquire greater energy until a
point is reached at which the lattice is broken down and the salt melts. The
ions, Na* and Cl-, are then mobile, and, in the absence of any electrical
pressure, move at random in the liquid. If two inert conductors are placed
in the liquid, say a platinum wire as cathode and a graphite rod as anode in
this case, the cations, Na*, begin to migrate towards the oppositely
charged pole, the cathode; correspondingly the anions, Cl~, begin to
migrate towards the anode. This movement of ions in both directions
corresponds to the passage of electric current.
When a sodium ion, Na*, reaches the cathode, it acquires from the
cathode an electron and is converted to a sodium atom. Correspondingly,
on reaching the anode, a chlorine ion, Ci~, gives up an electron to the
anode and is converted to a chlorine atom. The chlorine atoms liberated
can then combine in pairs to form molecules.
270 Physical Chemistry
At the cathode
Na*(aq) + e~ —~ Na(s)
At the anode
Cl" (aq) — e~ > 3C1,(g)
This accounts for the electrolytic decomposition products of the salt. In an
aqueous solution of an electrolyte the situation is more complex, because
H+ and OH- ions are always present from the ionization of water. In
aqueous common salt solution, for example, the situation is:
From sodium chloride: Na*(aq) Cl-(aq)
H,0() = H*(aq) + OH (aq)
Consequently four ionic species are present. The question at once arises—
which of them discharge during electrolysis ? This question is decided by a
number of factors of which the most important are:
1. The electrode potential of the element or group.
2. The relative concentrations of the ions.
3. The nature of the electrode concerned.
These factors are considered briefly below.

1. The electrode potential of the element or group. '


If all other factors are constant the order of discharge of ions will be
related to that of the electrochemical series. This order, excluding the less
common elements (and groups) is:

Cations:
most positive electrode potentials discharge first
Ag*t — Cu2+ — H* — Pb?2+ — Sn?+ — Fe?+ >
Zn2+ = Al&+ = Mg?t = Nat = Ca2t ss Kt

most negative
electrode potentials
discharge last
Anions:
OH- > I- > Br- — Cl- + NO; — SO2-
most negative most positive
electrode potentials — electrode potentials
discharge first discharge last

This situation in which, in identical conditions, the ions are discharged


in the order of their discharge potentials is complicated by the following
considerations.
Electrolysis and Conductance 271
2. Relative concentrations of ions
The general situation here is that the order of discharge potentials tends to
be reversed in an ionic pair if the ion which would normally discharge
second is present in relatively high concentration. The most important case
is that of OH~ and Cl-. The order of discharge potentials requires the
preferential discharge of OH™ ion and the later production of oxygen:
At the anode
OH-(aq) — e~ — (OH)
4(OH) — 2H,O(1) + O.(g)

If, however, the concentration of Cl~ is very high, this order is reversed
and the product is almost entirely chlorine.
At the anode
Cl" (aq) — e~ > 3Cl,(g)
In moderate concentrations of Cl~, mixtures of chlorine and oxygen are
produced with the proportion of chlorine decreasing with dilution. If two
ions are far apart in the E.C.S., concentration is an irrelevant factor, e.g.
SO2- cannot discharge before OH~.

3. Nature of the electrode


In some cases, the nature of the electrode can influence ionic discharge.
The most important case of this arises with the ion, Na*. If a common salt
solution is electrolysed with a platinum cathode, H* (from water) discharges
first.
2H* (aq) + 2e7 > H,(g)
If, however, a mercury cathode is used, the energy of discharge of sodium
ion is much reduced and it discharges and forms sodium amalgam.
Na*(aq) + e~ — Na(s)
Na(s) + Hg(l) — NaHg(l)
Hydrogen is not liberated in these conditions.

Electrolysis of aqueous solutions


Electrolytes with common cations, e.g. sodium chloride, sodium sulphate,
sodium hydroxide, tend to behave similarly at the cathode when elec-
trolysed. Similarly, electrolytes with common anions, e.g. concentrated
hydrochloric acid and concentrated solutions of sodium chloride and
potassium chloride, tend to behave similarly at the anode. These common
cathode and anode changes will now be presented for a number of the
better-known ions. The results can be paired to produce the effects of
electrolysing particular solutions. In general, the electrode is taken to be
inert.
972 Physical Chemistry

1. Aqueous solutions containing H*(aq), i.e. dilute acids


The acid supplies H+ (aq) and so does the water of the solution:
H,O(l) = H*(aq) + OH (aq)
Consequently there is only one cation to consider.
At the cathode
Hydrogen ion discharges by electron gain
Ht(aq) +e —-H
The hydrogen atoms combine in pairs to form molecules
H + H— H.(g)
As the H* discharges, the acidity tends to decrease at the cathode.
(There is a corresponding increase of acidity at the anode in the case of
H.SO, and HNO; see SO2- and NO; below). If oxygen is discharged
at the anode its volume is half of that of cathodic hydrogen.

2. Aqueous solutions of sodium (and potassium) compounds, containing the


ion, Nat (or K*)
The sodium compound provides the ion, Na*; water ionizes as
H,O(1) = H*(aq) + OH~(aq). The two cations present are Nat and H*.
At the cathode
H* ion discharges in preference to sodium ion:
Ht(aq)
+e —H

The product is hydrogen gas. As the H+ discharges, the equilibrium of


water ([H*][OH~] = K = 10~**) is disturbed. To restore K,,, water ionizes
further. This puts OH™ into excess and, with Na* attracted to the cathode,
this is equivalent to the production of a solution of sodium hydroxide. If
oxygen is the anode product, its volume is half of that of the cathodic
hydrogen.
Potassium compounds behave in a corresponding way.

3. Aqueous solutions of copper (and silver) salts, containing the ion, Cu?+*
(or Ag)
The copper salt produces the ion, Cu?* ; water ionizes as
H,O() = H*(aq) + OH~(aq)
The two cations present are Cu?+ and H+.
Electrolysis and Conductance 273

At the cathode
Cu? * ion discharges in preference to hydrogen ion. It discharges by electron-
gain.
Cu?*(aq) + 2e- — Cu(s)
Metallic copper is precipitated. The cathode itself is usually a copper rod
or plate.
Silver salts behave in a corresponding way, the cathode usually being a
silver plate or wire.

Ag*(aq) + e~ — Ag(s)

4. Aqueous solutions of soluble hydroxides, notably sodium hydroxide and


potassium hydroxide
The hydroxide produces the ion, OH~ ; water ionizes as
H,O()) = H*(aq) + OH (aq)
In this case, only one anion, OH~, is present.

At the anode
OH7 ion discharges by electron loss. The neutralized (OH) group is not
stable. It generates water and oxygen.
OH~-(aq) — e~ = (OH)
4(OH) — 2H,0(1) + O.(g)
If (as is usual) hydrogen is the corresponding cathode product, its volume
is twice that of the oxygen.

5. Aqueous solutions of soluble sulphates (including sulphuric acid)


The soluble sulphate produces the ion, SO2?~; water ionizes as H,O(l) =
H*(aq) + OH~(aq). Consequently, there are two anions present,
SO?- and OH-.

At the anode
OH- discharges, by electron loss, in preference to SO?~, and then oxygen
is produced as in (4) above. The relation to cathodic hydrogen is the same.
As OH7- discharges, the equilibrium of water
((H*|[OHgs) = Kee al ox 10-4. mol*:dmz*)
274 Physical Chemistry
is disturbed. To restore K,,, water ionizes further, putting H+ into excess.
This, with incoming SO2-, is equivalent to the production of sulphuric
acid (or an increase in its concentration). This increase is balanced, in
suitable cases, by a decrease in acidic concentration at the cathode (see
H?* earlier).
In electrolysing copper sulphate solution, the anode is often a copper
plate. In this case, no discharge of OH~ (or SO{-) will occur. Instead,
copper atoms of the anode will ionize by electron loss,
Cu(s) — 2e7 — Cu?* (aq)
and the ions will pass into solution. This process requires least energy.
The anode slowly dissolves. A corresponding equal precipitation of copper
occurs at the cathode.

6. Aqueous solutions of soluble chloride (including hydrochloric acid)


The soluble chloride produces the ion, Cl~; water ionizes as
H,O() = H*(aq) + OH~(aq)
Consequently, there are two anions present, Cl” and OH-.

At the anode
In very dilute solution, OH~ discharges by electron loss‘in preference to
Cl- Then oxygen is produced, as
OH~ (aq) — e~ — (OH)
4(OH) — 2H,0() + O.(g)
The volume of oxygen is half of that of hydrogen produced in the same
electrolysis at the cathode.
In very concentrated solution, the above order of discharge is almost
completely reversed by concentration effects and almost the sole product
is chlorine.
Cl-(aq) — e~ + 4C1.(g)
After saturation of the solution, the volume of chlorine is equal to that of
hydrogen produced in the same electrolysis at the cathode.
At intermediate concentrations of chloride, the above two effects will
occur mixed.

7. Aqueous solutions of nitrates, especially silver nitrate


The soluble nitrate produces the ion, NO,~; water ionizes as
H,0 (1) = H*(aq) + OH~(aq)
Consequently, there are two anions present, OH~ and NO,°~.
Electrolysis and Conductance 275
At the anode
OH- discharges, by electron loss, in preference to NOjZ, and oxygen is
produced.
OH" (aq)— e~ = (OH)
4(OH) > 211,00) + 0,(g)
The volume of the oxygen is half of that of hydrogen produced at the
cathode in the same eo Sete As OH™ discharges, the equilibrium of
water ({H*][OH- ie = | x 10-14 mol? dm-) is disturbed. To
restore K,,, water ionizes further. This puts H* into excess and, with
incoming NO3, is equivalent to the production of nitric acid (or an
increase in its concentration). This increase is balanced, in suitable cases,
by a corresponding decrease at the cathode (see H* earlier).
In the case of silver nitrate solution, an anode of silver is often used.
In that case, OH~ (or NO5) is not discharged. Instead, silver atoms of
the anode ionize as Ag — e~ = Ag’. The ions pass into solution, so the

hydrogen
(2 vols) cee re?

dilute sulphuric
acid

cathode anode
(platinum) (platinum)

QU00Q
OOQQY

= +
Figure 18.1. Electrolysis of dilute sulphuric acid
276 Physical Chemistry
anode slowly dissolves. There is a corresponding and equal deposition of
silver at the cathode. .
The apparatus of Figure 18.1 is suitable for the electrolysis of dilute
sulphuric acid, sodium (and potassium) sulphate, and sodium (and
potassium) hydroxide in solution. The result shown is the same for all
these solutions and is equivalent to the electrolysis of water. The cathodic
and anodic changes are detailed above.
The same apparatus is suitable for the electrolysis of concentrated
hydrochloric acid and concentrated solutions of sodium (and potassium)
chloride, but the anode must be changed to a carbon rod (since chlorine
produced in this way attacks platinum appreciably). The gaseous products
are then hydrogen and chlorine in equal volumes for all the solutions
mentioned (after chlorine has saturated the solution).
Copper sulphate and silver nitrate are usually electrolysed with copper
plates and silver plates, respectively, as both anode and cathode. The
plates are merely inserted (suitably spaced) into the solutions contained in
glass containers.

Faraday’s Laws of Electrolysis


These laws express the quantitative side of electrolysis. The first law is
expressed in the following way:

The mass of a given element liberated during electrolysis is directly


proportional to the magnitude ofthe steady current used and to the time
for which the current passes; that is, to the quantity of electricity
consumed during the electrolysis.

The first law can be illustrated by passing known steady currents: of


electricity through a solution of copper(II) sulphate between copper
electrodes for observed times. In each experiment the cathode is weighed
before the electrolysis and washed, dried, and weighed after it. The
increase of mass of the cathode (i.e. the mass of copper deposited) will be
found to be directly proportional to the product (steady current x time).
One mole of singly charged ions will require a transfer of 1 mole of
electrons for neutralization. Similarly, 1 mole of doubly charged ions will
require 2 moles of electrons. The transfer of 1 mole of electrons corresponds
to the passing of approximately 96 500 coulombs of electricity, which is
known as the Faraday constant. A coulomb is equivalent to the passage of
1 ampere for 1 second.
Faraday’s second law may thus be stated in the following way:
When the same quantity of electricity is passed through different
electrolytes, the masses of the different substances liberated are directly
proportional to the masses of the substances which require one mole of
electrons (1 faraday) for neutralization.
Electrolysis and Conductance 277
These masses are equal to:
the mass of | mole for singly charged ions,
the mass of 4 mole for doubly charged ions and
the mass of 4 mole for ions with three charges.
These masses have been known as the chemical equivalent masses of the
respective ions.
The second law can be illustrated by passing a suitable electric current
(the exact magnitude of which need not be known) in series for the same
time through, say, dilute sulphuric acid (with platinum electrodes),
copper(II) sulphate solution (with copper electrodes) and silver nitrate
solution (with silver electrodes). The volumes of hydrogen and oxygen
liberated in the first case (see page 275) are observed at known room
temperature and pressure, converted to s.t.p. and rendered as masses from
the relation: 1 dm* of hydrogen has a mass of 0.09 g at s.t.p. or 1 dm® of
oxygen at s.t.p. has a mass of 1.44 g. The copper (or silver) cathode is
weighed before the electrolysis, and washed, dried, and weighed after it.
The increase of mass is the mass of copper (or silver) liberated. It will be
found that if the mass of oxygen liberated is 8.00x g, the masses of hydro-
gen, copper, and silver are 1.008x, 31.8x, and 107.88x g.
8 g is the mass of 0.5 mole of oxygen atoms, 1.008 gis the mass of 1 mole
of hydrogen atoms, 31.8 g is the mass of 0.5 mole of copper atoms, and
107.88 is the mass of 1 mole of silver atoms.

Electrochemical equivalent of an element


This characteristic of an element is defined in the following way:
The electrochemical equivalent of an element is the mass of the element
liberated during electrolysis by the consumption of one coulomb of
electricity, i.e. by the passage of a steady current of one ampere for one
second.

Some Aspects of Conductance of Electrolytes


Except in certain exceptional circumstances, electrolytes obey the same
laws as metallic conductors with respect to electrical resistance (and
conductance); that is, at constant temperature, the electrical resistance of
a conductor is directly proportional to its length and inversely proportional
to its (constant) cross-sectional area. This can be expressed in the equation
l
R=aconstant x F

where R is the resistance (expressed in ohms), / is the length of the con-


ductor (in cm) and a its constant cross-sectional area (in cm”).
With these units, the constant of this equation is called the resistivity
of the material of which the conductor is composed. R becomes equal to
278 Physical Chemistry
the constant when both / and a are unity; that is, the resistivity of the
material is the resistance in ohms between opposite faces of a centimetre
cube of it.
The reciprocal of the resistivity is called the conductivity of the
material and is measured in reciprocal ohms per centimetre, that is, in
ohm-! cm-!. The conductivity is the characteristic of an electrolyte
which is measured (by means of the resistivity in actual practice), but it is
not the most valuable data about the electrolyte, because it varies with
dilution in a two-fold way. In the first place, dilution reduces the number
of ions (the conducting particles) in unit volume of the electrolyte and this
increases the resistance of the electrolyte; in the second place, weak
electrolytes increase their ionization on dilution, which decreases the
resistance of the electrolyte. These two effects are confused in the
conductivity values.
To sort out this confusion, a characteristic of the electrolyte is em-
ployed called its molar conductivity. This is obtained by multiplying the
conductivity of the electrolyte by the number of cm? containing the mass
of electrolyte which requires the passage of | Faraday for neutralization.
The effect of this is to cancel out the simple dilution factor in conductivity
as concentration varies, and, for weak electrolytes, to exhibit the effect of
variation in molar ionization. The molar conductivity of an electrolyte
is usually denoted by /,, where v is the number of cm? containing | mole
of the electrolyte. Consequently, ‘ ;
A, = (conductivity) x v
As we shall see a little later, the molar conductivity of an electrolyte at
infinite dilution (i.e. when | mole of it is dissolved in an infinitely large
volume of solution) is an important characteristic of it, and is denoted by
the symbol, /,,.

Molar conductivity and ionization


It has already been pointed out that, because of the dilution factor
involved in it, molar conductivity is independent of the effect of simple
dilution in reducing the conducting power of an electrolyte. It is affected
by two other factors — by the velocity of the ions and by the extent to
which the electrolyte is ionized, i.e. its degree of ionization. At first it was
assumed that the velocity of a given ion is constant at constant tempera-
ture. In this case the molar conductivity of an electrolyte is directly
related ot its degree of ionization. Since, for all electrolytes, ionization
becomes complete at infinite dilution, the relation was thought to hold:

degree of ionization («) of electrolyte at dilution, v = A»


foo}

when v is the volume (in cm?) containing | mole of the electrolyte, and
Electrolysis and Conductance 279
A, and A,, are the molar conductivity of the electrolyte at dilution, v, and
at infinite dilution.
For weak electrolytes (that is, electrolytes which are only slightly
ionized at the usual working dilutions), this relation still holds and is a very
valuable method of determining degree of ionization. Examples of its
working are given later when some other relevant factors have been
considered. For strong electrolytes, however, which are fully ionized (or
very nearly so) in all ‘dilute’ solutions, the above relation is not now
considered valid. In strong electrolytes of moderate concentration, the
ions are comparatively close together and each tends to interfere with the
progress of oppositely charged ions through the solution, giving the effect
of a mutual slowing down. This reduces the observed value of molar
conductivity below its ‘true’ value. Dilution increases the average distance
separating ions. Consequently they tend to interfere less with each other
and molar conductivity increases until, at infimite dilution, it reaches a
constant maximum value. This change was formerly ascribed to increased
ionization; the present view is that it is rather the result of increased
velocity of the ions, as their effeet on each other lessens. With weak
electrolytes, relatively few ions are present and their mutual interference
effects are almost negligible at all concentrations.
The above fraction, 4,/A., is now called the conductance ratio of an
electrolyte. For weak electrolytes only, it is also (with fair accuracy) a
measure of their degree of ionization.

Change of molar conductivity with dilution

The figures in the table below are typical of the changes of molar con-
ductivity with dilution in the case of a strong electrolyte (sodium chloride)
and a weak electrolyte (ethanoic acid).
The following features are notable about these figures. It will be seen
that, in both cases, molar conductivity increases as the solution becomes

Molarity NaCl CH,CO,H

M 74.2 1.32
0.5M 80.8 2.01
0.2 M 87.5 3.24
0.1M 91.8 4.60
0.05 M 95.5 6.48
0.01 M 101.7 14.3
0.002 M 105.3 30.2
0.001 M 106.3 41.0
280 Physical Chemistry
more dilute. In the case of sodium chloride (the strong electrolyte), how-
ever, the increase is relatively small (from 74.2 to 106.3, or about 43 per
cent), while in the case of the weak electrolyte, the relative increase is
much greater, from 1.32 to 41.0 or about 3000 per cent. Further, in the
case of the strong electrolyte the value of molar conductivity rises very
little between 0.01 M and 0.001 M and is reaching a constant value at
about 0.001 M. In the case of the weak electrolyte, the molar conductivity
is still rising at 0.001 M dilution and shows no tendency towards con-
stancy. These features are characteristic of the behaviour of strong and
weak electrolytes, though it must be remembered that an absolutely sharp
distinction between them cannot be drawn.
The general position is indicated approximately by the graphs in
Figure 18.2.

strong electrolyte

increasing
conductivity
molar
weak electrolyte

: “
increasing dilution 0.001M
Figure 18.2. Molar conductivity of electrolytes

Molar conductivity at infinite dilution


This characteristic of an electrolyte can be determined experimentally for
a strong electrolyte as follows. For a few of its solutions, the conductivity
is first determined. Experimental details are given on page 281. The
conductivity is then multiplied by the number of cm? of solution con-
taining 1 mole of the electrolyte. If a graph is plotted of molar con-
ductivity against the square root of concentration, a straight line is
obtained for low concentrations. If it is extrapolated to zero concen-
tration, i.e. infinite dilution, the value of molar conductivity at infinite
dilution can be read off.
For weak electrolytes, this method is less suitable. A method based on
the following facts is used. The values for molar conductivity of certain
Electrolysis and Conductance 281
, potassium salts and the corresponding sodium salts are (at infinite
dilution):

oo ee Ae ee ee ee
Potassium Sodium Difference

Chloride 130.1 109.0 21.1


Nitrate 126.5 105.3 Dien:
Sulphate 133.0 111.9 INNS

It will be seen that varying the cation produces, within experimental


limits, a constant difference in the molar conductivity. Similarly, the
KCI-KNOsz difference (130.1 — 126.5 = 3.6) is the same as the NaCl-
NaNOg difference (109.0 — 105.3 = 3.7) and the K2SO4_K NOs differ-
ence (133.0 — 126.5 = 6.5) is the same as the NagSO,—-NaNOs difference
(1i1.9 — 105.3 = 6.6). From general results such as these, Kohlrausch
reached the conclusion that the molar conductivity of an electrolyte at
infinite dilution is made up of a contribution from the cation and another
from the anion, and that these contributions are constant and indepen-
dent of each other. That is, the molar conductivity of an electrolyte at
infinite dilution is made up of the sum of two independent quantities
called the ionic conductivities of the two ions concerned. This is known as
the Law of Independent Migration of Ions.

Experimental determination of conductivity and molar conductivity of an


electrolyte
In principle, these determinations require the measurement of the
resistance of the electrolyte in a cell like that shown in Figure 18.3
between similar parallel electrodes of known area at a known distance
apart. In practice, this rather difficult measurement of exact dimensions is
avoided by relying on accurate determination of the conductivity of
potassium chloride solutions made by earlier workers. For example, the
conductivity of 0.02 M KCl at 298 K is 0.002 77 ohm-! cm~! and of 0.1
M KCI, at the same temperature, 0.0129 ohm-! cm~-!. From accepted
figures such as these, it is possible to calibrate a given cell and avoid
measurement of electrode dimensions and distances. This is done by
obtaining its cell constant. For example, the resistance of 0.02 M KCI at
298 K ina given cell (measured as below) was found to be 550.0 ohm, i.e.
the conductance of the electrolyte was 1/550.0 = 0.001 82 ohm=! cm~!.
From these figures, the cell constant of this particular cell is 0.002 77/
0.001 82 = 1.52(3). That is, all conductances found by this cell must be
multiplied by 1.52(3), and then yield true values for conductivity.
282 Physical Chemistry

Figure 18.3. Conductance cell

In measuring the resistance of electrolytes, certain precautions are


necessary. Among them are the following:
Ll; The solution must be made up in specially prepared ‘conductivity
water’. For ordinary measurements, distilled water is redistilled con-
taining potassium permanganate, in apparatus of resistance glass, with
ground joints, or corks protected by tin foil.
. Direct current must not be used because it decomposes the electrolyte.
This introduces the double error of altering the concentration of the
electrolyte and depositing decomposition products on the electrodes,
which will tend to set up a back e.m.f. and so increase the apparent
resistance.
To avoid polarization, the electrodes are usually covered with platinum
black by electrolysing platinic chloride solution in the cell before using
it for measurements. Direct current is passed in each direction for half-
minute intervals till the electrodes are blackened. (Such electrodes are
unsuitable for use with electrolytes on which platinum shows catalytic
activity.)

. Since the conductance of an electrolyte varies with temperature, the


cell should be placed in a thermostat (usually at 298 K).

For a determination, the lay-out is arranged as in Figure 18.4. AB is a


resistance wire of uniform cross-section. The value of the resistance in the
Electrolysis and Conductance 283

source of alternating current


Figure 18.4. Determination of resistance of an electrolyte

resistance box is adjusted by trial till a minimum sound is obtained in the


telephone with the contact C near the middle of AB. Then

resistance of electrolyte in cell = (resistance in box) x ce

_The reciprocal of the resistance gives the conductance of the electrolyte;


this when multiplied by the cell constant, gives the true specific conduc-
tance. A sample calculation is given below.
Electrolyte: 0.1 M zinc sulphate solution (at 298 K).
With 70 ohms in the resistance box, AC = 50.8 cm and BC = (100 —
50.8) = 49.2 cm.
That is, resistance of the electrolyte is
50.8
—— == 72.2
qidie hae 72. olohm

From this,

conductance of electrolyte = _ ohm=?

The cell constant (from the calculation above) being 1.52, the con-
ductivity of the electrolyte is

een SS 0.0211
779 =I cm eal
ohm!

Since 0.1 M means that | mole of the electrolyte is dissolved in 10 000


cm of solution, the molar conductivity of zinc sulphate at this dilution is
0.0211 x 10000 = 211 ohm—! cm?.
ay nme He
os

ci a5 aga o Iie dar, Sapa ake


aailTBN fmmih
} , ayes =~. : ; f

3 x. fant ie san teieayoa Thess”


a sn SOR:

ery acctumta a ovaaetanyon


sap aa
ateeaeBi) aris, roe (i= eb ve
wmiuits i hate Beaded
' Bes Hi ip) RAO WER Gih? =
~_ Gaya hat tj HAR sac
= tat er ee hie. 0s
- pons BS
bog ivitegls call rir vanatelans, ite, SB
. be «¢yp 4
se > ee Th gate
me ie
ve
ere a
:
5
werrave tie
ait eo
' :
coe
¥ . tA cel « ret *
y bah the Ove Ving anontiy hee vAey CAE
; &3

; nh | t cwtl he

y ‘
os
arygrt
- j SE nee
os
wee

VEERas DON
WE Sette 2).
22 MLL DS
ie Oona Aen ape
Py
b Maite 1 : i | a 1a
aeae =
as Spite 27 *)i 1B yt j PO
tyAe AP he
ett) uit,
ing age a
oh:

ADA ERO: LAS = UUAB Ries NARI ls ga ante


Questions on Chapter 1
1. Summarize the propositions which State Mitscherlich’s Law of Isomor-
together constitute the Atomic Theory phism. Briefly discuss its limitations.
of Dalton. Show that they lead to the At a certain period in the nineteenth
conclusions expressed in the Laws of century, the mass of zirconium which
Constant Composition and Multiple combined with 8.00g of oxygen was
Proportions. In the case of one of these known (correctly) to be 22.8g. The
laws, outline a piece of experimental accepted valency of the element was 3.
work by which it can be illustrated. Criticize this situation in the light of the
2. Outline the work of Landolt in fact that zirconium forms a compound
illustrating the Law of Conservation of isomorphous with the silicon compound
Matter. Relate the law to the Atomic known to be K.SiFg and containing
Theory of Dalton. To what extent has about 32% of zirconium. [K = 39,
this law been modified in the light of F = 19.]
recent discoveries? Explain why pure “7, Given the standard ?2C = 12.00,
samples of sodium chloride conform the relative atomic masses of hydrogen,
to the Law of Constant Composition silver, nitrogen, and chlorine are very
though chlorine has two isotopes of important as a basis for fixing other
masses 35 and 37 on the scale of O = 16. relative atomic masses. Describe two
3. Three oxides of nitrogen contain accurate experiments which have been
respectively 46.7%, 30.4%, and 63.6% performed to assist in the determination
of nitrogen. Show that these figures of the relative atomic masses of these
illustrate the Law of Multiple Propor- four elements.
tions. Relate this law to Dalton’s 8. A certain metallic element forms
Atomic Theory. (a) an oxide containing 23.1 % of oxygen,
4. Water, sulphur dioxide, and mag- (b) a chloride containing 43.3% of
nesium oxide contain 88.9%, 50.0%, chlorine, (c) two sulphides containing
and 40.0% of oxygen respectively. 37.6% and 25.6% of sulphur. The
Magnesium sulphide contains 57.1% of specific heat capacity of the element is
sulphur and hydrogen sulphide 94.1% 0.146 Jg-1 K+. What deductions can
of sulphur. Show that these figures you draw from these facts? [O = 16,
illustrate the Law of Reciprocal Pro- Si= 32> Cli= 35to%)
portions. Explain how this law can be 9. The two oxides of titanium contain
deduced from Dalton’s Atomic Theory. (a) 66.7% and (b) 60.0% of titanium.
5. 85.0cm® of hydrogen, measured The specific heat capacity of the metal is
dry at 16°C and 750 mmHg pressure, 0.54Jg-1K~1. Calculate the relative
were liberated by the action of 0.200 g atomic mass of titanium as an average
of a certain metal on excess of dilute derived from these data. What is the
acid. Calculate the mass of metal which percentage composition of the chloride
would liberate 22.4dm* of hydrogen of titanium corresponding to oxide (a)?
at s.t.p. If the specific heat capacity of [@le=="35:5:]
the metal is 0.46 Jg~1K~?, what are 10. Describe an experiment to dem-
the relative atomic mass and valency of onstrate the composition of hydrogen
the metal? [1 dm® of hydrogen at s.t.p. chloride by volume. State its result.
weighs 0.09 g.] Show that this result requires the hydro-
6. State what is meant by isomorphism. gen molecule to contain at least two
285
286 Physical Chemistry
atoms. What further evidence can be hydrogen molecule, prove the relatior
adduced to prove that the hydrogen between the relative molecular mas:
molecule is actually diatomic? and vapour density of a gas. Calculate
11. State (a) Gay-Lussac’s Law of the relative molecular mass of ar
Gaseous Volumes, (b) Avogadro’s Hypo- elementary gas, X, from the data:
thesis. Illustrate (a) by stating the
volume compositions of ammonia and Weight of evacuated globe = 211.470 g
steam. By applying Avogadro’s Hypo- Weight of globe filled with hydrogen
thesis to these compositions, and assum- = 211.830 g
ing the diatomicity of the hydrogen,
nitrogen, and oxygen molecules, deduce Weight of globe filled with X
the molecular formulae of ammonia
= 219.380 g
and steam. (Temperature and pressure remain con-
12. Assuming the diatomicity of the stant.)

Questions on Chapter 2
1. Calculate the mass of each of the following results on analysis: C, 40.0%;
following: (a) 2 moles of calcium atoms, H, 6.7%; O, 43.3%. Find its empirical
(b) 0.5 mole of nitrogen molecules, (c) formula. If its relative molecular mass
0.1 mole of sodium chloride, NaCl, is 60, what is its molecular formula?
(dy 0.2 mole of hydrated copper(I) The compound is found to release
sulphate, CuSO, . 5H.O, (e) 2.5 moles carbon dioxide from sodium carbonate
of ethanol, C.H;OH. solution. Can you suggest a structural
2. Calculate the number of moles in formula for the compound?
each of the following. Clearly state the 5. Barium chloride crystals are known
nature of the species concerned. (a) to be hydrated. 12.2 g of the crystals
3.9g of benzene, CgHe, (b) 41.42 of were heated to constant weight. The
lead, (c) 10.1 g of potassium nitrate, anhydrous barium chloride weighed
KNOsg, (d) 10 kg of magnesium nitride, 10.4g. Deduce the formula of the
MgzNae, (e) 8.6g of gypsum, CaSO. . hydrated crystals.
2H.0O. 6. Calculate the number of moles of
3. Calculate the molarity of each of each reactant and each product in the
the following solutions: reactions described below. Write an
(a) sodium bromide, NaBr, containing equation for each reaction, showing
1.03 g of NaBr in 50 cm® of solution, how you have used the information
(b) hydrogen ion, in a solution contain- provided. (Gas volumes are measured
ing 9.8 g of sulphuric acid, H2SOu,, in under usual room conditions, when one
200 cm® of solution, (c) magnesium mole of gas molecules occupies 24 dm.)
sulphate, MgSOx,, containing 24g of (a) 6.62 g of lead(ID) nitrate, Pb(NOs)2,
MgSO, in 100 cm? of solution, (d) pot- were heated to constant weight. 4.46 g
assium permanganate, KMnOQ,, con- of lead) oxide, PbO, were left. The
taining 7.9¢ of KMnO. in 1000 cm? other products of the reaction were
of solution, (e) ammonia, NHs, contain- 1.84 g of dinitrogen tetraoxide, N2Ox,
ing 11.2dm* (measured at s.t.p.) of and 480 cm® of oxygen gas. (b) 13 g of
NH,(g) dissolved in 250 cm? of solution. zine reacted with 200cm? of molar
4. A compound containing only car- copper(II) sulphate, CuSO,. 12.8 g of
bon, hydrogen, and oxygen gave the copper were precipitated and 200 cm‘
Questions 287
of a molar solution of zinc sulphate, (orthophosphoric acid), H3PO,. The
ZnSOx,, were left. (c) 75cm’ of 2M reaction produced 16.4g of sodium
sodium hydroxide, NaOH, reacted with phosphate, NasPO..
50cm* of molar phosphoric(V) acid

Questions on Chapter 4
1. What do you understand by the atomic mass scale. Do you consider
term isotopy? Illustrate it by reference that it would be more suitable to adopt
to the element chlorine. Explain why, a scale based upon F = 19.00? Give
in spite of the existence of chlorine your reasons.
isotopes, the relative atomic mass of 4. Account for the existence of three
this element appears constant at 35.5 in isotopes of hydrogen. Why has it been
all ordinary samples. thought desirable to allot separate
2. Give an account of two different chemical names and symbols to them?
types of experimental evidence which Describe one method of preparation of
support the view that certain elements deuterium oxide (heavy water) and con-
exhibit isotopy. Explain this phenom- trast its physical properties with those of
enon in general terms. Is there any protium oxide (ordinary water).
observed connection between the atomic 5. Explain why separation of isotopes
number of an element and the occur- of the same element must, in general,
rence of isotopy in the element? Analy- depend on physical methods. Outline
sis showed that 2.743 g of ordinary three such methods. Why, in spite of
lead chloride, PbClz, required 2.128 g of isotopy, is the order of relative atomic
silver for complete precipitation as masses of elements almost the same as
silver chloride; corresponding figures the order of atomic numbers?
for two samples of the chloride from 6. Briefly explain the principles which
lead associated with radioactivity were underlie the use of nuclear energy for
(a) 1.613 g required 1.255g of silver, the production of electricity. Outline
(b) 3.748 g required 2.901 g of silver. the problems involved in the disposal of
Calculate the relative atomic mass of the waste products from nuclear power
lead from each source and explain the stations.
results. [Ag = 107.9, Cl = 35.46.] 7. Give an account of the usefulness
3. Discuss the isotopy of oxygen and of isotopes in industry and medicine.
consider its relevance to the relative

Questions on Chapter 6
1. Explain briefly the basis on which 2. Why is it now considered correct
Mendeléeff s Periodic Table was origin- that potassium (K = 39.1) should follow
ally drawn up about 1870. Quote the argon (Ar = 39.9) in the Periodic
table in its modern form up to the element Table? What factors explain the lower
calcium. Using the portion of the table relative atomic mass of potassium ? Men-
quoted, illustrate the meaning of period- tion one other similar case in the table.
icity of properties by reference to (a) 3. Relate the classification of elements
valency towards hydrogen or chlorine, in the Periodic Table, from He to Ca,
(b) valency towards oxygen, (c) one to the electronic arrangements shown
other property of the elements. in the atoms of these elements.
288 Physical Chemistry
4. Give the names and full electron p block, other than noble gases (rare
structures of two elements in each case gases), (c) the d block. (An example of
from the following sections of the a full electron structure is: Ar, 1s?2s?2p®
Periodic Table: (a) the s block, (b) the 3s23p®; see Chapter 5.)

Questions on Chapter 7
1. Explain what is meant by the magnesium chloride, phosphorus tri-
terms electrovalency, covalency, and chloride, phosphorus trichloride oxide,
dative covalency. Illustrate them by and ammonium chloride were repre-
reference to the compounds calcium sented by the structures:
chloride, phosphorus trichioride oxide,
POCIs, and trichloromethane, CHCls. Cl
Briefly relate the properties of these |
compounds to the kinds of valency they Cl—Mg—Cl, Cl—P—Cl,
exhibit.
2. Discuss the valency types shown Cl H, .Hy, CG)
in the compounds calcium oxide, | ‘ls
ammonium chloride, and tetrachloro- CI—P=O, and N
methane, CCl,. Explain why ammon- | LEI
ium chloride can precipitate silver Cl H H
chloride from silver nitrate solution but
tetrachloromethane cannot do this. Criticize these formulae in the light of
3. Discuss in electronic terms (a) modern ideas about valency and suggest
the formation of ions of the type, suitable modifications,
MO,”"~, illustrating by reference to 6. Discuss the desirability of applying
ClO,~— and PO,3~, (b) the existence of a the term molecule to particles present
stable compound, BCl3-NHs3, (c) the in crystalline sodium chloride, ammonia
valency types shown in the formation of gas, ammonium chloride crystals, and
ammonium sulphate. chlorine gas.
4. What is meant by the term atomic 7. Account for the shapes of the
number of an element? Explain why following species: (a) BCls, (b) NHs,
this characteristic of an element is (c) CHa, (d) SFe. Why does boron
closely related to the chemical properties trichloride form an addition compound
shown by the element. Illustrate by with ammonia?
reference to the elements potassium and 8. What is meant by the term de-
chlorine. localization? Give examples of materials
5. Atacertain period, the compounds in which it is thought to occur.

Questions on Chapter 8
1. Account for the properties of the Give an illustrative example in each
allotropes of carbon in terms of their case.
structures. 3. Give a brief account of the use of
2. Explain the following terms: (a) X-ray diffraction in determining crystal
hexagonal close packing, (b) body- structures. Describe as fully as you can
centred cubic structure, (c) co-ordina- the structure of sodium chloride.
tion number, (d) molecular lattice.
Questions 289
Questions on Chapter 9
1. ‘The hydrogen bond is essential 3. Consider, as widely as you can,
to life.’ Discuss. the consequences of water losing its
2. Explain the nature of the forces ability to hydrogen bond with itself and
which hold the following materials with other materials.
together in the solid state: (a) pure
silver, (b) sodium chloride, (c) diamond,
(d) ice, (e) iodine.

Questions on Chapter 10
1. What do you understand by the Compare ‘true’ and colloidal solutions
term colloid? Discuss the relation with respect to three different properties.
between ‘true’ solution, colloidal solu- 3. Give a brief account of (a) the
tion, and suspension. Give an account Brownian movement,(b) electrophoresis,
of two distinct methods of producing (c) protection, (d) coagulation in colloid
colloidal solutions with one illustrative chemistry.
example for each. Discuss the process 4. What do you understand by
of dialysis as a means of purifying lyophilic and lyophobic sols? Mention
hydrosols. two examples of each with water as
2. What do you understand by the dispersion medium. Tabulate four con-
terms dialysis and peptization as used in trasting properties of these two types of
aqueous colloid chemistry? Give one sol,
example of the use of each process.

Questions on Chapter 11
1. Describe how you would find the Laboratory temperature = 20 °C
vapour density of trichloromethane by Temperature of sealing of bulb
the method of V. Meyer. Illustrate the = 73 °C; pressure, 755 mmHg
calculation by the following results:
0.119 g of trichloromethane used; 25.4 [1 dm® of dry air and 1 dm® of dry
cm® of air displaced at 17°C and 745 hydrogen at s.t.p. weigh 1.293 g and
mmHg pressure; vapour pressure of 0.09 g respectively.] Discuss the ad-
water at 17°C is 15 mmHg. [One dm° vantages and disadvantages of Dumas’
of hydrogen at s.t.p. has a mass of method.
0.09 g.]
3. In a V. Meyer determination,
2. Describe a determination of the 0.156g of a volatile liquid displaced
vapour density of ether by the method of 47.0 cm? of air measured over water at
Dumas. Use the following figures to 14 °C and 762 mmHg pressure. [Vapour
illustrate the calculation: pressure of water at 14 °C is 13 mmHg.]
Calculate the vapour density and relative
Mass of bulb and air = S102 8 molecular mass of the liquid.
Mass of bulb and ether vapour after 4. An element X forms a number of
sealing = 52.032, 2 gaseous or volatile compounds. The
Mass of bulb full of water (+ broken following data are representative:
tip) = gs
290 Physical Chemistry
science, giving one example of each use.
% of X by Vapour
At a certain temperature and under
Compound mass in the density of identical conditions, the volumes of
compound compound
oxygen and nitrogen dioxide effusing
1 Sia 14 from the same apparatus in the same
2, 2d 2D time were in the proportion of 5:3.
3 50.0 32 Calculate the relative molecular mass of
4 Bes 15 nitrogen dioxide at this temperature.
5 36.3 22 [O = 16.0.]
6 60.0 40 7. State Boyle’s Law and Charles’
Law. What is the mathematical formu-
What is the probable relative atomic lation of Boyle’s Law? Why do real gases
mass of X? not obey Boyle’s Law exactly? Give
5. When vaporized at 130°C in a a qualitative account of van der Waals
V. Meyer apparatus, 0.100 g of meth- attempt to express the behaviour of
anoic acid displaced 39.2cm® of air gases more precisely.
measured over water at 12 °C and 750 8. What was formerly meant by a
mmHg. Calculate the relative molecular permanent gas? Methane, CH4, was
mass of the acid as indicated by these placed in this class; carbon dioxide was
figures. If the correct molecular formula not. Interpret these facts in the light of
of the acid is H,COz, how do you later, more correct ideas. A gas, X, has
account for the calculated value ob- a critical temperature of 430K and a
tained above? [Vapour pressure of critical pressure of about 78 atm; corre-
water = 11 mmHg at 12°C. One dm?® sponding figures for gas, Y, are 155K
of hydrogen at s.t.p. weighs 0.09 g. H = and about 50 atm. State, with reasons,
i CS 12, OS Ta! suitable methods of liquefaction for X
6. Outline briefly two different uses and Y. What do the terms critical
for the process of gaseous diffusion in temperature and critical pressure mean?

Questions on Chapter 12
1. Describe the preparation of a 3. (a) A solution of a non-electrolyte
good semi-permeable membrane. Ex- of relative molecular mass 342 has an
plain the meaning of the term semi- osmotic pressure of 0.630 atm at 12 °C.
permeable and describe how you would Calculate the concentration of the
use the membrane to demonstrate the solution in g dm~°.
existence of osmotic pressure in a given (b) Calculate the osmotic pressure in
solution of cane sugar. State the laws atm of a solution of urea (formula mass
of osmotic pressure and point out the 60) at 24°C, which contains 2.00 g of
analogy between the behaviour of gases urea in 100 cm$ of solution.
and of unionized solids in dilute solution. (c) Calculate the temperature of a
2. Describe a reasonably accurate laboratory in which a solution contain-
method of measuring the osmotic pres- ing 5.00 g of urea (formula mass 60) in
sure of a given sugar solution. A 160 cm* of solution gave an osmotic
solution containing 4.00 g of a certain pressure of 12.30atm. [G.M.V. is
sugar in 100 cm® of aqueous solution 22.4 dm? at s.t.p.]
has an osmotic pressure of 2.76 atm 4. Describe in detail how you would
at 16 °C. Calculate the relative molecular determine the formula mass of urea by
mass of the sugar. [G.M.V. is 22.4 dm® measurement of the elevation of the
at s.t.p.] boiling point of water. Illustrate the
Questions 291
calculation by the following data: compound forms a eutectic with both
0.300 g of urea in 30 g of water elevate A and B, the one with A having the
the boiling point of the solvent by lower melting point. Sketch (with
0.310 °C. [K, = 1.68 °C per 1000 ¢ of temperature axis vertical) a rough graph
water.] to express these facts. Consider what
5. Describe how you would determine happens during the cooling of various
the relative molecular mass of naph- liquids containing A and B, choosing
thalene by observations on the freezing enough cases to illustrate all the typical
point of benzene, in which naphthalene behaviour possible.
is readily soluble. [The freezing point 10. Explain the following terms,
of benzene is about 4 °C.] Illustrate the giving one example of each: (a) eutectic,
calculation from the data: 6.50¢ of (b) cryohydrate, (c) deliquescence, (d)
naphthalene in 200g of benzene de- efflorescence. Why does the addition of
pressed the freezing point of the solvent sufficient common salt to an icy path
by 1.250 °C. [K; = 5.10°C per 1000g produce a liquid on the path? What is
of benzene.] the relation of this fact to the use of
6. What do you understand by the ice-salt in freezing mixtures to give about
terms saturated solution and super- —15°C.?
saturation? Mention the usual condi- 11. State the Distribution (or Par-
tions under which a supersaturated tition) Law. Give an account of one set of
solution can be obtained. Illustrate by experiments by whichit can be illustrated.
reference to one example. How would 18 g of a compound X distribute them-
you prepare a saturated solution of selves between water and an equal
potassium nitrate at 5O °C and then use volume of an immiscible solvent Y so
it to determine the solubility of potassium that 2 g of X are in the water. Calculate
nitrate at that temperature? to the nearest integer the percentage of
7. Describe in outline two methods of X left in water if 1000 cm® of water
finding the solubilities of very sparingly containing 1g of X are extracted by
soluble salts, such as silver chloride or (a) one litre of Y, (b) half a litre of Y
lead sulphate. twice, (c) one-third of a litre of Y three
8. A salt, A, forms no compound times, so that one litre of Y is used in
with water but, at —25°C, forms a each case. Comment briefly on the ex-
eutectic of composition 60% water and perimental implications of these figures.
40% A. The solubility of A increases, 12. State the Partition Law as it
but not very rapidly, with rise of temp- relates to a compound in the same mole-
erature from —25 °C. Draw, with the cular state in both solvents. Discuss
temperature axis vertical, a rough graph briefly the variations produced in the
expressing these facts over the range statement of the law by (a) ionization,
of about —30°C to +30 °C. Describe (b) polymerization. The distribution of
what happens during the cooling of an acid between benzene and water
various liquids involving water and A, shows the following figures:
so as to show all the typical behaviour
possible. What is an alternative name Concentration Concentration
for the eutectic in this case? Why is it in water in benzene
not considered to be a compound in
spite of its constant melting point and 0.0450 0.504
composition at constant pressure? 0.0549 0.750
9. Two metals, A and B, B having the 0.0639 1.018
higher melting point, form a compound 0.0810 1.634
containing about 40% of A. This
292 Physical Chemistry
If the acid is monomeric in water, show steam, the liquid boiled at 99.3 °C and
that it is dimerized in benzene. of the distillate, 90 cm® were water anc
13. Describe in essential outline, 14 cm? were X. [Barometric height =
methods suitable for finding the solu- 760 mmHg; vapour pressure of water a
bility of (a) a gas slightly soluble in 99.3 °C = 740 mmHg. Density of X =
water, (b) a gas very soluble in water. 1.22 gcm~*.]
Assuming air to contain 79 % of nitrogen, 15. The distillation of a methanol-
21% of oxygen, and 0.04% of carbon water mixture can produce the pur
dioxide (by volume) and that all the alcohol, but distillation of a water
gases obey Henry’s Law, calculate the ethanol mixture gives a distillate with
composition of air boiled out from at best, about 96% of the alcohol
water. [The absorption coefficients are: Explain this difference.
nitrogen, 0.02, oxygen, 0.04, carbon 16. Hydrogen chloride forms a con
dioxide, 1.80.] stant-boiling mixture with water. Explair
14. Give an account of the principles the meaning of this statement, illustrate
involved in the process of steam distilla- it graphically and discuss the results o'
tion. To what type of compound can distilling liquids containing water anc
this process be suitably applied? Cal- hydrogen chloride. Choose enougt
culate the relative molecular mass of an examples to illustrate all possible types o:
organic compound, X, from the data: behaviour.
when the compound was distilled in

Questions on Chapter 13
1. Definethe terms: standard enthalpy formation of ethanol, C2HsOH, giver
of formation and standard enthalpy of that its enthalpy of combustion i:
combustion of a compound. Describe — 1421 kJ mol-!, and that carbon di.
in essential outline how the enthalpy of oxide and water have enthalpies of for.
combustion of a solid compound such mation of —405 and —284 kJ mol-}
as glucose, CgsHi20¢, can be determined. respectively.
Assuming this result, explain how you 5. Define the term enthalpy of neutra
could obtain the enthalpy of formation of lization. Calculate the enthalpy o:
glucose, given that water and carbon neutralization of caustic soda by nitric
dioxide are exothermic compounds with acid from the data: 250 cm? of 0.1 M
standard enthalpies of formation of a caustic soda solution were added tc
kJ mol~? and 6 kJ mol}, respectively. 250 cm’ of 0.1 M nitric acid in a laggec
2. Define the terms enthalpy of solution calorimeter. Both solutions were initi
of a compound and enthalpy of neutra- ally at 13.70 °C and the final temperature
lization. Describe in essential outline was 14.32°C. The water-equivalent o
how you would determine one of these, the calorimeter system was 50 g and the
using a named example. specific heat capacities of the solution:
3. State Hess’s Law of Heat Summa- can be taken as the same as that of water
tion. Show that it is essentially an 6. The combustion of carbon di
aspect of the general Law of Conserva- sulphide is exothermic and the enthalpy
tion of Energy. Devise a set of experi- of combustion of the compound i
ments, based on the conversion of 1108 kJ mol~?. Given that carbon di
barium oxide to a dilute solution of oxide and sulphur dioxide are exothermi
barium chloride, to illustrate Hess’s Law. compounds, with enthalpies of formatio1
4. Calculate the standard enthalpy of of 405 and 293 kJ mol~1, respectively
Questions 293
calculate the enthalpy of formation of 7. Discuss the measurements which
carbon disulphide. Comment on the must be made in order to determine the
stability of this compound at various lattice energy of sodium chloride.
temperatures, considering the resuit 8. What factors influence the solu-
obtained in the light of Le Chatelier’s bility in water at room temperature of an
Principle. ionic compound ?

Questions on Chapter 14
1. Discuss briefly what is meant by H2(g) + L(g) = 2HI(g)
the term chemical equilibrium. In illustra-
tion, consider the reaction: is about 61. If, in a similar experiment, 2
moles of hydrogen and 6 moles of iodine
CH;COOH() + C.H;OH() = are used, calculate the molar composition
CH;COOC-H;(1) + H.0() of the equilibrium mixture. [Use K =
occurring in the liquid phase. What is the 61.] Outline the experimental method
effect, in the light of Le Chatelier’s required to obtain such a set of observa-
Principle, of (a) increasing the relative tions.
concentration of the alcohol, (b) in- 4. At a certain high temperature, the
creasing the relative concentration of equilibrium constant of the reaction:
water? Why does the superficially N2(g) + O2(g) = 2NO(g),
similar reaction:
is 8 x 10-*. If air is a mixture of
HNO,(aq) -+- NaOH(aq) > nitrogen:oxygen in the proportions of
NaNOs,(aq) + H,0()), 4:1 by volume, calculate the percentage
proceed rapidly and almost completely of nitrogen oxide by volume in the gas
from left to right? produced when air reacts to equilibrium
2. If a mixture of 1 mole of ethanol at this temperature.
and 1 mole of glacial ethanoic acid is 5. The equilibrium constant of the
allowed to reach equilibrium at 25 °C, reaction:
two-thirds of the acid is esterified. CO.(g) + He(g) = CO(g) + H20(g)
Calculate the molar composition of the
mixtures obtained at equilibrium at this is 1.6 at 1000 °C. What is the percentage
temperature starting from (a) 1 mole of by volume of each constituent in the
the acid and 8 moles of the alcohol, (b) mixture at equilibrium (at 1000 °C)
1 mole each of ethanoic acid, alcohol produced from a mixture of (a) 75 cm®
and water. This reaction proceeds with of hydrogen and 75 cm® of carbon
very little heat change. What would you dioxide, (b) 50 cm® each of carbon
expect to be the effect of temperature dioxide and carbon monoxide and 100
change on the equilibrium? cm? of hydrogen ?
3. In an experiment of Bodenstein, 6. Ammonium hydrogen sulphide
20.57 moles of hydrogen and 5.22 moles being a solid, in what circumstances
of iodine were allowed to react to can the Equilibrium Law be applied to
equilibrium at about 405 °C. At this the equilibrium:
point, the mixture contained 10.22 moles NH,HS(g) = NH3(g) + H2S(g)?
of hydrogen iodide. Show from the
data that the equilibrium constant of the If the system is in equilibrium at ¢ °C.
reaction: with the total pressure at 400 mmHg of
294 Physical Chemistry
mercury, what is the partial pressure of (b) As for (a) in the reaction:
ammonia if hydrogen sulphide is added N.(g) + O2(g) = 2NO(g);
till its partial pressure is 500 mmHg with-
out change of volume at ¢ °C? Under BkJ absorbed.
what conditions would the partial pres- (c) Temperature and pressure in a
sure of hydrogen sulphide be 640 mmHg system of ice-water in equilibrium.
at t°C in the same volume?
7. State Le Chatelier’s Principle. (d) Temperature in a system of
Consider its application in predicting the potassium nitrate-water in equilibrium
results of varying: (the dissolution of the salt being
(a) Temperature, pressure, and rela- endothermic).
tive concentrations of reagents in the (e) Temperature and pressure in a
reactions: system of oxygen-water in equilibrium
N.(g) + 3H2(g) = 2NHa(g); (the dissolution of oxygen being
AkJ evolved. exothermic).

Questions on Chapter 15
1. What is meant by the ‘abnormal’ tion of the acid in 0.1 M solution at this
behaviour of electrolytes in aqueous temperature. Hencecalculate its approxi-
solution with respect to osmotic pressure mate degree of ionization in 0.05 M
and its related effects? Outline the solution.
ionic theory of Arrhenius as it was first 5. What is the osmotic pressure of a
put forward about 1880 and show how decimolar solution of ethanoic acid at
it explains the abnormalities. Mention 25 °C if its dissociation constant at this
one considerable modification of the temperature is 1.8 x 10~> mol dm~*?
theory in recent years. Answer in atmospheres. [G.M.V. is
2. Using the weak electrolyte, ethanoic 22.4 dm? at s.t.p.]
acid, in illustration, obtain a mathe- 6. Using the ‘insoluble’ salt, barium
matical expression which states the sulphate, in illustration, explain what
variation of its ionization with con- is meant by the term solubility product.
centration. What is the name given to Write expressions for the solubility
this expression? A molar aqueous product of lead iodide, silver chromate,
solution of a weak monobasic acid was and calcium phosphate. Assuming the
found to freeze at —1.91 °C. Calculate solubility product of barium sulphate to
the degree of ionization of the acid and be 1 x 107?° mol? dm~° at about room
its dissociation constant. [K; = 1.86 °C temperature, calculate (a) the solubility
per 1000 g of water.] of the salt in g dm~°, (b) its solubility in
3. Ammonium hydroxide is 1.4% 0.001 M sulphuric acid, assuming the
ionized at 25°C in 0.1 M solution. acid fully ionized. [BaSO, = 233.]
Calculate the dissociation constant of 7. The solubility product of barium
this base. Calculate the approximate carbonate is 7 x 10~°mol?dm-~® at
degree of ionization of the base in 16°C. What is the greatest mass of
0.01 M solution. What is the hydroxyl barium carbonate that can be dissolvec
ion concentration in this solution and, in 1500 cm of water at this temperature‘
hence, its pH? [BaCO3 = 197.]
4. If the dissociation constant of 8. The solubility product of bariur
formic acid is 2.14 x 10-4 mol dm-3 oxalate (ethanedioate), BaC.Ouz, at 18 °C
at 25°C, calculate the degree of ioniza- is 1.2 x 10~-7mol? dm~°®. Calculate th:
Questions 295
mass of barium oxalate that would be acids by caustic soda are: hydro-
precipitated from 1 dm® of its saturated chloric, —57.3 kJ mol; sulphuric,
solution at 18 °C by dissolving in it 0.67 g —114.6 kJ mol-; ethanoic, — 56.0 kJ
of anhydrous sodium oxalate. Assume mol-!. Comment on these figures.
both salts fully ionized. [Ba = 137, 14. Discuss, in ionic terms, the nature
C = 12,0 = 16, Na = 23.] and mode of operation of a typical
9. Explain, in ionic terms, the follow- acid—alkali indicator. Show by approxi-
ing facts: (a) silver chloride is almost mate graphs the change of pH during
insoluble in water but passes readily into titrations with strong and weak acids
solution when excess ammonia is added, and bases. Hence show what indicators
(b) iodine is readily soluble in a con- are suitable for the following titrations:
centrated solution of potassium iodide (a) NaOH-HCl, (b) KOH-H-2C.Ou, (c)
but only sparingly soluble in water, (c) NH.,OH-HCIl. Discuss the position
lead) chloride is appreciably more regarding choice of indicator for an
soluble in concentrated hydrochloric ethanoic acid—ammonia titration.
acid than in water, (d) the addition of 15. What do you understand by the
potassium cyanide solution to a solution term Aydrolysis as applied to inorganic
of silver nitrate gives a white precipitate salts? Discuss the situation prevailing in
which redissolves in excess of potassium aqueous solutions of: (a) aluminium
cyanide solution. chloride, (b) ammonium ethanoate, (c)
10. State, in ionic terms, what is sodium ethanoate. Explain (i) why a
meant by (a) a neutral aqueous liquid, dilute solution of sodium cyanide has
(b) the hydrogen ion index, pH, of an an almond smell, (ii) why a solution of
aqueous solution. (c) neutralization in ammonium chloride, if swallowed by an
aqueous solution. Calculate the pH of: individual, tends to turn his blood acidic.
(i) 0.01 M HCI, (ii) 0.001 M KOH, Briefly describe one method of estimating
(iii) an aqueous solution containing approximately the pH of any one of
7.3 g of HCl per dm%, (iv) an aqueous the above salt solutions.
solution containing 4g of NaOH per 16. What is meant by a buffer solution?
dm*°, (v) 0.01 M ethanoic acid in which State the purpose of such a solution
the degree of ionization is0.044. [Assume and, illustrating by one actual case,
the electrolytes fully ionized except explain why it is able to fulfil it. Mention
where it is stated otherwise. H = 1, appropriate buffer solutions for (a)
Cl = 35000 =—a1Gs Nal 2ayGe—s 12s moderately acidic conditions, (b) mod-
11. Outline the development in mean- erately alkaline conditions.
ing of the term acid. Explain the terms 17. Describe in outline a method for
strong and weak as applied to acids and determining the degree of ionization of
summarize the characteristic properties a weak electrolyte in aqueous solution.
of an acid. What is the relation be- (a) A 0.1 molar solution of sodium
tween an acid and one of its salts? chloride freezes at —0.335 °C. Calculate
12. What is meant by the basicity the osmotic pressure of the solution (in
of an acid? Illustrate by reference to mm of mercury) at 8°C, assuming
sulphuric acid and describe two methods that the apparent ionization of the
by which you could show this acid to salt is unaffected by the temperature
be dibasic. change. [K; = 1.86°C per 1000g of
13. Describe in outline two methods water; G.M.V. is 22.4 dm® at s.t.p.]
by which the strength of two acids can (b) 2.00g of potassium chloride
be compared. Briefly indicate the dissolved in 65g of water raise the
limitations of the methods. The enthal- boiling point of the liquid by 0.400 °C.
pies of neutralization of the following Calculate the apparent degree of ion-
296 Physical Chemistry
ization of the salt. [K, = 0.52 °C per —0.0935 °C. (b) The freezing point of a
1000 g of water.] solution containing 0.36 mole of calcium
18. Calculate the apparent degree of nitrate in 1000 g of water is —1.68 °C.
ionization of calcium nitrate in the 19. A 0.05 molar solution of magnes-
given solution from the data (at 760 ium sulphate acts as if the electrolyte
mmHg): (a) the freezing point of a - is 70% ionized. Calculate the osmotic
solution of 2.30 g of cane sugar, relative pressure of this solution at 20°C in
molecular mass 342, in 134 g of water is mmHg. [G.M.V. is 22.4 dm? at s.t.p.]

Questions on Chapter 16
1. State briefly in electronic terms stating the experimental conditions re-
what is understood by oxidation and quired, of the oxidizing actionof chlorine,
oxidizing agent. Illustrate by reference hydrogen ion, and potassium per-
to the action of chlorine on an iron(II) manganate, and of the reducing action of
salt. Show that the older ideas of hydrogen sulphide, tin(II) chloride, and
oxidation as ‘combination with oxygen’ hydriodic acid. In one oxidizing action
and ‘increase in the proportion of the and one reducing action of the above,
electronegative part of the molecule’ can interpret the change in electronic terms.
be reconciled with the electronic inter- 4. The following are all cases of
pretation. Illustrate by reference to oxidation, in an older meaning of this
suitable metallic compounds. term. Discuss the application of elec-
2. Rewrite the ‘molecular’ equations tronic ideas of oxidation to them.
given below in ionic terms; then, for
each molecular or ionic species on the
(a) 2Ca(s) + O.(g) > 2CaO(s)
left-hand side of each equation, state (b) C(s) + O2(g) + CO.(g)
whether it is oxidized or reduced or (c) 2NazSO3(aq) + O2(g) >
left unchanged, giving reasons for your
2Na.SO.(aq)
answer in electronic terms: (d) 2H2S(g) + O2(g) > 2H20() + 2S(s)
Discuss the oxidizing behaviour of (a)
(a) 2Na(s) + Cl.(g) ~ 2NaC\s)
sulphuric acid, (b) nitric acid on copper
(b) Zn(s) + H2SO.(dil. aq) >
and zinc. Wherever you can, interpret
ZnSO,(aq) + H2(g)
the reactions in electronic terms.
(c) 2Na2S203(aq) + Ia(s) > 5. Describe briefly how standard redox
NazS,O0¢(aq) + 2Nal(aq)
potentials may be measured. Explain
(d) SnCl,(aq) + HgClo(aq) >
how they may be used to predict (a)
Hg(l) + SnCl,(aq) the e.m.f. values of cells, (b) the like-
(e) 2FeSO,(aq) + Br.(l) + H.SO.(dil.
lihood of a redox reaction occurring.
aq) — Fe2(SO.)s(aq) + 2HBr(aq) Give two examples in each case. [Use
3. Give one example in each case, the data given in the table on page 250.]}

Questions on Chapter 17
1. Define the term catalyst. Outline catalyst promotion, (b) catalyst poison-
the two principal theories of catalysis ing, (c) autocatalysis, (d) negative cata-
and illustrate them by two examples in lysis (inhibition).
each case. Discuss, giving one example 2. Give three characteristics of a
in each case, what is meant by (a) catalyst. Describe experiments by which
Questions 297
you would try to show that manganese- What is meant by activation energy?
(IV) oxide is catalytic for the decomposi- How may the activation energy of a
tion of potassium chlorate by heat. reaction of your choice be found?
Mention a possible mechanism for this 5. Dilute hydrochloric acid reacts with
catalysis. Are there any experimental a solution of sodium thiosulphate to
facts supporting the mechanism you produce, initially, an opaque suspension
suggest? of sulphur. Describe a series of experi-
3. Distinguish between the terms ments to determine how the rate of this
order and molecularity of a reaction. reaction varies with (a) concentration
Describe, for a reaction of your choice, of the reactants, (b) temperature.
how the order of a reaction may be 6. Write an essay entitled ‘Catalysis
determined. in Industry’. Include economic con-
4. How does the rate constant of a siderations where appropriate.
reaction depend on the temperature?

Questions on Chapter 18
1. What are the products of electro- copper is deposited in the first cell,
lysis of the following in aqueous solution calculate (a) the mass of silver deposited
with inert electrodes: (a) sulphuric in the second, (b) the volume of oxygen
acid, (b) silver nitrate, (c) copper(II) liberated in the third at 15 °C and 740
sulphate? Give suitable equations. mmHg pressure. [0 = 16, Cu = 63.5,
What difference is made if (b) is per- Ag = 108. Molar volume of a gas at
formed with a silver anode and (a) with s.t.p. 1s 22.4 cm3.]
a copper anode? Give a sketch of 4. Discuss the principal factors which
apparatus suitable to demonstrate (a). determine the products obtained at
2. State Faraday’s Laws of Elec- the cathode and anode in the electrolysis
trolysis and describe, in outline, experi- of aqueous solutions containing ions
ments by which they can be illustrated. Suchieas. Na™, Cut. As" sOysae
(a) 0.396 g of copper is deposited from Cl-, and the ions derived from water.
copper(II) sulphate solution in 40 min Give suitable examples.
by a steady current of half an ampere. 5. Explain what is meant by (a) con-
Calculate the mass of copper deposited by ductivity, (b) molar conductivity, (c)
passage of 1 mole of electrons. [Assume molar conductivity of an electrolyte.
that one Faraday = 96000 C.] (b) An Explain why (c) can be determined
electric current is passed in series through experimentally for a strong electrolyte
dilute sulphuric acid and silver nitrate such as potassium chloride, but not for
solution for the same time. 112 cm® of a weak electrolyte such as ethanoic acid.
hydrogen are liberated in the acid Describe in outline the determination of
(measured wet at 16 °C and 750 mmHg.) the resistivity of 0.05 M sodium chloride
Calculate the mass of silver deposited at solution. If the value is 210 ohms, cal-
the cathode in the other cell. [Ag = culate the molar conductivity of this
108, H = 1.0, V.P. of water at 16°C electrolyte at this dilution.
is 14mmHg; 1 dm? of hydrogen at s.t.p. 6. What do you understand by
has a mass of 0.09 g.] Kohlrauch’s Law of Independent Mi-
3. Electric current is passed in series gration of Ions? The molar con-
for the same time through solutions ductivity of the following electrolytes
of copper(II) sulphate, silver nitrate, at infinite dilutation are: AgCl, 119.8;
and dilute sulphuric acid. If 0.315 g of AgNOQ3, 116.2; AgsSOu, 122.8; KCl,
298 Physical Chemistry
130.1; KNOs3, 126.5; K2SO,, 133.1. 7. Why cannot the molar con-
Show that these figures are in accordance ductivities at infinite dilution of
with the law. Describe, in outline, ethanoic acid be determined directly?
how any one of these figures can be Indicate briefly one method which can
determined. be used for this determination.

General Questions
The questions in the following pages are verting one form into another, and vice
included by permission of the Examining versa, and indicate the chief differences
Boards concerned: Joint Matriculation between them. (Lond. Inter.)
Board (J.M.B.); Oxford and Cambridge 4. Explain the following terms and
Schools Examination Board (0.C.); illustrate them with examples drawn
Examination Board for Wales (W.); from your experience in qualitative
London University Entrance and Schools analysis: (a) solubility product, (b)
Examination Council (L.); London common ion effect, (c) complex ion
University Intermediate B.Sc. (Lond. formation, (d) amphoteric behaviour.
Inter.); University of Durham (D.). (L.)
1. Measurement has shown that even 5. The atom of an element A contains
the most highly purified distilled water twelve protons. Give your reason fot
has a small electrical conductivity. How regarding the element as a metal ot
has this been explained? Calculate the non-metal and write the electronic
PH values in solutions formed by adding formulaof its compound with chlorine
48, 50, and 52cm® of 0.1M caustic The specific heat capacity of A is
soda separately to 50cm® of 0.1M 1.045 J g-1K~-+. What is its approxi:
hydrochloric acid. The ion product for mate relative atomic mass? Outline ¢
water may be taken as K, = 107-1* method for obtaining additional informa:
mol? dm~°, (D.) for the accurate determination of the
2. Show how the electronic theory of relative atomic mass. Comment briefly
valency provides an explanation of the on the adoption of 12C = 12.000 as the
regular variation of valency in passing standard for relative atomic mass. (W.
from group to group of the periodic 6. Explain fully what you understanc
classification of the elements. Tabulate by three of the following: (a) negative
as completely as the facts allow the catalyst, (b) the heat of formation of :
following information about the elements metallic oxide, (c) the vapour pressurt
Al, C, Ca, Cl, N, Na, Ne, S, rearranging of a liquid, (d) the valency of nitroge:
the elements in the order of their groups: in ammonium chloride. (0.C.
(a) the group in the Periodic Table, (b) 7. List the three main fundamenta
the formula of the characteristic hydride, particles which are constituents of atoms
(c) the action of the hydride on water, Give their relative charges and masses
(d) the formula of the oxide characteristic Explain concisely from the standpoin
of the groups to which the element of atomic structure: (a) the difference i
belongs, (e) the action of this oxide on chemical properties between a metal an
water. (J.M.B.) a non-metal, (b) the difference in valenc
3. Define the expression allotropic between sodium and magnesium, (<
modification and give three examples of isotopes, (d) oxidation and reductior
allotropy. In the case of one example Illustrate your answer by reference t
give an account of a method for con- suitable examples. (J.M.B
Questions 299
8. Trace the developments in the made the basis of the separation of
Classification of the elements which the metals when they are present as
culminated in the work of Lothar chlorides in an aqueous solution. Explain
Meyer and Mendeléeff. Indicate briefly the theory underlying the method.
how modern views of atomic structure (J.M.B.)
have provided a more satisfactory basis 13. Comment on, illustrate, or explain
of classification than that of Mendeléeff. the following statements: (a) hydrogen
(D.) chloride does not obey Henry’s Law,
9. Of what value is the knowledge of (b) metals can displace hydrogen from
the specific heat capacity of an element in sodium hydroxide, (c) a strong elec-
deciding its relative atomic mass? trolyte does not obey Ostwald’s Dilution
Illustrate your answer by reference to Law, (d) some allotropes differ in
the cases of (a) copper, (b) argon. chemical properties, others in physical
(Lond. Inter.) properties only. (0.C.)
10. Calculate the pH value of a
14. What are the distinctive charac-
solution of (a) 0.01 M sodium hydroxide,
teristics of the metallic and non-metallic
(b) 0.01 M ethanoic acid. The dissocia-
elements? Discuss their distribution in
tion constant of ethanoic acid is 1.8 x the periodic classification of the elements.
10-> mol dm-* at room temperature. By reference to any one group of the
Explain the observation that when Periodic Table, illustrate the gradation
equivalent amounts of (a) and (b) are of the above characteristics with in-
mixed the pH value of the resulting creasing relative atomic mass. (W.)
solution is greater than 7. What indi-
15. Explain the following statements:
cator would you use for titrating ethanoic
(a) it is incorrect to refer to concentrated
acid against sodium hydroxide? Give
sulphuric acid as strong sulphuric acid,
reasons for your choice of indicator.
(b) on dissolving in water hydrogen
(W.) chloride ionizes and dissociates but
11. What is meant by the electro-
potassium chloride only dissociates, (c)
chemical series? Arrange the elements
water is acid to phenolphthalein but
Ag, Fe, H, K in order of their positions
alkaline to methyl orange, (d) it is
in this series. Explain why (a) a steel
inadmissible to speak of the solubility
knife becomes coated with copper when
product of ethanoic acid. (J.M.B.)
dipped into a copper(I) sulphate solu-
tion, (b) galvanizing with zinc affords 16. Explain and illustrate four of the
more lasting protection to sheet iron following terms: (a) atomic number,
than tinning, (c) when sodium amalgam (b) isotope, (c) isobar, (d) complex ion,
is added to water the sodium alone (e) covalent bond. (L.)
reacts with the water, (d) copper will 17. What do you understand by a
dissolve in dilute nitric acid but not reversible reaction? Discuss how a
in dilute sulphuric acid. (J.M.B.) system in equilibrium is affected by
12. Define the terms velocity constant changes in the quantities of the various
and equilibrium constant. What is substances present, and illustrate your
meant by the common ion effect? Two answer by reference to (a) the reaction
metals D and E form chlorides DCl2 between arsenic(III) oxide and iodine,
and ECl, which are both soluble in (b) one other reaction. (Lond. Inter.)
water and sulphides DS and ES which are 18. How do the volumes of (a) ideal
both insoluble in water. The solubility gases, (b) real gases vary with temper-
product of DS is very much less than ature and pressure? Illustrate your
that of ES. Suggest a method by which the answer by referring (graphically or
insolubilities of the sulphides may be otherwise) to the real gases hydrogen,
300 Physical Chemistry
oxygen, and carbon dioxide, and indi- you draw from this about the state of
cate the two chief reasons for the failure sulphuric acid in water ? (ED
of real gases to obey the ideal, gas 24. Devise experiments to demon-
laws. What do the terms critical tem- strate that: (a) aqueous mercury(I}
perature and critical pressure signify? chloride is less ionized than aqueous
(J.M.B.) sodium chloride, (b) ethanoic acid is a
19. An element X forms a sulphate stronger acid than phenol, (c) iron(II)
which contains 64.06% of X and gives chloride exists as double molecules in
an alum with iron(II] sulphate. Find the vapour phase, (d) phosphorus
the relative atomic mass of X, explaining pentachloride dissociates on heating.
the arguments used in your calculation. (L.)
What other experiments would you 25. State Raoult’s law of vapour
suggest to confirm this relative atomic pressure lowering and explain qualita-
mass? (L.) tively why a solution of a non-volatile
20. Define osmosis and osmotic pres- solute in a volatile solvent has a higher
sure of a solution. Show how you can boiling point than the solvent. At 25 °C
demonstrate the existence of the former, a saturated aqueous solution of calcium
and how the latter can be measured for hydroxide, Ca(OH)z, possesses a vapour
an aqueous solution. (O.C.) pressure of 23.765 mmHg. At the
21. Explain, with illustrative ex- same temperature, the vapour pressure
amples, and distinguish between three of of water alone is 23.790 mmHg.
the following pairs of chemical terms: Assuming the calcium hydroxide to be
(a) allotrope and isotope, (b) iso- completely ionized in solution, what is
morphism and isomerism, (c) dialysis its solubility in moles per dm® at 25 °C?
and electrophoresis, (d) transition point ‘ (L.)
and eutectic point. (L.) 26. It is known that ammonia dis-
22. Suggest explanations for the tributes itself between equal volumes of
following: (a) The addition of a deci- trichloromethane and water in the
molar monobasic acid to an equal molar ratio of 1:26 respectively. In an
volume of a decimolar monoacidic experiment 25cm® of molar aqueous
base does not necessarily produce a ammonia were mixed with 25 cm® of
neutral solution. (b) Aqueous solutions 0.1 M aqueous copper(II) sulphate. The
of copper(II) sulphate and sodium mixture was shaken with 50cm*® of
ethanoate have opposite effects on trichloromethane for several minutes
litmus. (c) Determination of the relative and allowed to settle. The trichloro-
molecular mass of ethanoic acid in methane layer was separated off and
benzene solution by the cryoscopic repeatedly shaken with several portions
method gives a result which does not of distilled water. The aqueous extracts
conform to the formula C2H,O2. (L.) were combined and required 11.4 cm? of
23. What do you understand by the 0.05 M hydrochloric acid for neutra-
term colligative property? Describe a lization, using methyl red as indicator.
method for the determination of one (a) What information can you deduce
such property and show how this from the above results about any
method might be used for the determina- compound formed between copper(II)
tion of the degree of ionization of an sulphate and ammonia? Suggest what
inorganic acid. The depression of its structure would be. (b) Account,
freezing point observed for an aqueous as far as possible, for the greater solu-
1M solution of HCl is practically bility of ammonia in water than in
identical with that of a 1 M solution of trichloromethane. (L.)
H2S8O, in water. What inference can 27. State Avogadro’s Hypothesis.
Questions 301
20.0 cm® of a mixture of two gaseous 5% solution of glucose, CgHi20¢, in
hydrocarbons was mixed with 80.0 cm? water was found to give the same boiling
of oxygen. After explosion in a eudio- point elevation as a 3.3% aqueous
meter, the volume of the gas mixture solution of the simple carbohydrate
was found to have decreased to 60.0 cm. erythrose. If the composition of simple
Reaction with potassium hydroxide carbohydrates may be represented by
solution led to a reduction in volume of the general formula C,H2,O,, what is
30.0 cm®. After reaction with alkaline the molecular formula of erythrose?
pyrogallol solution, a further decrease [C = 12.0,H = 1.00,O = 16.0.] (L.)
in volume by 30.0 cm? was measured. 31. Discuss and explain the following
(a) Find the composition of the hydro- phenomena: (a) When an aqueous
carbon mixture. (b) Suggest a method solution of ethanoic acid is titrated
suitable for the separation of the gases with a standard solution of sodium
in the mixture. (L.) hydroxide, using methyl orange or
28. Give a concise account of the prin- methyl red as the acid/base indicator,
ciples underlying the following methods no satisfactory titration end-point can
of separation of chemical substances: be observed. (b) When aqueous am-
(a) steam distillation, (b) fractional monia solution is cautiously added to
distillation, (c) fractional crystallization, a solution of copper(II) sulphate, a
(d) diffusion of gases through a mem- blue precipitate results which is soluble
brane. (L.) in excess of the ammonia solution to give
29. A current of 0.2 amperes is passed a deep blue solution. (c) When dry
for 60 minutes through 100 cm® of an acid hydrogen chloride gas is dissolved in
solution of 0.5 M copper(I) chloride, the dry methylbenzene or dry benzene, the
cathode being platinum and the anode solution (i) fumes in air, (ii) is found to
being carbon. What mass of copper is be a non-conductor of electricity. When
deposited at the cathode and what this fuming solution is shaken with
would be the new concentration of the water, it is observed that (iii) the fuming
solution in moles of copper(I) chloride ceases, (iv) the aqueous layer conducts
per dm*? What ionic equilibria are electricity, (v) the aqueous layer turns
involved in this reaction? If a solution of blue litmus red. (L.)
0.5 ™M copper(II) chloride in 100 cm® 32. (a) Explain what is meant by the
was used with copper electrodes, how term ‘ionic product of water’ and
long would be required to deposit the discuss how the pH scale is related to it.
same mass of copper using the same (b) In an attempt to determine the ionic
current? What is the new concentration product of water by a conductivity
of the solution in moles of copper(I) procedure, Kohlrausch and Heydweiler,
chloride per dm® in this case? [Cu = in 1894, redistilled water 42 times under
63.5, 1 faraday = 96500 ampere reduced pressure and obtained water of
seconds.] (L.) conductivity 4.1 x 10-8’ ohm-! cm= at
30. State Raoult’s Law and explain 18 °C. Given that the ionic molar con-
concisely its application in the deter- ductivity at 18 °C of H*t and OH~ are
mination of relative molecular masses 316 ohm! cm? mol and 176 ohm!
by the boiling point method. Outline cm? mol-!, respectively, calculate the
the experimental procedure which you value of the ionic product of water at
would adopt for the determination of this temperature. Briefly explain why it
the relative molecular mass of a non- was necessary for Kohlrausch and
electrolyte which is soluble in water. Heydweiler to purify their water in so
Draw a sketch of the apparatus that you elaborate a manner. (G2)
would employ. In an experiment, a 33. Describe the phenomenon of
302 Physical Chemistry
osmosis, illustrating your answer with Number of Logarithm of
one appropriate example. Define the electrons removed ionization energy
term ‘osmotic pressure’ and describe, in
outline only, a method for the deter- 2.69
mination of the osmotic pressure of a 3.66
solution of sugar. Calculate the osmotic 3.84
pressure at 25 °C of an aqueous solution 397,
of cane sugar, Cy2H22011, containing 4.12
3.42¢dm-*. (C= 12,H = 1,0 = 16; 4.22
G.M.V. at s.t.p.. = 22.4 dm*.] (L.) 4.30
34. A student has measured the pH 4.41
values of a number of dilute aqueous 4.46
salt solutions and reports the follow- 5245
ing results:
pe AONDAMNRWN5.30
—EH
RP
OU

Solute PH value
(a) Plot a graph of the logarithm of
(i) Ammonium sulphate 4.5 ionization energies against the number
(ii) Sodium cyanide 11 of electrons removed. (b) Explain how
(iii) Sodium hydrogen- from your graph information about the
carbonate ed electronic structure of a sodium atom
(iv) Iron(I) chloride 4 may be obtained. (c) Discuss how a
knowledge of successive ionization
energies of the elements can help the
chemist in explaining the chemical
(a) Explain the experimental results properties of elements. (L.)
obtained by the student. (b) The pH 37. The solubility product of mag-
values given above were obtained by nesium hydroxide, Mg(OH).z, is 1.25 x
means of indicator paper. Give a 10-7° mol? dm-°%. (a) Calculate the
brief account of the theory of acid/base solubility (in gdm~*) of magnesium
indicators. (L.) hydroxide in water. (b) Suppose you are
35. X-ray diffraction is one of the given 1dm*® of a 0.01M aqueous
most powerful techniques for the elucida- ammonia solution. What is the maxi-
tion of the structure of crystalline solids. mum quantity (in moles) of magnesium
Give a brief outline of the principles on hydroxide that can be dissolved in it?
which this technique is based. _Crystal- [Kaissoctation =1389<810=2eiimol Sidi?
line solids may be classified, according for NHa(aq)]. (c) How does the solu-
to the nature of their structural units, as bility product of Mg(OH)z compare
ionic, molecular or atomic crystals. For with the solubility products of the
each of these classes give an appropriate other alkaline earth metal hydroxides ?
example, state how the particles of the [Mg = 243,H=1,0=16] (L,)
substances chosen are held together in 38. ‘Covalent substances tend to be
the solid state, and discuss briefly the volatile, of low melting point, soluble
relationship between the structure and in non-aqueous solvents but not in
properties of these materials. (L.) water, and non-electrolytes; whereas
36. The following table lists the electrovalent substances tend to have the
logarithmic values of the successive opposite properties.’ Discuss this state-
ionization energies (in kJ mol~+) of the ment critically, referring particularly to
element sodium. silicon dioxide, SiOz, hydrogen chloride,
Questions 303
HCl, sodium chloride, NaCl, trichloro- perimental reaction rates. Give a concise
methane, CHCls, and ammonium chlo- account of each of these theories, and
ride, NH,Cl. (L.) discuss the major similarities and/or
39. The kinetic theory leads to the differences that exist between them. (L.)
equation 41. It is customary to classify mixtures
pV = 4nmC?, for an ideal gas. of two miscible liquids according to
whether or not they obey Raoult’s Law.
(a) What do the letters p, V, n, m, and é (a) State Raoult’s Law and show
in this equation stand for? Give one diagrammatically how (i) the vapour
consistent set of units in which they may pressure and (ii) the boiling point, of an
be expressed. (b) Draw a rough sketch ideal binary mixture vary with compo-
to show how the speeds of the molecules sition. (b) Give one example each of
of a gas are distributed at some temp- binary mixtures showing positive and
erature, 7,. On the same sketch, show negative deviations from Raoult’s Law
the distribution of the speeds at some and give approximate vapour pressure/
higher temperature, T2. Discuss briefly composition diagrams for these. Discuss
the bearing this shift in distribution has the reason(s) for these deviations in
on the rates of reactions in gases. (c) terms of molecular interactions. (L.)
State Avogadro’s Law and show that it 42. What properties must a substance
necessarily follows for an ideal gas from possess if it is to be suitable for use as
the equation above. (d) Real gases an acid-base indicator? A _ certain
deviate to a greater or less extent from indicator has an acid dissociation
‘ideal’ behaviour. Discuss qualitatively constant of 10-5 mol dm~3; ina solution
those assumptions made in deducing the of pH 1, the indicator is blue, whereas
ideal gas equation which are not true for when the pH is 13 the colour is yellow.
real gases and the effect they have on the Explain carefully (a) how this indicator
behaviour of the latter. Under what could be used to find the pH of a sample
conditions does the behaviour of a gas of slightly cloudy pond-water, and (b)
like carbon dioxide tend to that of an why the indicator would be suitable for
ideal gas? (L.) titrating hydrochloric acid with either
40. The collision theory and the ammonia solution or sodium hydroxide
transition state theory represent two solution but not for titrating ethanoic
different scientific models which the acid with sodium hydroxide solution or
chemist uses in order to interpret ex- with ammonia solution. (L.)

Scholarship Level
1. ‘Oxidation and reduction are elec- associate with (a) electrovalent linkage,
tron transfer reactions.’ Discuss this (b) covalent linkage? Comment on the
statement. Name two compounds which types of chemical bond present in (c)
can function as both oxidizing and HCl, (d) KCl, (e) NH.Cl, (f) CO, (g)
reducing agents. Give an illustrative HNOs. (Z.)
reaction in each case. (W.) 3. How would you remove the chem-
2. Give an account of variable valency ically active constituents of air in order
as exhibited by the elements of the to show that part of it is chemically
first two periods of the periodic system inert? What part did the discovery of
and discuss the explanation of the the noble gases play in the development
phenomena given by the modern elec- of modern chemical ideas? (0.C.)
tronic theory. What properties do you 4. Describe and explain what takes
304 Physical Chemistry
place when an electric current passes electropositive character is related to
between platinum electrodes immersed the occurrence and mode of extraction
in: (a) an aqueous solution of copper(II) of an element and to the properties of
sulphate, (b) a hot aqueous solution of its oxide and chloride. What connection
potassium chloride, (c) fused sodium has the electropositive character with
ethanoate, (d) a concentrated aqueous the position of an element in the Periodic
solution of potassium hydrogensulphate Table? (L.)
at 0°C, (e) a solution of silver cyanide 10. Explain, briefly, thetheory govern-
in potassium cyanide solution. (J.M.B.) ing each of the following processes, and
5. Discuss the following, with illus- in each case give a practical illustration:
trative examples where necessary: (a) (a) ether extraction, (b) distillation under
the determination of relative atomic reduced pressure, (c) fractional re-
mass by Cannizzaro’s method, (b) the crystallization to constant melting point,
laws of conservation of mass and of (d) drying in a desiccator. (L.)
energy, (c) the rule of Dulong and Petit, 11. The equilibrium constant, K,, for
(d) the reducing power of a metal in the thermal dissociation of ammonium
relation to its position in the electro- chloride according to
chemical series. (L.)
6. Discuss the application of the NH.Cl(s) = NH,(g) + HCl(g)
electronic theory of valency to account is 0.36 atm?, at 327 °C.
for the properties of the following: (a) Outline an experimental rogaine
methane, sodium chloride, ammonia, which you would adopt for the deter-
phosphorus pentachloride. It is often mination of the equilibrium constant
found difficult to assign to a molecule for this reaction. (b) State the expression
or an ion a single electronic formula for the equilibrium constant, Kp, of the
which adequately represents its proper- above reaction. Suppose that exactly
ties. Cite such an example and comment 0.2 moles of NH.Cl(s) were introduced
very briefly on the present interpretation into an evacuated container of 10 dm?
of its structure. (W.) capacity and then heated to 327°C.
7. Give a brief account of the present What is the quantity (in moles) of
theory of atomic structure and show NH.,Cl that would remain as a solid?
how it accounts for (a) the periodic (c) Consider a closed reaction vessel
system of the elements, (b) the existence containing an equilibrated mixture of
of isotopes, (c) the principal types of ammonia, hydrogen chloride, and solid
valence bonds occurring in chemical ammonium chloride. Explain whether
compounds. (J.M.B.) the addition of more NH.Cl(s) to this
8. Give a definition of catalysis by system would increase, decrease or
describing three characteristics of a leave unchanged the partial pressure of
catalysed reaction. Show, by describing ammonia. (L.)
three experiments each illustrating a 12. The ‘strength’ of acids is custom-
different catalysed reaction, how you arily expressed in pK-units, where pK
could demonstrate the characteristics denotes the negative logarithm to the
you have mentioned. Describe briefly base 10 of the dissociation constant of
two suggestions as to the mechanism of an acid. For an investigation of the
catalysis. Give one reaction to illustrate strength of hydrofluoric acid in aqueous
each type of mechanism you mention. solution, the following procedure was
(J.M.B.) adopted: 25.00 cm® of a dilute hydro-
9. What do you understand by the fluoric acid solution was added to 50.00
electropositive character of an element? cm? of a 1.000M sodium hydroxide
Show by appropriate examples how the solution contained in a laboratory flask.
Questions 305
The excess sodium hydroxide was then (experimental) thermochemical data, (b)
back-titrated with 1.000M_ sulphuric information about the solid structure of
acid solution, requiring 6.25 cm® of the an ionic compound, especially the pack-
latter. In a separate experiment the ing arrangement of the ions and inter-
PH of the hydrofluoric acid solution was ionic distances. Discuss why method
found to be 1.66. (a) From the above (b) appears to be unsatisfactory when
information, calculate the dissociation applied to substances such as silver
constant and the pK value of hydro- chloride and zinc sulphide. (L.)
fluoric acid. (b) Briefly comment upon 16. The Arrhenius equation, k =
the procedure adopted for the titration. Eq

(c) Suggest a method that might be used Ae? describes the variation of the
for the determination of the pH of the rate constant, k, of a chemical reaction
above solution. (d) How does the with temperature. In this equation, e is
strength of hydrofluoric acid compare the base to the natural logarithm, R the
with that of the other hydrogen halide gas constant, and A and E, are constants
acids? Suggest reasons for any differ- for a particular reaction. (a) Explain
ences you may infer. (L.) concisely the physical significance of the
13. Discuss and explain the following two constants A and E,. (b) The
observations: (a) The boiling point of decomposition of N.O; was studied over
methane is considerably lower than that arange of different temperatures and the
of the corresponding silicon hydride following data obtained:
(SiH,, monosilane), whereas the boiling
points of ammonia and of water are Temp-
1/T logk
higher than those of phosphine and of ee Tlie Gerayned |<dein's=3)
hydrogen sulphide respectively. (b)
Aniline is a weaker base than ammonia,
but ethylamine is a stronger base than 298 | 3.36 x 10-2 | —4.76
ammonia. (c) 1 M aqueous solutions of 308 | 3.25% 10-2 | —4.18
hydrogen chloride, hydrogen bromide, 318 |3.14x 10-2 | —3.60
and hydrogen iodide have pH values of 328 | 3.05 x 10-2 | —3.12
0.09, 0.06, and 0.02 respectively, whereas 338 «| 2.96 x 10-2 | —2.62
the pH of a 1M aqueous solution of
hydrogen fluoride is 1.7. (L.) Plot a graph of log k against 1/T and
14. Outline the essential assumptions from it obtain a value for E,, using the
made about the nature of a gas in the modified Arrhenius equation
kinetic theory of gases and discuss
Ey
critically the extent to which these log k = log A = 23RT’
assumptions are justified. Show how
the kinetic theory of gases accounts for where R = 8.3 J K~1 moil-?. Also cal-
(a) Boyle’s Law, (b) Avogadro’s Law, culate a value for A. (L.)
and (c) Dalton’s Law of Partial Pressures. 17. Discuss briefly the procedure by
Under what conditions would you which three of the following could be
expect these laws to apply strictly to real determined experimentally: (a) The
gases? How have any of these laws been number of molecules in a mole. (b)
modified to explain the behaviour of The first ionization potential of an
real gases? (L.) element. (c) The atomic number of a
15. What is meant by the term given element. (d) The standard redox
‘lattice energy’? Outline the theoretical potential for the system Fe?* /Fe?*. (L.)
principles underlying the methods of 18. ‘Chemical bonding, irrespective
evaluating lattice energies from (a) of whether it is ionic or covalent, is
306 Physical Chemistry
basically the result of electrostatic 21. Atroom temperature the electrode
interactions between electrons and nuclei. potential for the system Fe**(aq),
The essential difference between the Fe? +(aq)|Pt is given by the equation
two types of bonding lies in the distribu-
tion of electrons.” Explain and discuss je = E? “pb 0.06 logio
[Fe?*] (1)
this statement. (L.) [Fe?*]
19. The direction of change of a
For this system, E® = 0.77 V. Similarly
chemical system at constant temperature
for the silver electrode, Ag *(aq)|Ag(s)
and pressure may be said to be governed
by two factors: the enthalpy change and E= E® + 0.06 logio[Ag*] (2)
the entropy change for the given
reaction. (a) Explain what you under- and E® =0.80V. (a) Calculate the
stand by the term ‘entropy’. (b) Discuss, electrode potential of the half-cell
with suitable examples of your own Fe?*(aq), Fe**+(aq)|Pt when the
choice, the way in which the enthalpy and iron(II) ion concentration is 1.0 M and
entropy of a reaction determine the the iron(I]) ion concentration is 0.6 M
direction of the change. (GLa) (b) What concentration of silver ions
20. The atomic spectrum of hydrogen (Agt) in contact with metallic silver
consists of a number of lines arranged would give the same electrode potential
in series (called e.g., ‘Balmer’ series, as you have calculated in (a)? (c) What
‘Paschen’ series, etc.). Describe how, in would happen if a solution 1.0 M with
principle, these spectra could be ob- respect to Fe®*, 0.6 M with respect to
served and recorded. What information Fe?+, and containing silver ions of the
about the hydrogen atom can be concentration you have found in (b),
obtained from these line spectra? was in contact with; metallic silver?
Discuss briefly (a) the relation between Explain. (d) Use equations (1) and (2)
the atomic spectrum of an element and to derive a value for the equilibrium
its ionization energy; (b) the information constant K for the reaction
about the chemical properties of an Fe** (aq) + Ag(s) = Fe?*(aq) +
element that can be obtained from its Ag*(aq).
ionization energies. (L.) (L.)

Answers to Numerical Questions (pages 285-306)


Chapter 1 (c)0.1, formula units; (d) 100,
formula units; (e) 0.05, formula
3s 452=1
units
5. 560g; 56.0; 2
3. (a) 0.2; (b)1.0; (c)2.0; (d) 0.05;
6. Valency 4; relative atomic mass
(e) 2.0
91.2
4. CH,O; C.H.02; CH;COOH
8. Relative atomic mass 186; valen-
5. BaCl..2H2O
cies 4 and 7
9. 48.0; 31.0% Ti
12. 43.9 Chapter4
2. 207.4; 206.2; 207.9
Chapter 2
Chapter 11
1. (a) 80g; (b) 142; (c) 5.85g;
(d) 49.92; ()115¢g 1. 58.9
2. (a) 0.05, molecules; (b) 0.2, atoms; 2. 36.9
Questions 307
3. 39.3 3s" 199° 107° mofdm-*: *4.4°;
4. 16 pH 10.6
5. 60.8; 32% dimerized 4. 0.045; 0.063
6. 89 5. 2.48 atm
6. (a) 2.3 x 10-5gdm-® (b) 23 x
(Om Ge-dinime
Chapter 12 7. 0.025 g
2. 344 8. 0.077 g
3. -(a) 9.22gdm-3; (b) 8.1 atm; 10. @) 2; (i) 11; (iii) 0.7; (iv) 13;

4.
(c)
60
14.8 °C 17.
et
(a) 3154mmHg; (b) 0.86
5. 133 18. 0.75

13. Nz, 63.5; Oz, 33.7; CO2,2.9% 19 1552 mmHg


14. 127

a is Chapter 18
Beer
apt
2. (a)31.7g; (b) 0.996 g
4. —242kJ mol-? 3. (a) 1.07 g; (b) 60.2 cm®
5. —56.8 kJ mol-* 5. 95.2 ohm7! cm? mol-!
6. +117kJ mol-?

General Questions
Chapter 14
12.692, 7.0; 11,39
2. (a) ester and water 0.97, acid 0.03,
10. (a)12; (b)3.4
alcohol 7.03 mole; (b) ester 0.54,
19. 85.6 (Rb)
water 1.54, acid and alcohol 0.46
25. 1.94 x 10-2? mo! dm-*
mole 32. K,, = 0.695 <x 10->1* mol? dm=-°
3. HI 3.88, H. 0.06, I, 4.06 mole 37. (a) 0.0184 g dm-®
Ag tec,
5. (a) CO and H.O, 28%; COz and
Hz, 22%; (b) COs, 10.53. Ha,
39:5:. CO, 39.5; H,0;d457, Scholarship Questions
6 NHs; 80mm; 62.5mm 11. (b) 0.078 mole
12. (a) Kaiss = 3.24 x 10-* mol
dm~*; pK = 3.49
Chapter 15 21. (a) 0.783 V;_ (b) 0.521 mol dm~°;
2. 0.027; 7.5 x 10-*
mol dm-* (d) 3.17 mol dm~$
308 Physical Chemistry
TABLE OF RELATIVE ATOMIC MASSES TO FOUR SIGNIFICANT FIGURES
(Scaled to the relative atomic mass 12C =12 exactly)

Values quoted in the table, unless marked * or f, are reliable to at least +1 in the fourth signifi-
cant figure. A number in parentheses denotes the atomic mass number of the isotopes of longest
known half-life.
Relative Relative
At. Name Sym- Atomic At. Name Sym- Atomic
No. bol Mass No. bol Mass
1 Hydrogen H 1.008 53 Iodine I 126.9
2 Helium He 4.003 54 Xenon Xe 131.3
3 Lithium Li 6.941*F 55 Caesium Cs 132.9
4 Beryllium Be 9.012 6 Barium Ba 13763.
5 Boron B 10.817 57 Lanthanum La 138.9
6 Carbon Cc 12.01 58 Cerium Ce 140.1
7 Nitrogen N 14.01 59 Praseodymium Pr 140.9
8 Oxygen O 16.00 60 Neodymium Nd 144.2
9 Fluorine F 19.00 61 Promethium Pm (145)
10 Neon Ne 20.18 62 Samarium Sm 150.4
11 Sodium Na 22.99 63 Europium Eu 152.0
12 Magnesium Mg 24.31 64 Gadolinium Gd 157.3
13. Aluminium Al 26.98 65 Terbium Tb 158.9
14 Silicon Si 28.09 66 Dysprosium Dy 162.5
15 Phosphorus P 30.97 67 Holmium Ho 164.9
16 Sulfur S 32.06¢ 68 Erbium Er 167.3
17 Chlorine Cl 35.45 69 Thulium Tm 168.9
18 Argon Ar 39.95 70 Ytterbium Yb 173.0
19 Potassium K 39.10 71 Lutetium Lu 175.0
20 Calcium Ca 40.08¢ 72 Hafnium Hf 178.5
21 Scandium Sc 44.96 73 Tantalum Ta 180.9
22 Titanium Ti 47.90* 74 Wolfram (Tungsten) W 183.9
23 Vanadium Vv 50.94 75 Rhenium Re 186.2
24 Chromium Cr 52.00 76 Osmium Os 190.2
25 Manganese Mn 54.94 77 ~=Iridium cp lt 192.2
26 Iron Fe 55.85 78 Platinum PE 195.1
27 Cobalt Co $8.93 79 Gold Au 197.0
28 Nickel Ni 58.70 80 Mercury Hg 200.6
29 Copper Cu 63.55 81 Thallium Tl 204.4
30 Zinc Zn 65.38 82 Lead Pb 207.2
31 Gallium Ga 69.72 83 Bismuth Bi 209.0
32 Germanium Ge a9 ™ 84 Polonium Po (209)
33 Arsenic As 74.92 85 Astatine At (210)
34 Selenium Se 78.96* 86 Radon Rn (222)
35 Bromine Br 79.90 87 Francium Fr (223)
36 Krypton Kr 83.80 88 Radium Ra (226)
37 Rubidium Rb 85.47 89 Actinium Ac (227)
38 Strontium Sr 87.62t 90 Thorium Th 232.0
39 Yttrium Y 88.91 91 Protactinium Pa (231)
40 Zirconium Zr 91.22 92 Uranium U 238.0F
41 Niobium Nb 92.91 93 Neptunium Np 237)
42 Molybdenum Mo 95.94* 94 Plutonium Pu 244
43 Technetium Tc (97) 95 Americium Am 243
44 Ruthenium Ru 101.1 6 Curium Cm 24
45 Rhodium Rh 102.9 97 Berkelium Bk 24
46 Palladium Pd 106.4 98 Californium Cf 251
47 Silver Ag 107.9 99 Einsteinium Es (3a
48 Cadmium Cadi 1124 100 Fermium -Fm (25

| Scere od) ee ie
49 Indium In 114.8 101 Mendelevium Md = (258
Im: : wrenci
52 Tellurium Te 127.6 be el ce

* Values so marked are reliable to + 3 in the fourth significant figure.


+ Values so marked may differ from the relative atomic masses of th e relevant elements in some naturally
occurring samples because of a variation in the relative abundance of the isotopes.
Index

ABSOLUTE TEMPERATURE, 114 Base, 215, 219


Acid, 215, 219 Beckmann thermometer, 138, 141
characteristic properties, 215 Benzene, structure, 82
comparison of strength, 220 Berthollide compounds, 4
salt, 219 Berzelius, 8
strong, 217 Beta rays, 38
weak, 217 Binding energy, 36
Activation energy, 261 Body-centred cubic, 89
Adsorption, catalyst, 265 Bohr, 55
coefficient, 164 Boiling point constant, 135
indicator, 112 Bomb calorimeter, 181
Air, liquefaction, 123 Bond, angles, 79
Alpha particles, 23, 25 energies, 185
Aluminium chloride, hydrolysis, 226 polarity, 97
Ammonia, bonding in, 71 Born-Haber cycle, 189
manufacture, 201, 206 Boron trichloride, ammonia com-
shape, 79 pound, 74
Ammonium chloride, bonding in, 75 Boron trifluoride, shape, 80
ion shape, 80 Boyle’s Law, 7, 113
Andrews, 120 Bragg equation, 65
Arrhenius, 261 Bredig’s arc, 106
Aston, 30 Breeder reactor, 43
Atom, 5 Bronsted and Lowry theory, 217
Atomic, bomb, 41 Brownian movement, 109
emmission spectra, 46 Buffer solution, 233
mass, 7, 10, 15
number, 27 CAESIUM CHLORIDE, structure, 92
orbitals, 55, 80 Cannizzaro, 10
theory of Dalton, 4 Carbon-14 dating, 45
volume, 61 Carbon dioxide, bonding in, 72
Atomicity, 10 Carboxylic acids, 84, 101
Aufbau Principle, 52 Catalysis, 262
Autocatalysis, 268 negative, 267
Avogadro, constant, 17 theories of, 265
hypothesis, 8 Cathode rays, 21
Azeotropic mixture, 170 Cell constant, 282
Azimuthal quantum number, 50 Cells, e.m.f. of, 248
Chadwick, 23
BALMER SERIES, 47 Classification of elements, 58
309
310 Physical Chemistry
Cloud chamber, 25 Electrochemical equivalent, 277
Colloids, 105 Electrode potential, 248
preparation of, 106 Electrolysis, 269
properties of, 108 laws of, 276
Combining masses, 6 Electrolytes, 269
Complex ions, 76, 241 Electron, 21
Compounds, polar, 97 affinity, 189
Conductance cell, 282 density, 65
Conductivity, 288 distribution in atoms, 49, 62
Conjugates, 217 measurement of e/m, 22
Convergence frequency, 48 pairing, 77
Co-ordinate covalency, 73 spin, 53
Co-ordination number, 88 wave nature, 22
Copper(II) hydroxide, soluble in Electrophoresis, 110
ammonia, 242 Electrovalency, 67
Covalency, 70 Element, 1
Covalent, lattice, 94 Elevation of boiling point, 135
molecules, shape of, 78 Emission spectra, 46
radii, 103 Endothermic reactions, 179, 193
Critical pressure and temperature, 122 Energetics, 178
Cryohydrate, 158 Energy, cycles, 183
Crystal structures, 86 levels, 49
Cubic close-packing, 88 Enthalpy of, atomization, 185
combustion, 181
DALTon’s, Atomic Theory, 4 formation, 181 t
Law of Partial Pressures, 114 hydration, 191
Dative covalency, 73 neutralization, 222
Degree, of dissociation, 150, 208 reaction, 180
of ionization, 150, 208 solution, 191
Deliquescence, 162 Entropy, 193
Delocalization, 82 Equation, determination of, 19
energy, 188 Equilibrium, chemical, 194
Depression of freezing point, 137 Law, 194, 207
Deuterium, 35 Equivalent masses, 6
Dialysis, 107 Ethane, bonding in, 72
Diamond structure, 94 Ethereal extraction, 174
Diffraction, X-ray, 64 Eutectic, 158
Diffusion, 32, 115 with compound, 159
Dilution Law (Ostwald), 208 Exothermic reactions, 179
Dipole, 98
Dispersion medium, 105
Dissociation constant, 208 FACE-CENTRED CUBIC, 88
Distillation of mixtures, 167 Faraday’s Laws, 276
Distribution Law, 173 First-order reaction, 257
Dobereiner’s Triads, 58 Fission, 41
Dumas, 129 Fluorite, structure, 94
Formula, empirical, 19
EFFLORESCENCE, 162 molecular, 19
Einstein, 3 Fractional distillation, 169
Index 311
Free energy, 193, 207, 252 Ice, structure of, 100
Freezing equilibrium, 197 Indicators, 229
Freezing-point constant, 138 choice of, 231
Fundamental particles, 21 Inhibition, 263
Fusion, 42 Intermediate compound catalysis, 265
Intermolecular bonding, 97
GAMMA-RADIATION, 38 Interstitial sites, 90
Gas, constant, 114 Iodine, structure of, 96
Laws, 113 soluble in KI, 242
Ionic, bonding, 67
deviations from, 119
Gay-Lussac’s Law, 7 crystals, 89
lattices, 89
Geiger, 24
Gels, 105
product of water, 212
radii, 91
Giant lattices, 86
Ionization, degree of, 150, 208
Gibb’s free energy, 193
energy, 48
Graham, 32, 115
Ions, size of, 91
Graphite, structure, 95
Isoelectric point, 111
Gypsum, 101
Tsomorphism, 14
Isotopes, 29
HABER PROCESS, 201, 206 artificial, 43
Half-cell, 248 separation, 36
Half-life, 40 uses, 44
method, 258 Isotopy, 29
Heavy water, 35 of chlorine, 29
Heisenberg, 56 of hydrogen, 34
Hess’s Law, 183 of lead, 33
Heterogeneous, catalysis, 264 of neon, 32
equilibria, 201
Hexagonal close-packing, 87 JOULE—THOMSON EFFECT, 123
Homogeneous, catalysis, 264 for helium, 124
equilibria, 194 for hydrogen, 124
Hybridization, 81
Hydrates of, KINETIC THEORY, 115
copper(II) sulphate, 162 Kohlrausch’s Law, 281
iron(II) chloride, 161
Hydrogen, bomb, 42 LANDOLT, 2
bond, 98 Landsberger, 135
diatomicity, 9 Law of, Boyle, 7, 113
electrode, 248 Charles, 7, 113
ion index, 213 Conservation of Mass, 2, 5
isotopy, 34 Constant Composition, 3, 5
spectrum, 47 Dulong and Petit, 13
Hydrogen fluoride, 99 Equilibrium, 194
Hydrogen sulphide, Gay-Lussac, 7
precipitations by, 238 Graham, 32, 115
Hydrolysis, of salts, 223 Henry, 165
degree of, 228 Hess, 183
Hydroxonium ion, 75 Isomorphism, 14
312 Physical Chemistry
Law of,—(continued) NAPHTHALENE, structure of, 96
Kohlrausch, 281 Negative catalysis, 267
Multiple Proportions, 5 Nernst, 251
Octaves, 58 Neutralization, 219
Radioactive Change, 40 Neutron, 23
Reciprocal Proportions, 6 Newlands, 58
Lattice, 86 Nitrate ion, bonding in, 85
covalent, 94 Nuclear, energy, 40
energy, 189 fission, 41
giant, 86 fusion, 42
ionic, 89 reactions, 38
metal, 86 reactors, 42
molecular, 95 theory of the atoms, 24
Lavoisier, 216 Nucleus, 25, 29
Le Chatelier’s Principle, 204, 207
Ligands, 76 ONE-COMPONENT SYSTEMS, 131
Line spectra, 46 Orbital, 55, 80
Liquefactions of gases, 121 Order of reaction, 257
Lothar Meyer, 61 Osmosis, 142
Lowering of vapour pressures, 133 Osmotic pressure, 142
Lyman series, 47 Laws, 145
Lyophilic sols, 108 Ostwald, 208
Lyophobic sols, 108 Oxidation, 243
number, 246
MAGNETIC QUANTUM NUMBER, 54 Oxidizing agent, 244 ,
Marsden, 24
Mass, defect, 36 Pp ORBITALS, 56
number, 30 Partial pressure, 166
spectrometer, 14, 30 Partition Law, 173
Mendeléeff, 59 Pauli, 55
Mercury(II) iodide, 242 Peptiziation, 106
Metallic bonding, 86 Periodic classification, 58
Methyl cyanide, bonding in, 72 anomalies of atomic mass, 59
Millikan, 22 Pfeffer, 144
Mitscherlich, 14 pH, 213
Molar conductivity, 278 Phase equilibria, 131
Molarity, 18 Phosphorus(V) chloride, bonding in,
Mole, 17 78
fraction, 114 Phosphorus(IL1) chloride, bonding in,
Molecular mass, relative, 78
by boiling point elevation, 135 Pi bond, 81
by freezing point depression, 138 eG, PN
by osmotic pressure, 148 pKo, 221
by vapour pressure lowering, 133 Planck’s constant, 47
of gases, 125 Polar bonds, 97
Molecular orbital, 80 Positive rays, 22
velocities, 118 Principal quantum number, 49, 54
Molecularity, 260 Proton, 22
Moseley, 26 acceptor, 219
Index 313
Proton,—(continued) Solvent extraction, 174
donor, 219 Spectra, hydrogen, 47
Pyknometer, 167 Spin quantum number, 54
Standard, electrode potential, 249
QUANTUM NUMBERS, 54 enthalpy change, 178
redox potential, 250
RADIOACTIVITY, 37 Steam distillation, 172
decay series, 39 Subsidiary quantum number, 50
Radiography, 45 Sulphate ion, bonding in, 85
Raoult’s Law, 133 Supersaturated solution, 152
Rast, 140
Rates of reaction, 254 THERMOCHEMICAL DEFINITIONS, 180
Reaction mechanism, 260 Thermodynamics, First Law of, 183
Redox, 243 Thixotropy, 112
potentials, 248 Thomson, J. J., 21
Reducing agent, 244 Thorium decay series, 39
Reduction, 243 Titration curves, 232
Relative atomic mass, 7, 10, 15 Tracer techniques, 44
Rutherford, 23 Two-component systems, 132

SALT HYDRATES, 161 ULTRA-MICROSCOPE, 109


classification of, 219 Uranium-235, 32
hydrolysis, 223
degree of, 228
VALENCY, 12
Saturated solution, 151
Vapour density, relative, 11, 126
Second-order reactions, 259
Vapour pressure, lowering, 133
Semi-permeable membrane, 143
over salt hydrates, 161
Sigma bond, 81 van der Waals’, equation, 120
Silver chloride,
forces, 103
soluble in ammonia, 241
radii, 103
Silver cyanide, Victor Meyer, 126
soluble in KCN(aq), 241
Sodium chloride, structure of, 91
Water, bonding in, 71
Sol, 105
ionic state of, 211
lyophilic, 108
shape, 79
lyophobic, 108
Wilson, 25
preparation of, 106
properties of, 108
Solubility, curves, 154 X-RAYS, 22
of gases, 163 diffraction, 64
of solids, 151 spectra, 26
product, 235
very low, 155 ZINC BLENDE, structure of, 93
e Ra e
Hes Mattes ey 2
ave i?

iroege eata: ay
Lotek e: =o, PKA
[ite - awe oe a...Stim
y. : i ore 6

Sh Ait Men ‘ ihat) &


; mi
wis $eb .coxirrertinth
a Shop ‘ i te i‘
tod

rod on | . .
7
9 ter OU SOG OLS | . 1
. ~ ae : .
F i “jun, ¥ se
. ’ ‘ =e
; m4
i”
=v ar ' ;
a =\ Oc
a rae
i iuiint
Level ‘hace |

TEXTBOOKS
Advanced Level Inorganic Chemistry
A. Holderness

Principles of Physical Chemistry


David Mansfield

Principles of Organic Chemistry


Peter R.S. Murray

Organic Chemistry: A Conceptual Approach


G.H. Williams

Structural and Comparative Inorganic Chemistry


P.R.S. Murray and P.R. Dawson

QUESTION BOOKS

Problems and Worked Examples in eer


to Advanced Level
A. Holderness and J. Lambert

Structured Questions in ‘A’ Level Chemistry


R.A. H. Hillman, M. C. V. Cane, and C. T. McCarty

EXPERIMENTAL TESTS
Practical Chemistry
J. Lambert and T.A. Muir

Comprehensive Qualitative Analysis for Advanced


Level Chemistry
E.N. Lambert and M. J. Mohammed

Heinemann Experimental Chemistry Series


General Editors: A.J. Mee and M. Rogers

Heinemann Educational Books

You might also like